Skip to main content

Full text of "Elements Of Fractional Distillation"

See other formats


>;  CD 

8]<OU  158477 


Osmania  University  Librarq 

Call  No. 


Author 
Title 


JLLsgLttl- 

Thi«  book  should  be  retufned'on V  beYore  the  date  last  marked  below 


Elements  of 
Fractional  Distillation 


BY 

CLARK  SHOVE  ROBINSON 

AND 

EDWIN   RICHARD  GLLLILAND 

Revised  and  Rewritten 

t>y 

EDWIN    RICHARD  GILT.II.ANO 

Professor  of  Chemical  Engineering 
Afassachusetts  Institute  of  Technology 


FOURTH  EDITION 
SKCOND 


McGRAW-HILL  BOOK   COMPANY,  INC. 

NKW  YORK      TORONTO      LONDON 
1950 


ELEMENTS  OF  FRACTIONAL  DISTILLATION 

Copyright,  1922,  1930,  1939,  1950,  by  the  McGraw-Hill  Book  Company,  Inc 
Printed  in  the  United  States  of  America.  All  rights  reserved.  This  book,  or  parti 
thereof,  may  not  be  reproduced  in  any  form  without  permission  of  the  publishers 


THE  MAPLE  PRESS  COMPANY,  YORK,  PA. 


PREFACE  TO  THE  FOURTH  EDITION 

The  firs^^ditipn.of-this  book  and  the  early  revisions  were  the  result 
of  the  efforts  of  Professor  Robinson,  and  he  took  an  active  part  in 
guiding  the  revision  of  the  previous  edition.  His  death  t  made  it 
necessary  to  prepare  this  edition  without  his  helpful  guidance  and 
counsel. 

The  present  revision  differs  extensively  from  the  previous  edition. 
The  material  has  been  modified  to  bring  it  more  closely  into  line  with 
the  graduate  instruction  in  distillation  at  Massachusetts  Institute  of 
Technology.  Much  greater  emphasis  has  been  placed  on  the  measure- 
ment, prediction,  and  use  of  vapor-liquid  equilibria  because  it  is 
believed  that  this  is  one  of  the  most  serious  limitations  in  design  calcu- 
lations. Greater  emphasis  has  also  been  placed  upon  the  use  of 
enthalpy  balances,  and  the  treatment  of  batch  distillation  has  been 
considerably  expanded.  Unfortunately,  the  design  calculations  for 
this  type  of  operation  are  still  in  an  unsatisfactory  status.  Azeotropic 
and  extractive  distillation  are  considered  as  an  extension  of  conven- 
tional multicomponent  problems.  The  sections  on  column  design 
and  column  performance  have  been  completely  rewritten  and  increased 
in  scope.  In  all  cases  quantitative  examples  have  been  given  because 
it  has  been  found  that  this  greatly  aids  the  student  in  understanding 
descriptive  material. 

During  the  last  15  years  a  large  number  of  design  methods  have  been 
proposed  for  multicomponent  mixtures,  some  of  which  are  reviewed  in 
Chapter  12.  Most  of  these  do  not  appear  to  offer  any  great  advantage 
over  the  conventional  Sorel  method,  and  it  is  believed  that  the  law  of 
diminishing  returns  has  been  applying  in  this  field  for  some  time.  It  is 
hoped  that  the  present  edition  will  stimulate  some  of  these  investi- 
gators to  transfer  their  efforts  to  more  critical  problems,  such  as 
vapor-liquid  equilibria,  batch  distillation,  transient  conditions  within 
the  distillation  system,  and  column  performance. 

EDWIN  RICHARD  GILLILAND 
CAMBRIDGE,  MASS. 
July,  1960 


PREFACE  TO  THE  FIRST  EDITION 

The  subject  of  fractional  distillation  has  received  but  scant  atten- 
tion from,  writers  in  the  English  language  since  Sidney  Young  published 
his  book  " Fractional  Distillation "  in  1903  (London).  French  and 
German  authors  have,  on  the  other  hand,  produced  a  number  of  books 
on  the  subject,  among  the  more  important  of  which  are  the  following: 

"La  Rectification  et  les  colonnes  rectificatriccs  en  distillerie," 
E.  Barbet,  Paris,  1890;  2d  ed.,  1895. 

"Der  Wirkungsweise  der  Rcctificir — und  Destillir— Apparate," 
E.  Hausbrand,  Berlin,  1893;  3d  ed.,  1910. 

"Theorie  der  Verdampfimg  und  Verfliissung  von  gemischcn  und  der 
fraktionierten  Destination,"  J.  P.  Kuenen,  Leipzig,  1906. 

"Theorie  der  Gewinnung  und  Trennung  der  atherischen  Olc  durch 
Destination,"  C.  von  Rechenberg,  Leipzig,  1910. 

"La  Distillation  fractione*e  et  la  rectification,"  Charles  Manlier, 
Paris,  1917. 

Young's  "Fractional  Distillation,"  although  a  model  for  its  kind, 
has  to  do  almost  entirely  with  the  aspects  of  the  subject  as  viewed 
from  the  chemical  laboratory,  and  there  has  been  literally  no  work  in 
English  available  for  the  engineer  and  plant  operator  dealing  with  the 
applications  of  the  laboratory  processes  to  the  plant. 

The  use  of  the  modern  types  of  distilling  equipment  is  growing  at 
a  very  rapid  rate.  Manufacturers  of  chemicals  are  learning  that  they 
must  refine  their  products  in  order  to  market  them  successfully,  and 
it  is  often  true  that  fractional  distillation  offers  the  most  available 
if  not  the  only  way  of  accomplishing  this.  There  has  consequently 
arisen  a  wide  demand  among  engineers  and  operators  for  a  book  which 
will  explain  the  principles  involved  in  such  a  way  that  these  principles 
can  be  applied  to  the  particular  problem  at  hand. 

It  has  therefore  been  the  purpose  of  the  writer  of  this  book  to 
attempt  to  explain  simply  yet  accurately,  according  to  the  best  ideas  of 
physical  chemistry  and  chemical  engineering,  the  principles  of  frac- 
tional distillation,  illustrating  these  principles  with  a  few  carefully 
selected  illustrations.  This  book  is  to  be  regarded  neither  as  a  com- 
plete treatise  nor  as  an  encyclopedia  on  the  subject  but,  as  the  title 
indicates,  as  an  introduction  to  its  study. 


Viii  PREFACE  TO  THE  FIRST  EDITION 

In  general,  it  has  been  divided  into  five  parts.  The  first  part  deals 
with  fractional  distillation  from  the  qualitative  standpoint  of  the 
phase  rule.  The  second  part  discusses  some  of  the  quantitative 
aspects  from  the  standpoint  of  the  chemical  engineer.  Part  three 
discusses  the  factors  involved  in  the  design  of  distilling  equipment. 
Part  four  gives  a  few  examples  of  modern  apparatus,  while  the  last 
portion  includes  a  number  of  useful  reference  tables  which  have  been 
compiled  from  sources  mostly  out  of  print  and  unavailable  except  in 
large  libraries. 

The  writer  has  drawn  at  will  on  the  several  books  mentioned  above, 
some  of  the  tables  being  taken  nearly  bodily  from  them,  and  has  also 
derived  much  help  from  Findlay's  " Phase  Rule"  (London,  1920)  and 
from  "The  General  Principles  of  Chemistry"  by  Noyes  and  Sherrill 
(Boston,  1917).  He  wishes  especially  to  express  his  gratitude  for 
the  inspiration  and  helpful  suggestions  from  Dr.  W.  K.  Lewis  of  the 
Massachusetts  Institute  of  Technology  and  from  his  other  friends  and 
associates  at  the  Institute  and  of  the  E.  B.  Badger  &  Sons  Company. 
Finally,  he  wishes  to  express  his  appreciation  of  the  assistance  of 
Miss  Mildred  B.  McDonald,  without  which  this  book  would  never 
have  been  written. 

CLARK  SHOVE  ROBINSON 

CAMBRIDGE,  MASS. 
June,  1920, 


CONTENTS 

PREFACE  TO  THE  FOURTH  EDITION  .  v 

PREFACE  TO  THE  FIRST  EDITION  vii 

INTRODUCTION 1 

1.  Determination  of  Vapor-Liquid  Equilibria.  .        .    .        .3 

2.  Presentation  of  Vapor-Liquid  Equilibrium  Data 16 

3.  Calculation  of  Vapor-Liquid  Equilibria   .  .  .26 

4.  Calculation  of  Vapor-Liquid  Equilibria  (Continued)      .    .       .79 

5.  General  Methods  of  Fractionation  .  .    .    101 

6.  Simple  Distillation  and  Condensation.  107 

7.  Rectification  of  Binary  Mixtures.    .  118 

8.  Special  Binary  Mixtures .  192 

/9/  Rectification  of  Multicomponent  Mixtures     .  ...        ,   214 

10.  Extractive  and  Azeotropic  Distillation  ...        .   285 

i^l.  Rectification  of  Complex  Mixtures  .  ....   325 

12.  Alternate  Design  Methods  for  Multicomptonent  Mixtures      ,  336 

13.  Simultaneous  Rectification  and  Chemical  Reaction.    .  .  361 

14.  Batch  Distillation 370 

15.  Vacuum  Distillation 393 

16.  Fractionating  Column  Design  403 

17.  Fractionating  Column  Performance 445 

18.  Fractionating  Column  Auxiliaries 471 

APPENDIX 479 

INDEX 481 

ix 


INTRODUCTION 

Definition  of  Fractional  Distillation.  By  the  expression  fractional 
distillation  was  originally  meant  the  process  of  separating  so  far  as  it 
may  be  feasible  a  mixture  of  two  or  more  volatile  substances  into  its 
components,  by  causing  the  mixture  to  vaporize  by  suitable  application 
of  heat,  condensing  the  vapors  in  such  a  way  that  fractions  of  varying 
boiling  points  are  obtained,  re  vaporizing  these  fractions  and  separating 
their  vapors  into  similar  fractions,  combining  fractions  of  similar  boil- 
ing points,  and  repeating  until  the  desired  degree  of  separation  is  finally 
obtained. 

Purpose  of  Book.  Such  a  process  is  still  occasionally  met  with  in 
the  chemical  laboratory,  but  it  is  a  laborious  and  time-consuming 
operation  which  has  its  chief  value  as  a  problem  for  the  student,  for  the 
purpose  of  familiarizing  him  with  some  of  the  characteristic  properties 
of  volatile  substances.  It  is  possible  to  carry  on  a  fractional  distilla- 
tion by  means  of  certain  mechanical  devices  which  eliminate  almost  all 
of  this  labor  and  time  and  which  permit  separations  not  only  equal  to 
those  obtained  by  this  more  tedious  process  but  far  surpassing  it  in 
quality  and  purity  of  product.  The  purpose  of  this  book  is  to  indicate 
how  such  devices  may  be  profitably  used  in  the  solution  of  distillation 
problems. 

Origin  of  Fractional  Distillation.  Like  all  the  older  industries, 
fractional  distillation  is  an  art  that  originated  in  past  ages  and  that 
developed,  as  did  all  the  arts,  by  the  gradual  accumulation  of  empirical 
knowledge.  It  is  probable  that  its  growth  took  place  along  with  that 
of  the  distilled  alcoholic  beverages,  and  to  the  average  person  today 
the  word  "still"  is  synonymous  with  apparatus  for  making  rum, 
brandy,  and  other  distilled  liquors.  To  France,  which  has  been  the 
great  producer  of  brandy,  belongs  the  credit  for  the  initial  development 
of  the  modern  fractionating  still. 

Physical  Chemistry  and  Fractional  Distillation.  Fractional  dis- 
tillation has  labored  under  the  same  sort  of  burden  that  the  other 
industrial  arts  have  borne.  Empirical  knowledge  will  carry  an 
industry  to  a  certain  point,  and  then  further  advances  are  few  and  far 
between.  It  has  been  the  function  of  the  sciences  to  come  to  the  rescue 

1 


2  FRACTIONAL  DISTILLATION 

of  the  arts  at  such  times  and  thus  permit  advancement  to  greater  use- 
fulness. The  science  that  has  raised  fractional  distillation  from  an 
empirical  to  a  theoretical  basis  is  physical  chemistry.  By  its  aid  the 
study  of  fractionation  problems  becomes  relatively  simple,  and  it  is  on 
this  account  that  the  subject  matter  in  this  book  is  based  upon  physical 
chemistry. 


CHAPTER  1 
DETERMINATION  OF  VAPOR-LIQUID  EQUILIBRIA 

The  separation  of  a  mixture  of  volatile  liquids  by  means  of  fractional 
distillation  is  possible  when  the  composition  of  the  vapor  coming  from 
the  liquid  mixture  is  different  from  that  of  the  liquid.  The  separation 
is  the  easier  the  greater  the  difference  between  the  composition  of  the 
vapor  and  that  of  the  liquid,  but  separation  may  be  practicable  even 
when  the  difference  is  small.  The  relation  between  the  vapor  and 
liquid  compositions  must  be  known  in  order  to  compute  fractional  dis- 
tillation relationships.  Usually  this  is  obtained  from  information 
concerning  the  composition  of  the  vapor  which  is  in  equilibrium  with 
the  liquid.  On  this  account  a  knowledge  of  vapor-liquid  equilibrium 
compositions  is  usually  essential  for  the  quantitative  design  of  frac- 
tional distillation  apparatus.  In  most  cases  the  study  is  made  on  the 
basis  of  the  composition  of  the  vapor  in  equilibrium  with  the  liquid. 
However,  this  is  not  a  fundamental  requirement  and  any  method  that 
would  allow  the  production  of  a  vapor  of  a  different  composition  than 
that  of  the  condensed  phase,  whether  equilibrium  or  not,  could  be  used 
for  separation.  However,  most  of  the  equipment  employed  depends 
on  the  use  of  a  vaporization  type  of  operation,  and  the  equilibrium 
vapor  is  a  good  criterion  of  the  possibilities  of  obtaining  a  separation. 

The  methods  for  obtaining  vapor-liquid  equilibrium  compositions 
can  be  considered  in  two  main  classifications:  (1)  the  experimental 
determination  of  equilibrium  compositions  and  (2)  the  theoretical 
relationships. 

EXPERIMENTAL    DETERMINATIONS    OF    VAPOR-LIQUID    EQUILIBRIA 

The  measurement  of  vapor-liquid  equilibrium  compositions  is  not 
simple.  A  highly  developed  laboratory  technique  is  therefore  needed 
to  obtain  reliable  data  in  any  of  the  several  methods  described  here. 

Circulation  Method.  A  common  method  for  obtaining  vapor-liquid 
equilibrium  (Refs.  11,  13,  16,  23,  27,  35)  is  by  circulating  the  vapor 
through  a  system  and  bringing  it  into  repeated  contact  with  the 
liquid  until  no  further  change  in  the  composition  of  either  takes 
place.  A  schematic  diagram  of  such  a  system  is  shown  in  Fig.  1-1. 
The  vapor  above  the  liquid  in  vessel  A  is  removed,  passed  through 

3 


FRACTIONAL  DISTILLATION 


chamber  5,  and  recirculated  by  pump  C  through  the  liquid  in  A. 
While  the  system  appears  simple,  in  actual  practice  it  involves  a  num- 
ber of  complications: 

1.  The  system  must  be  completely  tigkt;  otherwise  the  total  quan- 
tity of  material  will  continually  vary  and  the  equilibrium  compositions 
of  the  vapor  and  liquid  will  also  change. 

2.  The  quantities  of  liquid  and  of  vapor  when  equilibrium  is  obtained 
must  remain  constant  and  not  vary  during  the  recirculation.     To  keep 


Heat-* 
exchanger 

Equilibrium       I 
cell 


--Vapor  sample 


FIG.  1-1.     Circulation  apparatus  for  vapor-liquid  equilibrium  measurements. 

them  constant  it  is  necessary  for  the  system  to  remain  isothermal  and 
for  the  total  volume  to  remain  constant.  The  chief  variation  in  the 
volume  of  the  system  is  due  to  the  fact  that  it  is  usually  found  expedient 
to  use  a  reciprocating  pump.  The  error  due  to  this  variation  is  usually 
minimized  by  making  the  displacement  volume  of  the  pump  small. 
The  pumps  are  generally  of  a  mercury-piston  type;  t.e.,  a  mercury 
column  is  forced  up  and  down  in  a  steel  or  glass  cylinder  serving 
as  the  piston  of  the  pump.  This  makes  it  possible  to  have  an  essen- 
tially leakproof  pump  and  allows  the  pumping  operation  to  be  carried 
out  with  very  little  contamination  of  the  circulating  vapors. 
3.  This  type  of  system  has  been  used  most  successfully  under  con- 


DETERMINATION  OF  VAPOR-LIQUID  EQUILIBRIA  5 

ditions  where  the  vapor  does  not  condense  at  room  temperature.  If 
it  were  necessary  to  operate  the  pumping  system  at  a  high  temperature 
to  avoid  condensation  of  the  vapor,  difficulties  might  be  encountered 
due  to  the  vapor  pressure  of  the  mercury,  in  which  case  other  lower 
vapor  pressure  metallic  liquids  should  be  suitable. 

4.  Another  condition  that  could  cause  the  relative  volumes  of  vapor 
and  of  liquid  to  vary  is  the  rate  of  flow.     The  rate  of  recirculation 
varies  the  pressure  drop  through  the  apparatus  and  thereby  changes 
the  quantity, of  vapor  present.     In  most  cases  the  rate  of  recirculation 
is  such  that  the  pressure  differential  for  recirculation  is  not  great. 
Both  the  volume  variation  due  to  pumping  and  the  pressure  changes 
due  to  recirculation  can  be  made  less  detrimental  by  making  the  vol- 
ume of  the  liquid  in  vessel  A  large. 

5.  It  is  necessary  to  ensure  that  there  is  no  entrainment  of  liquid 
with  the  vapor  leaving  A.     If  liquid  is  carried  over  to  vessel  B,  the 
vapor  sample  will  be  contaminated.     This  entrainment  is  eliminated 
by  the  use  of  low  velocity  and  by  efficient  entrainment  separators  in 
the  upper  part  of  A. 

6.  Another  precaution  is  the  necessity  to  prevent  any  condensation 
of  the  vapor  during  recirculation.     If  any  vapor  condenses,  the  con- 
densate  will  be  of  different  composition  and  the  results  will  be  in  error. 

This  type  of  apparatus  has  been  used  for  a  variety  of  systems. 
It  is  particularly  suitable  for  very  low  temperature  studies  such  as 
those  involved  in  the  equilibria  associated  with  liquid  air.  In  this 
case  vessel  A  is  maintained  in  a  low-temperature  cryostat,  and  the 
recycle  vapor  stream  is  heat-exchanged  with  the  exit  vapor;  the  rest 
of  the  system  is  maintained  at  essentially  room  temperature.  One  of 
the  difficulties  with  the  operation  is  the  fact  that  the  vapor  sample  is 
obtained  as  a  vapor  and,  unless  the  pressure  is  high,  the  quantity  of 
vapor  obtained  in  vessel  B  may  be  so  small  as  to  offer  difficulties  in 
analysis. 

The  system  has  the  great 'ad  vantage  that  a  vapor  can  be  repeatedly 
bubbled  through  the  liquid  until  equilibrium  is  obtained.  Theoretically 
exact  equilibrium  is  not  obtained  because  of  the  fact  that  there  are 
pressure  differentials  in  the  system.  Thus  the  vapor  entering  at  the 
bottom  of  A  must  be  under  a  pressure  higher  than  the  vapor  leaving  A, 
at  least  by  an  amount  equal  to  the  hydrostatic  head  of  the  liquid  in  A. 
Since  the  vapor-liquid  equilibria  depend  on  pressure,  it  is  obvious  that 
there  cannot  be  exact  equilibrium.  However,  the  change  in  the 
composition  of  the  equilibrium  vapor  due  to  this  small  change  of  pres- 
sure is  small  in  most  cases.  It  could  be  serious  in  the  critical  region 


FRACTIONAL  DISTILLATION 


where  the  vapor  is  very  compressible.  Basically  this  system  is  one  of 
the  best  for  obtaining  true  equilibrium. 

Bomb  Method.  In  the  bomb  method  (Refs.  3,  4,  12,  14,  36)  the 
liquid  sample  is  placed  in  a  closed  evacuated  vessel.  It  is  then  agitated 
by  rocking,  or  by  other  means,  at  constant  temperature  until  equilib- 
rium is  obtained  between  the  vapor  and  the  liquid.  Samples  of  the 
vapor  and  the  liquid  are  then  withdrawn  and  analyzed. 

The  method  appears  simple,  but  it  involves  certain  difficulties. 
During  sampling  there  are  pressure  changes  due  to  the  removal  of 
material,  and  these  pressure  changes  can  be  large  in  magnitude.  In 
order  to  avoid  them,  it  is  customary  to  add  some  fluid,  such  as  mercury, 
to  the  system  while  the  samples  are  being  removed  in  order  to  prevent 


Rocking 
mechanism 


-  Sampling  line 


fjf  -  Constant  temperature 


FIG.  1-2.     Bomb  apparatus. 

any  vaporization  or  condensation.  Another  difficulty  with  the 
method  is  the  fact  that  in  most  cases  it  is  necessary  to  use  sampling 
lines  of  small  cross  sections.  These  may  fill  up  with  liquid  during  the 
initial  part  of  the  operation,  and  this  liquid  may  never  come  to  the  true 
equilibrium.  It  is  necessary  to  purge  the  sampling  lines  to  remove 
such  liquid.  This  liquid  holdup  is  particularly  serious  in  the  case  of 
the  vapor  sample  since  in  quantity  it  may  be  large  in  comparison  to  the 
sample.  A  schematic  diagram  of  the  bomb-type  apparatus  is  shown 
in  Fig.  1-2. 

Dynamic  Flow  Method.  Another  method  that  has  been  widely 
used  (Refs.  10,  21,  25,  37)  for  the  determination  of  vapor-liquid  equi- 
libria is  one  in  which  a  vapor  is  passed  through  a  series  of  vessels  con- 
taining liquids  of  a  suitable  composition.  The  vapor  entering  the 
first  vessel  may  be  of  a  composition  somewhat  different  from  the 
equilibrium  vapor,  but  as  it  passes  through  the  system  it  tends  to 
approach  equilibrium,  If  all  the  vessels  have  approximately  the  same 
liquid  composition,  the  vapor  will  more  nearly  approach  equilibrium 
as  it  passes  through  the  unit,  The  number  of  vessels  employed  should 
be  suoh  that  the  vapor  entering  the  last  unit  is  of  essentially  equilib- 
rium composition. 


DETERMINATION  OF  VAPOR-LIQUID  EQUILIBRIA  7 

This  system  has  the  advantage  that  it  is  simple  and,  in  certain  cases, 
it  is  possible  to  dispense  with  the  analysis  of  the  liquid  sample,  i.e.,  the 
liquids  can  be  made  of  a  known  composition,  and  since  the  change  in 
the  last  vessel  is  small,  it  is  possible  to  assume  that  the  composition  of 
the  liquid  in  this  case  is  equal  to  that  originally  charged.  A  schematic 
diagram  of  such  a  system  is  shown  in  Fig.  1-3. 

It  is  obvious  that  it  cannot  be  an  exact  equilibrium  system  because 
of  the  fact  that  a  pressure  drop  is  involved  in  passing  the  vapor  through 
the  system;  i.e.,  there  are  pressure  variations  which  will  affect  equi- 
librium. There  is  also  the  danger  of  entrainment,  although  this  can 
be  minimized  by  low  velocities. 

In  a  great  many  cases,  the  gas  introduced  into  the  first  vessel  has 
.been  carrier  gas  of  low  solubility  and  not  a  component  of  the  system. 


FIG.  1-3.     Dynamic  flow  method. 

Thus,  in  the  determination  of  the  vapor-liquid  equilibria  for  systems 
such  as  ammonia  and  water,  ammoniacal  solutions  are  placed  in  the 
vessels,  and  a  gas  such  as  nitrogen  is  bubbled  into  the  first  of  these  and 
the  resulting  nitrogen-ammonia-water  vapor  mixture  is  passed  through 
the  succeeding  vessels  obtaining  a  closer  approach  to  equilibrium. 
Equilibrium  obtained  in  such  a  manner  is  not  the  true  vapor-liquid 
equilibria  for  the  water  vapor-ammonia  system.  It  closely  approaches 
true  equilibrium  for  the  binary  system  under  a  total  pressure  equal  to 
the  partial  pressure  of  the  ammonia  and  the  water  vapor  in  the  gaseous 
mixture.  Even  this  is  not  exact.  The  carrier  gas  has  some  solubility 
in  the  liquid  phase,  and  the  partial  pressure  of  these  added  constituents 
modifies  the  energy  relations  of  the  liquid  and  vapor  phases.  Usually 
for  low-pressure  operation  these  errors  are  not  large  in  magnitude,  but 
as  the  pressure  becomes  higher  the  errors  are  serious  and  the  method 
can  give  erroneous  results  if  the  true  vapor-liquid  equilibria  for  mix- 
tures without  the  carrier  gas  are  desired. 

Dew  and  Boiling-point  Method.  In  essence  this  technique  consists 
in  introducing  a  mixture  of  known  composition  into  an  evacuated 
equilibrium  container  of  variable  volume  (Refs.  6,  7,  9,  15,  17,  18,  20, 
28).  The  system  is  maintained  at  a  constant  temperature,  and  by 
varying  the  volume  the  pressure  is  observed  at  which  condensation 


8  FRACTIONAL  DISTILLATION 

commences  and  is  completed.  The  dew-  and  bubble-point  curves  of 
pressure  vs.  temperature  for  a  number  of  these  prepared  samples  are 
obtained  and,  by  cross-plotting,  conditions  of  phase  equilibrium  may 
be  found  by  locating  points  at  which  saturated  liquid  and  saturated 
vapor  of  different  compositions  exist  at  the  same  temperature  and 
pressure. 

The  pressures  are  determined  in  two  ways.  One  involves  the 
measuring  and  plotting  of  the  PV  isotherm,  the  dew  point  and  bubble 
point  being  indicated  by  the  discontinuities  in  the  curve  at  the  begin- 
ning and  the  end  of  condensation.  The  other  employs  a  glass  or 
quartz  equilibrium  cell,  and  the  conditions  are  determined  visually. 

The  advantages  of  this  method  are  that  it  allows  the  critical  condi- 
tions to  be  determined,  gives  data  on  specific  volumes  of  mixtures  at 
high  pressures,  and  requires  no  analysis  of  the  phases  since  the  total 
composition  of  the  mixture  is  accurately  determined  gravimetrically 
upon  charging. 

There  are  disadvantages,  however.  For  certain  conditions  the  dew 
and  bubble  points  are  not  sharply  defined ;  hence  they  require  measure- 
ments to  be  made  with  highly  refined  precision  instruments.  The 
simpler  units  using  mercury  as  a  variable  volume  confining  fluid  cannot 
be  used  below  the  freezing  point  of  mercury.  In  addition,  the  mate- 
rials used  must  be  very  pure  and  free  in  particular  from  traces  of  fixed 
gases,  for  in  the  critical  region  the  saturation  pressure  is  quite  sensitive 
to  small  amounts  of  fixed  gases.  Further,  a  large  amount  of  experi- 
mental work  must  be  done  in  order  to  define  completely  and  accurately 
the  phase  equilibria  over  all  ranges  of  liquid  and  vapor  composition. 
The  major  limitation,  however,  is  the  fact  that  the  method  can  be  used 
only  on  binary  systems.  As  the  phase  rule  dictates  that  more  complex 
systems  are  not  a  unique  function  of  pressure  and  temperature,  dew 
and  bubble  points  alone  cannot  define  the  composition  of  two  phases  in 
equilibrium. 

Dynamic  Distillation  Method.  The  four  previous  methods  involved 
repeated  contact  of  the  vapor  with  the  liquid  and  thus  afforded  the 
time  necessary  for  the  attainment  of  equilibrium.  The  dynamic  dis- 
tillation method  (Refs.  2,  5,  11,  19,  24,  26,  34,  39)  involves  a  different 
procedure  (see  Fig.  1-4).  In  this  system  a  distilling  vessel  is  connected 
to  a  condenser  and  a  receiver. 

In  the  simplest  case,  a  small  sample  of  distillate  is  taken,  and  the 
compositions  of  this  sample  and  the  liquid  in  the  still  are  determined. 
During  such  a  distillation  the  composition  of  the  distillate  and  the 
liquid  in  the  still  changes,  and  the  samples  represent  average  values. 


DETERMINATION  OF  VAPOR-LIQUID  EQUILIBRIA 


9 


To  reduce  this  composition  variation  the  quantity  of  liquid  in  the  still 
is  made  large  in  comparison  to  the  quantity  of  distillate.  Frequently 
successive  samples  of  condensate  are  obtained,  and  these  are  analyzed 
and  the  composition  plotted  vs.  the  quantity  of  liquid  that  has  been 
distilled.  An  extrapolation  of  this  curve  back  to  zero  quantity  of 
liquid  removed  is  taken  as  the  composition  of  the  vapor  in  equilibrium 
with  the  original  liquid. 

Top  heater 


Charge 


FIG.  1-4.     Dynamic  distillation  apparatus. 

The  method  involves  a  new  assumption,  namely,  that  the  vapor 
obtained  by  boiling  a  liquid  is  in  equilibrium  with  the  liquid.  There 
has  been  no  adequate  proof  of  this  assumption,  and  theoretical  con- 
siderations would  tend  to  indicate  that  equilibrium  should  not  be 
obtained.  The  few  experimental  data  that  are  available  would  indi- 
cate that  the  difference  in  the  composition  between  the  vapor  obtained 
in  this  manner  and  the  true  equilibrium  is  not  great  in  most  cases,  but 
in  a  few  systems  significant  differences  have  been  found. 

After  the  vapor  leaves  the  liquid,  any  condensation  in  the  upper  part 
of  the  distilling  vessel  will  change  the  composition  of  the  vapor  and 


10  FRACTIONAL  DISTILLATION 

therefore  introduce  errors.  Such  condensation  is  usually  reduced  or 
eliminated  by  having  the  upper  part  of  the  system  jacketed  and  at  a 
higher  temperature  than  the  condensation  temperature  of  the  vapor. 
However,  this  higher  temperature  can  introduce  errors;  for  example, 
in  such  a  boiling  system  there  is  a  certain  amount  of  spray  and  splash- 
ing. The  spray  that  lands  on  the  heated  walls  will  tend  to  vaporize 
totally  and  give  a  vapor  of  the  composition  of  the  liquid  rather  than  of 
the  equilibrium  composition. 

The  pressure  involved  in  such  a  system  is  of  course  essentially  that 
prevailing  in  the  receiver,  and  this  method  can  be  used  either  for  nor- 
mal pressures,  high  pressures,  or  vacuum.  The  exact  temperature  of 
the  operation  is  usually  not  known  because  the  liquid  is  generally 
superheated.  The  vapor  and  the  liquid  therefore  are  not  in  thermal 
equilibrium,  and  it  is  doubtful  whether  they  are  in  true  composition 
equilibrium.  The  apparatus  has  been  extensively  used  because  of  its 
simplicity,  and  the  results  are  of  sufficient  accuracy  to  be  of  real  value 
in  distillation  calculations. 

In  order  to  obtain  a  closer  approach  to  equilibrium,  various  com- 
plicating arrangements  have  been  used;  for  example,  Rosanoff  modified 
the  system  to  obtain  a  second  contact  of  the  vapor  with  the  liquid. 

Continuous  Distillation  Methods.  Continuous  distillation  methods 
involve  distilling  a  liquid,  condensing  the  vapor  sample,  and  recycling 
the  condensate  back  to  the  still.  A  schematic  drawing  of  such  an 
equilibrium  still  is  given  in  Fig.  1-5.  This  system  was  developed  by 
Yamaguchi  (Ref .  38)  and  Sameshima  (Ref .  29)  and  has  been  modified 
and  improved  by  a  number  of  other  investigators  (Refs.  1,  8,  22,  30,  31, 
32,  33).  This  method  has  been  widely  used  and  has  the  great  advan- 
tage that  it  is  simple,  and  the  unit  can  be  placed  in  operation  and 
allowed  to  come  to  a  steady  state  without  any  great  amount  of  atten- 
tion. The  same  precautions  relative  to  entrainment,  condensation 
and  total  vaporization  of  splashed  liquid  must  be  observed  in  the  still 
as  was  indicated  for  the  dynamic  distillation  method.  The  condensate 
collects  until  the  level  is  high  enough  to  flow  over  the  trap  and  back  to 
the  still.  At  the  end  of  the  distillation,  this  condensate  is  removed  and 
analyzed  to  determine  the  composition  of  the  vapor,  and  a  sample  is 
removed  from  the  still  to  determine  the  still  composition. 

This  method  suffers  from  the  same  difficulties  as  the  dynamic  dis- 
tillation method  in  that  it  is  open  to  the  question  of  whether  the  vapor 
formed  by  boiling  a  liquid  is  in  equilibrium  with  the  liquid.  It  is  also 
difficult  to  obtain  the  true  liquid  temperature  because  of  the  super- 
heating effects.  The  pressure  is  maintained  by  the  pressure  in  the  exit 


DETERMINATION  OF  VAPOR-LIQUID  EQUILIBRIA 


11 


tube,  and  in  normal  pressure  determinations  this  is  usually  open  to  the 
atmosphere.  This  theoretically  offers  the  possibility  of  errors  in  that 
it  allows  Oxygen  and  nitrogen  to  dissolve  in  the  condensate  sample, 
which  is  then  recycled  back  to  the  still.  At  low  pressures  the  solu- 
bility of  such  gases  is  usually  small  and  the  error  is  slight,  but  in  high- 
pressure  operations  the  use  of  this  gas  system  can  lead  to  serious  errors. 


FIG.  1-5.     Continuous  distillation  equilibrium  still. 

The  gas  pressuring  system,  however,  is  extremely  desirable  in  that  it 
regulates  the  condenser  cooling  capacity  so  that  it  exactly  balances 
heat  input  to  the  still.  At  high  pressures  the  errors  become  so  serious 
that  this  benefit  must  be  foregone.  Figure  1-6  indicates  a  type  of 
apparatus  in  which  the  heat  input  and  removal  are  adjusted  so  that 
the  pressure  remains  constant  without  the  necessity  of  a  sealing  gas. 
Another  source  of  error  in  the  system  is  possible  because  the  con- 
densate returned  to  the  still  is  of  a  different  composition  from  the 
liquid  in  the  still  and  in  general  is  of  lower  boiling  point.  If  this 
vaporizes  before  it  is  completely  mixed  with  all  of  the  liquid  in  the  still, 
this  vapor  composition  will  not  be  an  equilibrium  vapor. 


12 


FRACTIONAL  DISTILLATION 


Although  the  apparatus  appears  to  be  of  the  recirculation  type  and 
it  might  be  supposed  that  the  successive  contacts  would  tend  to  give 
a  closer  approach  to  equilibrium,  this  is  not  the  case.  If  the  vapor 
evolved  from  the  liquid  is  not  an  equilibrium  vapor,  this  type  of  recycle 


Vent*. 


Condenser, 
Top  heater 


Relay  controlling  heat 
supply  to  still 


SHU 
thermo  • 
couple 


St/it 
sample 


Hot 


Bottom  heater 
FIG.  1-6.     Continuous  distillation  still  for  high-pressure  operation. 

system  does  not  give  a  closer  approach  with  repeated  recycling  since 
new  vapor  is  formed  and  the  recycled  material  is  not  brought  to  equi- 
librium by  successive  contacts.  The  recycling  does  give  a  steady-state 
condition,  but  the  approach  to  equilibrium  is  only  that  obtained  by 
boiling  the  liquid. 

In  order  to  eliminate  some  of  the  sources  of  error  in  continuous  dis- 
tillation systems,  various  modifications  ^have  been  made.     The  most 


DETERMINATION  OF  VAPOR-LIQUID  EQUILIBRIA 


13 


important  of  these  (Refs.  1,  8)  would  appear  to  be  one  in  which  the 
condensate  stream  is  re  vaporized  before  it  is  returned  to  the  still;  i.e., 
the  heat  is  added  to  revaporize  the  condensate  stream  instead  of  form- 
ing a  new  vapor  in  the  still.  Such  an  apparatus  is  shown  in  Fig.  1-7, 
In  this  case,  the  result  is  equivalent  to  recycling  the  vapor,  and  the 
operation  tends  to  be  equivalent  to  the  usual  recycle  system.  It  is 
more  difficult  to  operate  than  the  conventional  continuous  distillation 
system.  The  condensate  must  be  completely  vaporized.  If  any 


Thermometer 
well 


FIG.  1-7.     Continuous  distillation  still  with  re  vaporized  condensate. 

liquid  is  allowed  to  return  to  the  still,  the  purpose  of  the  system  is 
defeated  and  the  rate  of  distillation  decreases;  i.e.,  less  vapor  leaves  the 
still.  If  the  vapor  returned  to  the  still  is  greatly  superheated,  it  will 
cause  additional  evaporation  in  the  still  and  the  operation  will  speed 
up.  By  proper  adjustment  a  satisfactory  balance  can  be  obtained. 
It  is  believed  that  this  apparatus  is  a  definite  improvement  over  the 
regular  continuous  distillation  system,  and  comparative  data  on  the 
same  system  taken  with  this  and  the  usual  continuous  distillation  sys- 
tem show  definite  differences  of  the  type  that  would  be  anticipated. 

Both  the  continuous  distillation  system  and  the  modifications  of  it 
suffer  from  the  difficulty  that  the  vapor  must  be  totally  condensable 
under  the  operating  conditions^  This  is  usually  not  a  serious  difficulty, 


14  FRACTIONAL  DISTILLATION 

It  is  also  necessary  that  the  condensate  be  a  homogeneous  mixture. 
,  Thus,  if  the  condensate  separates  into  two  layers,  the  operation  is  not 
satisfactory.  The  other  vapor-liquid  equilibrium  methods  are  suitable 
for  multilayer  systems  either  in  the  still  or  in  the  vapor  sample. 

ACCURACY  OF  VAPOR-LIQUID  EQUILIBRIUM  DATA 

A  satisfactory  investigation  of  the  accuracy  of  the  various  experi- 
mental methods  has  not  been  made,  and  there  is  real  question  con- 
cerning a  large  amount  of  the  published  experimental  vapor-liquid  equi- 
librium data.  The  circulation  and  the  bomb  methods  have  the 
potentiality  of  giving  high  accuracy,  and  the  value  of  the  results  depends 
on  the  care  employed  by  the  experimentalist  in  eliminating  sources  of 
error.  The  dynamic  distillation  and  the  continuous  distillation 
methods  involve  the  assumption  that  the  vapor  produced  by  boiling  a 
liquid  is  of  a  composition  that  is  in  equilibrium  with  the  remaining 
liquid.  The  adequacy  of  this  assumption  has  not  been  proved  experi- 
mentally, but  there  are  experimental  results  which  cast  doubt  on  its 
validity  for  all  cases.  Analysis  of  the  published  data  obtained  by 
employing  the  continuous  distillation  method  would  indicate  that  it 
may  give  values  for  the  differences  of  the  vapor  and  liquid  compositions 
at  a  given  liquid  composition  that  are  within  10  to  15  per  cent  of  the 
true  values.  When  results  for  the  same  system  are  compared  on  this 
basis,  it  is  not  uncommon  to  find  deviations  of  ±  10  per  cent  between 
different  investigators  employing  essentially  the  same  techniques.  A 
critical  study  of  the  methods  of  determining  vapor-liquid  equilibria  is 
needed. 

References 

1.  ADEY,  S.  M.  thesis  M.I.T.,  1941. 

2.  BALY,  Phil.  Mag.,  (V),  49,  517  (1900). 

3.  BEBGANTZ,  Sc.D.  thesis,  M.I.T.,  1941. 

4.  BOOMER  and  JOHNSON,  Can.  J.  Research,  16B,  328  (1938). 

5.  BROWN,  Trans.  Chem.  Soc.,  35,  547  (1879). 

6.  CALINGAERT  and  HITCHCOCK,  J.  Am.  Chem.  Soc.,  49,  750  (1927). 

7.  CAUDET,  Compt.  rend.,  130,  828  (1900). 

8.  COLBURN,  JONES,  and  SCHOENBORN,  Ind.  Eng.  Chem.,  35,  666  (1943). 

9.  CUMMINGS,  Ind.  Eng.  Chem.,  23,  900  (1931). 

10.  DOBSON,  J.  Chem.  Soc.,  127,  2866  (1925). 

11.  DODGE  and  DUNBAR,  J.  Am.  Chem.  Soc.,  49,  591  (1927), 

12.  FEDORITBNKO  and  RUHEMANN,  Tech.  Phys.,  U.S.S.R.,  4,  1  (1937). 

13.  FERGUSON  and  FUNNELL,  /.  Phys.  Chem.,  33,  1  (1929). 

14.  FREETH  and  VERSCHOYLB,  Proc.  Royal  Soc.  (London),  A  130,  453  (1931). 

15.  HOLST  and  HAMBURG,  Z.  physik.  Chem.,  91,  513  (1919). 

16.  INGLIS,  Phil.  Mag.  (VI),  11,  640  (1906). 


DETERMINATION  OF  VAPOR-LIQUID  EQUILIBRIA  15 

17.  KAY,  Ind.  Eng.  Chem.,  30,  459  (1938), 

18.  KTJBNEN,  VERSCHOYLE,  and  VAN  UBK,  Communs.  Phys.  Lab.,  Univ.  Leiden, 
No.  161  (1922). 

19.  LEHFELDT,  Phil.  Mag.  (V),  46,  42  (1898). 

20.  MEYEK,  Z.  physik.  Chem.,  A  175,  275  (1936). 

21.  ORDOBFF  and  CARRELL,  /.  Phys.  Chem.,  1,  753  (1897). 

22.  OTHMER,  Ind.  Eng.  Chem.,  20,  743  (1928). 

23.  QUINN,  Sc.D.  thesis,-  M.I.T.,  1940. 

24.  RAYLEIGH,  Phil.  Mag.  (VI),  4,  521  (1902). 

25.  RIGNAULT,  Ann.  chim.  et  phys.,  (3)  16,  129  (1845). 

26.  ROSANOFF,  BACON,  and  WHITE,  J.  Am.  Chem.  Soc.,  36,  1993  (1914). 

27.  ROSANOFF,  LAKE,  and  BREITHUT,  /.  Am.  Chem.  Soc.,  48,  2055  (1909). 

28.  SAGE  and  LACEY,  Ind.  Eng.  Chem.,  26,  103  (1934). 

29.  SAMESHIMA,  J.  Am.  Chem.  Soc.,  40,  1482,  1503  (1918). 

30.  SCATCHARD,  RAYMOND,  and  GILMANN,  J.  Am.  Chem.  Soc.,  60,  1275  (1938). 

31.  SCHEELINE  and  GILLILAND,  Ind.  Eng.  Chem.,  31,  1050  (1939). 

32.  SIMS,  Sc.D.  thesis,  M.I.T.,  1933. 

33.  SMYTH  and  ENGEL,  /.  Am.  Chem.  Soc.,  61,  2646  (1929). 

34.  TAYLOR,  /.  Phys.  Chem.,  4,  290  (1900). 

35.  TOROCHESNIKOV,  Tech.  Phys.,  U.S.S.R.,  4,  337  (1937). 

36.  VERSCHOYLE,  Trans.  Roy.  Soc.  (London),  A  230,  189  (1931). 

37.  WILL  and  BREDIG,  Ber.,  22,  1084  (1889). 

38.  YAMAGUCHI,  /.  Tokyo  Chem.  Soc.,  34,  691  (1913). 

39.  ZAWIDSKI,  Z.  physik.  Chem.,  36,  129  (1900). 


CHAPTER  2 
PRESENTATION  OF  VAPOR-LIQUID  EQUILIBRIUM  DATA 

It  is  usually  desirable  to  present  the  experimental  vapor-liquid 
equilibrium  data  graphically.  A  number  of  methods  of  presentation 
have  been  developed,  but  the  most  important  are  the  temperature- 
composition  and  the  vapor-liquid  composition  diagrams. 

Phase  Rule.  The  method  of  presentation  must  be  consistent  with 
the  number  of  variables  involved.  For  equilibrium  conditions  the 
number  of  independent  variables  can  be  obtained  from  the  phase  rule 
which  states  that  the  number  of  phases  <j>  ^lus  the  degrees  of  variance 
F  is  equal  to  the  number  of  components  C  plus  2. 

0  +  p  =  C  +  2  (2-1) 

In  the  usual  vapor-liquid  equilibria  two  phases  are  involved:  liquid  and 
vapor.  ?Iowever,  in  some  systems  more  than  one  liquid  phase  may  be 
encountered.  For  the  two-phase  system  the  phase  rule  states  that  the 
degrees  of  freedom  or  variance  are  equal  to  the  number  of  components. 
Thus  a  binary  system  has  two  degrees  of  freedom  and  can  be  repre- 
sented by  two  variables  *  on  rectangular  coordinates.  Three-com- 
ponent systems  involve  three  degrees  of  freedom  and  are  usually 
presented  on  triangular  coordinates.  Multicomponent  systems  with 
more  than  three  components  are  difficult  to  present,  and  special 
methods  are  employed  for  such  systems. 

Temperature-Composition  Diagrams.  By  fixing  the  total  pressure 
of  a  two-component  system,  a  temperature-composition  diagram  can 
be  made.  Figure  2-1  shows  a  temperature-composition  curve  for 
carbon  tetrachloride-carbon  bisulfide  at  a  total  pressure  of  760  mm. 
Any  point  on  the  curve  ABC  gives  the  composition  x,  of  a  mixture  of 
CCU  and  082,  which  boils  at  a  pressure  of  760  mm.  at  a  temperature  £, 
where  t  is  in  degrees  centigrade  and  x  is  the  mol  fraction  of  CS2.  The 
use  of  the  "mol  fraction "  greatly  facilitates  calculations  of  vapor- 
pressure  phenomena.  It  is  the  ratio  of  the  number  of  molecular 
weights  of  one  component  in  a  mixture  divided  by  the  sum  of  the  num- 
ber of  the  molecular  weights  of  all  components.  Mol  per  cent  is  equal 
to  100  times  the  mol  fraction.  The  line  ADC  represents  the  com- 
position of  the  vapor  that  is  in  equilibrium  with  the  liquid  at  any 

16 


PRESENTATION  OF  VAPOR-LIQUID  EQUILIBRIUM  DATA       17 

given  temperature.  Thus  a  liquid  with  the  composition  xi  will  have 
a  vapor  pressure  of  760  mm.  at  the  temperature  t^  and  the  vapor  in 
equilibrium  with  it  will  have  the  composition  yi  =  #2. 

Starting  with  a  mixture  of  the  composition  xi,  at  a  constant  total 
pressure  equal  to  760  mm.,  and  at  a  temperature  below  £2,  there  will  be 
but  one  phase  present,  the  liquid  mixture  of  CCU  and  CSa.  As  the 
temperature  is  raised,  only  a  liquid  phase  will  be  present  until  the 
vertical  line  at  Xi  intersects  the  curve  ABC,  when  a  vapor  phase  of 


80 
A 

:  70 


^60 


S. 

E 


50 


40 


T 


v 
\X4 


0      01      02     03     04     05     06    07     08     09     10 
CC14  Mo!  fraction  C$2  CS2 

FIG.  2-1.     Boiling-point  curve  for  CCh-CSa  mixtures. 

the  composition  x2  will  appear.  Since  there  are  now  two  phases  and  the 
pressure  is  fixed,  there  can  be  but  one  variable,  temperature,  and  the 
composition  of  the  phases  will  depend  upon  it.  Let  the  temperature 
then  be  raised  to  some  point  tz  and  the  liquid  and  vapor  compositions, 
being  no  longer  independent  variables,  must  change  accordingly, 
which  they  do  along  the  curves  ABC  and  ADC,  respectively,  the  liquid 
now  having  a  composition  x*,  and  the  vapor  in  equilibrium  with  it  a 
composition  y*  =  rc4.  It  should  be  remembered  that  the  quantity  of 
CC14  and  CS2  in  the  system  has  not  changed  during  this  process;  there- 
fore, the  change  in  the  compositions  of  the  liquid  and  the  vapor  includes 
such  a  corresponding  change  in  the  relative  proportions  of  each  phase 
that  the  total  composition  of  the  system  remains  the  same,  x\.  Fur- 
thermore, the  relative  proportions  of  the  liquid  phase  and  the  vapor 
phase,  at  the  temperature  /3,  are  as  the  distances  FG  and  EF.  It  will 


18 


FRACTIONAL  DISTILLATION 


be  seen  that,  as  the  temperature  is  raised  farther,  the  proportion  of 
liquid  phase  decreases  until,  when  the  temperature  reaches  a  point 
corresponding  to  the  intersection  of  the  vertical  line  x\  and  the  curve 
ADC,  which  occurs  at  a  temperature  U,  the  vapor  has  the  same  com- 
position as  the  original  liquid,  and  the  liquid  phase  disappears.  At 
higher  temperatures,  there  is  but  one  phase,  and  the  system  again 
becomes  trivariant,  so  that  at  constant  pressure  it  is  possible  to  vary 


1.U 

0.8 
S. 

5*  °-6 

o 

§ 

t; 
|    0.4 
"o 

<M 

OQ.2 

( 

^ 

/ 

/ 

^ 

/ 

/ 

/ 

/ 

/ 

/ 

/ 

/ 

/ 

/ 

/ 

/ 

/ 
' 

/ 

y 

/ 

/ 

/ 

l/_ 

)             0.2             0.4            0.6             0.8            1. 

vug 

Fia.  2-2.     Equilibrium  y,x  data  for  CC14-CS2  mixtures. 

both  the  temperature  and  the  composition  of  the  vapor.  This  is  the 
region  of  superheated  vapor. 

If  the  foregoing  process  is  reversed,  the  steps  can  be  followed  in  the 
same  way.  Starting  with  superheated  vapor  of  a  composition  3/5  =  xi 
and  at  a  temperature  /6,  condensation  will  first  occur  when  the  vertical 
line  2/5  cuts  the  vapor  line  ADC,  when  liquid  of  a  composition  #5  will 
separate  out.  Further  cooling  will  change  both  the  composition  of  the 
liquid  and  the  vapor  along  the  lines  ABC  and  ADC,  respectively,  until 
the  liquid  has  reached  the  composition  x\  when  all  the  vapor  will  have 
disappeared. 

Vapor-Liquid  Equilibrium.  It  is  possible  to  plot  the  same  data  as 
were  used  in  Fig.  2-1  as  vapor  composition  vs.  liquid  composition  at 
either  constant  pressure  or  constant  temperature.  The  data  presented 
in  Fig.  2-1  were  for  constant  pressure  so  they  have  been  replotted  in 


PRESENTATION  OF  VAPOR-LIQUID  EQUILIBRIUM  DATA      19 

Fig.  2-2  for  the  same  conditions.  In  this  curve  the  composition 
yi  =  x2  is  plotted  as  ordinate  with  composition  x\  as  abscissa,  y*  «  x\ 
as  ordinate  with  z3  as  abscissa,  and  so  forth.  This  particular  relation 
is  very  useful  in  distillation  calculations.  It  does  not  give  so  much 
information  as  Fig.  2-1,  owing  to  the  elimination  of  temperature. 
However,  in  most  distillation  calculations  it  is  desired  to  make  a  given 
separation  between  the  components  and  the  temperatures  are  allowed 


120 


100 


8. 

I  60 


40 


20 


I  -  Benzene  -  toluene 
TST-  /sobufano/- wafer" 


~  Acetone -chloroform 
H  -Acetone  -carbonctisutfide 

I        I        I        I        I        I 


3  02  0.4  06  0.8 

x,Mol  fraction  in  liquid 

Fio.  2-3.     Temperature-composition  diagram. 


1.0 


to  adjust  themselves  accordingly.  On  Pig.  2-2  the  45°  line  represents 
a  vapor  of  the  same  composition  as  the  liquid.  If  the  temperature  is 
important,  this  variable  can  be  plotted  vs.  the  liquid  composition  on 
the  same  figure. 

The  curves  given  in  Figs.  2-1  and  2-2  are  termed  the  normal  type. 
However,  there  are  several  other  common  types  of  curves,  In  Fig. 
2-3  temperature-composition  diagrams  for  constant  total  pressure  are 
given  for  four  different  types  of  binary  mixtures,  and  in  Fig.  2-4  the 
corresponding  vapor-liquid  diagrams  are  given  for  the  four  same 

mixtures. 

Type  I  is  normal,  i.e.,  the  composition  of  the  equilibrium  vapor  is 
always  richer  in  the  same  component  than  the  composition  of  the 
liquid,  thus  by  repeated  operations  it  is  possible  to  obtain  complete 
separation. 


20 


FRACTIONAL  DISTILLATION 


In  type  II  the  temperature-composition  diagram  passes  through  a 
minimum,  and  the  vapor-liquid  composition  diagram  crosses  the 
diagonal.  Thus  there  are  mixtures  that  have  lower  boiling  points  than 
either  of  the  pure  components  at  the  same  pressure.  In  other  words, 
the  mixture  is  the  minimum  boiling-point  type.  When  such  tempera- 
ture-composition diagrams  are  encountered,  the  vapor-liquid  com- 
position curve  will  always  cross  the  45°  line.  In  the  region  below  this 
intersection  with  the  diagonal,  the  equilibrium  vapor  is  richer  in  one 


02 


[  "Benzene-toluene 
H  -  Acetone -carbondisutfide 
HI  -  A  cetone  -  ch/o  ro  fo  rm 
IF  -  Isobutanol-  water 


Q  0.2         *  0.4  0.6  0.8 

x,Mo!  fraction  in  liquid 
FIG.  2-4.     Vapor-liquid  equilibrium  curves. 


1.0 


component  than  the  liquid;  above  this  intersection,  the  vapor  is  poorer 
in  this  component  than  the  corresponding  liquid  from  which  it  came. 
Thus  the  volatilities  have  reversed.  Where  the  vapor-liquid  curve 
crosses  the  45°  line,  the  vapor  has  the  same  composition  as  the  liquid 
and  operations  based  on  producing  an  equilibrium  vapor  from  this 
liquid  would  not  be  able  to  separate  mixtures  of  this  composition. 
This  particular  composition  is  called  a  constant  boiling  mixture,  or 
azeotropic  mixture,  sincejt^will  yaporize^without  any  change  in  Com- 
position and,  theipfore,  without  any  change  in  temperature  during  the 
evaporation.  *~ 

TyplFHTTs  the  reverse  of  type  II.  In  this  case  there  are  mixtures 
that  have  boiling  points  higher  than  either  of  the  pure  components  at 
the  same  pressure.  It  will  be  noted  in  Fig.  2-4  that  the  curve  of  type 


PRESENTATION  OF  VAPOR-LIQUID  EQUILIBRIUM  DATA       21 

III  also  crosses  the  45°  line  but  curve  II  cuts  the  45°  line  with  the 
slope  less  than  1  while  curve  III  crosses  the  45°  line  with  the  slope 
greater  than  1.  Curve  III  is  of  the  maximum  boiling-point  type,  and 
the  particular  composition  at  which  the  curve  crosses  the  45°  line  is 
called  a  maximum  constant  boiling  mixture  or  a  maximum  boiling 
azeotrope. 

The  curve  of  type  IV  is  similar  to  that  of  type  II  except  that,  for  a 
considerable  range  of  composition,  the  temperature  of  the  liquid  phase 
is  constant.  This  curve  type  is  characteristic  of  a  partly  miscible 
liquid  system.  In  the  immiscible  region,  two  liquid  phases  are  present 
and  the  phase  rule  indicates  that  the  boiling  temperature  of  the  mix- 
ture must  be  constant.  In  the  diagram,  the  over-all  composition  of 
the  liquid  is  plotted  as  x  regardless  of  whether  one  or  two  phases  are 
present.  There  is  no  single  liquid  phase  that  has  a  composition  equal 
to  the  value  given  in  the  two-phase  region.  The  y,x  data  for  this 
system  are  given  in  Fig.  2-4.  In  this  case  the  y,x  curve  crosses  the 
diagonal  in  the  two-phase  region;  thus  at  this  intersection  the  com- 
position of  the  vapor  is  the  same  as  that  of  the  combined  liquid  phases. 
Such  a  mixture  can  be  evaporated  to  dryness  at  constant  pressure 
without  change  in  composition  or  temperature.  The  mixture  of  this 
particular  composition  is  termed  a  pseudo-azeotrope.  This  terminol- 
ogy is  sometimes  applied  to  any  two-phase  mixture,  but  the  original 
usage  of  the  term  azeotrope  by  Wade  and  Merriman  (Ref .  50)  implied 
that  the  liquid  could  be  evaporated  to  dryness  without  change  in 
composition.  Only  the  two-phase  mixture  corresponding  to  the  inter- 
section of  the  y,x  curve  with  the  y  =  x  line  can  be  evaporated  without 
changing  composition. 

In  other  cases  the  y,x  curve  may  not  cross  the  diagonal  in  the  two- 
phase  region,  and  such  mixtures  do  not  form  pseudo-azeotropes,  but 
they  may  form,  and  usually  do,  true  azeotropes  in  one  of  their  single- 
phase  regions. 

Literature  Data.  The  vapor-liquid  equilibria  for  a  large  number  of 
mixtures  have  been  experimentally  determined,  and  Table  2-1  lists 
some  of  the  more  reliable  and  useful  determinations. 

The  data  given  in  the  table  represent  a  large  amount  of  experi- 
mental effort,  and  owing  to  the  difficulties  of  making  such  determina- 
tions a  number  of  investigators  have  tried  to  develop  theoretical  and 
empirical  methods  of  predicting  such  vapor-liquid  equilibria  from  the 
physical  properties  of  the  pure  components.  While  certain  correla- 
tions have  been  developed  by  this  method,  reliable  experimental  deter- 
minations are  still  to  be  preferred  to  any  such  calculated  values. 


22 


FRACTIONAL  DISTILLATION 
TABLE  3-1.    VAPOB-LIQUID  EQUILIBRIUM  DATA 


System 

Constant 
TorP 

Tech- 
nique * 

Ref. 

Acetaldehyde—  water  

760  mm. 

_ 

19 

Acetic  acid-acetic  anhydride 
—acetone  

750  mm. 
760  mm. 

C.D. 
C.D. 

27 
55 

-acetone—  water  . 
—benzene  

760  mm. 
758  mm. 

C.D, 
C.D. 

55 
27 

—ethyl  acetate  

760  mm. 

C.D. 

13 

-methyl  amyl  ketone  .  . 
-water  

Acetone-carbon  disulfide  . 

-chloroform  
—ethyl  ether  ... 

760  mm. 
125,  250,  300,  760  mm. 
760  mm. 
35.2°C. 
760  mm. 
760  mm. 
30°C. 

C.D. 
C.D. 
C.D. 
B.M. 

D.D. 
C.D. 

29 
14 
55 
17 
19 
38 
39 

—  isopropanol  

25°C. 

D.F. 

32 

—  methanol  

755  mm. 

C.D. 

28 

—  w-butanol  

760  mm. 

C.D. 

7 

—nitrobenzene  

20°C. 

D.F. 

52 

—  trichloroethylene    .   .  . 

755  mm. 

C.D. 

47 

—water     .       ,      

25°C. 

B.M. 

1 

760  mm. 
200,  350,  500,  760  mm. 
50,  100,  200  p.sa.a. 
150,  300,  760  mm. 

C.D. 
C.D. 
C.D. 
C.D. 

5 
29 
31 
30 

1  to  20  atm. 

3 

Aniline—  benzene  

70°C. 

C.D. 

24 

Benzene-carbon  disulfide  
~carbon  tetrachloride 

19.9°C. 
40°C. 

B.M. 
C.D. 

17 
40 

—chloroform  

760  mm. 
760  mm. 

D.D. 

38 

48 

—  cyclohexane  

740  mm. 

C.D. 

16 

-ethyl  bromide  
-ethylene  chlorohydrin  . 
-ethylene  dichloride  
—methanol  

760  mm. 
760  mm, 
100,  200,  400,  760  mm. 
40°C. 

C.D. 
C.D. 
C.D. 

48 
44 
4 
22 

—  n-hexane  

760  mm. 
735  mm. 

C.D. 
C.D. 

53 

46 

—nitrobenzene  

20°C. 

D.F. 

52 

—phenol  ,  ,  

70°C. 

C.D. 

25 

—  n-propanol  

40°C. 

C.D. 

22 

—2  2  3  trimethylbutane  

740  mm. 

C.D. 

16 

*B,M.  «•  bomb  method 
C.D.  —  continuous  distillation 
C.M.  —  circulation  method 


D.B.  ••  dew-point,  bubble-point  method 
D.D.  «  dynamic  distillation 
D.F.  «  dynamic  flow 


PRESENTATION  OF  VAPOR-LIQUID  EQUILIBRIUM  DATA      23 
TABLE  2-1.     VAPOR-LIQUID  EQUILIBRIUM  DATA  (Continued) 


System 

Constant 
TorP 

Tech- 
nique* 

Ref. 

750  mm. 
760  mm. 
760  mm. 
760  mm. 
760  mm. 
760  mm. 
760  mm. 
20°C. 
19  8°C. 
29.2°C. 
20.0°C. 
17  0°C. 
20  0°C. 
745  mm. 
49.9°C. 
745  mm. 
760  mm. 
760  mm. 
762  mm. 
100°C. 
757  mm. 
20°C. 
760  mm. 
740  mm. 
25°C. 
740  mm. 
10,  50  mm. 

0  1  mrn. 
0  1  mm. 
-25.5,  -12,  0,  12, 
23°C. 
760  mm. 
20°C. 
760  mm. 
30°C, 
760  mm. 
95,  190,  380  mm. 
760  mm. 
39.8°C. 
20°C. 
20°C. 
120°C. 

D.D. 
C.D. 
C.D. 
D.D. 
C.D. 
D.D. 
D.D. 
B.M. 
B.M. 
B.M. 
B.M. 
B.M. 
D.F. 

D.D. 

C.D. 
C.D. 
C.D. 

D.F. 
D.D. 
C.D. 
D.F. 
C.D. 
C.D. 

C.D. 
C.D. 
C.M. 

C.D. 
D.F. 
C.D. 
B.M. 
C.D. 
C.D. 
C.D. 
D.D. 
D.F. 
D.F. 
C.D. 

35 
7 
44 
45 
44 
45 
38 
17 
17 
17 
17 
17 
52 
48 
56 
48 
48 
34 
8 
20 
48 
52 
37 
16 
51 
16 
4 

33 
33 
26 

13 
10 
15 
12 
23 
2 
21 
5 
52 
52 
20 

Ti-Butanol—  /t-butyl  acetate           

-ethylene  chlorohydrin 
—water     

i-Butanol-ethylene  chlorohydrin  
—water             .             .... 

Carbon  disulfide-carbon  tetrachloride  
-chloroform  
—  cyclohexane  
-ethyl  ether.   .           .... 
—  isobutyl  chloride  

—  isopentane             

-nitrobenzene  
Carbon  tetrachloride-ethanol  ...   . 
-ethyl  acetate 

—ethyl  ether  .... 
-tetrachloroethylcne 
-toluene       .        ... 
Chlorobenzene-ethylenc  bromide  
Chloroform-methanol  
—nitrobenzene  

Cvclohexane—cyclohexane  

-ethanol     
223  trimethylbutanc 

Diethvl  benzene  o-dichlorobenzene  

Diethyl  hexyl  phthalate-diethyl  hexyl  seba- 
cate                              

Dioctyl  phthalate-diethyl  hexyl  sebacate 
Ethane—  ethylene        ... 

TfltKannl   f»tlrv1  acetate                  

—  ethyl  ether        ....      .             •  •  • 

methanol—  water     

—  7^-heptane          

24 


FRACTIONAL  DISTILLATION 


TABLE  2-1.     VAPOR-LIQUID  EQUILIBRIUM  DATA  (Continued) 


System 

Constant 
^orP 

Tech- 
nique * 

Ref. 

Ethylene  chlorohydrin-toluene 

760  mm. 

C.D. 

44 

Formic  acid—  water  

750  mm. 

C.D. 

27 

Furfural—  2-methylfuran  

738  mm. 

C.D. 

18 

—water  

760  mm. 

19 

w-Heptane-w-pentane  .  . 

10.3,  20.6,  31.1atm. 

D.B. 

9 

-toluene  

760  mm. 

C.D. 

6 

w-Hexane-methanol  

45°C. 

B.M. 

12 

Hydrochloric  acid-water.    .    .  . 

751  mm. 

C.D. 

28 

Isobutylene-propane  .... 

14.1,  21.1,28.2, 

C.D. 

41 

35  2,  42.2atm. 

Isopentane-propane  . 

0  to  180°C. 

D.B. 

49 

Isopropanol—  nitromethane-water 

760  mm. 

C.D. 

42 

—water  

760  mm. 

C.D. 

42 

760  mm. 

C.D. 

21 

760  mm. 

— 

19 

Methanol-methyl  acetate  . 

39  8°C. 

D.D. 

5 

-nitrobenzene.    .                  .    . 

20°C. 

D.F. 

52 

-water  

59.4°C. 

D.D. 

54 

60,  115,  165p.s.i.a. 

C.D. 

31 

200,  350,  500,  760  mm. 

C.D. 

29 

Methyl  ethyl  ketone-water 

200,  350,  500,  760  mm. 

C.D. 

29 

Nitric  acid—  water  

760  mm. 



19 

Nitrogen-oxygen     

0.4  to  45atm. 

C.M. 

11 

Nitromethane-water  

760  mm. 

C.D. 

42 

n-Octane-toluene  . 

760  mm. 

C.D. 

6 

Phenol—  wi-cresol 

760  mm. 

D.D. 

36 

-o-cresol  

760  mm. 

D.D. 

36 

-water  

40,  260,  760  mm. 

D.D. 

36 

43.4,  58.4,  73.4, 

C.D. 

43 

98.4°C. 

n-Propanol-water  

30.5°C. 

D.D. 

54 

Pyridine-water  .  .  .                     .          ... 

80.1°C. 

C.D. 

56 

Sulfuric  acid-water  .  .                  

760  mm. 

-— 

57 

References 

1.  BAERE,  Me  VICAR,  and  FERGUSON,  /.  Phys.  Chem.,  34,  1310  (1930). 

2.  BEEBE,  COULTER,  LINDSAY,  and  BAKER,  2nd.  Eng.  Chem.t  34,  1501  (1942). 

3.  BOSNJAKOVIC,    "Technische    Termodynamik    II,"    diagrams,   T.   Steinkopf, 
Dresden,  1937. 

4.  BRAGG  and  RICHARDS,  Ind.  Eng.  Chem.,  34,  1088  (1942). 

5.  BREDIG  and  BAYER,  Z.  physik.  Chem.,  130,  15  (1927). 

6.  BROMILEY  and  QUIGGLE,  Ind.  Eng.  Chem.,  25,  1136  (1933). 

7.  BRUNJES  and  FURNAS,  Ind.  Eng.  Chem.,  27,  396  (1935). 

8.  CAREY,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1929. 

9.  CXJMMINGS,  STONE,  and  VOLANTE,  Ind.  Eng.  Chem.,  25,  728  (1933). 


PRESENTATION  OF  VAPOR-LIQUID  EQUILIBRIUM  DATA       25 

10.  DESMAROUX,  Mem.  poudres,  23,  198  (1928). 

11.  DODGE  and  DUNBAR,  Chem.  Met.  Eng.,  36,  336  (1929). 

12.  FERGUSON,  FREED,  and  MORRIS,  Landolt-Bernstein  Tabellen,  eg  IIIc,  2470, 
and  /.  Phys.  Chem.,  7,  87  (1933). 

13.  FURNAS  andLEiGHTON,  Ind.  Eng.  Chem.,  29,  709  (1937). 

14.  GILMONT  and  OTHMER,  Ind.  Eng.  Chem.,  36,  1061  (1944). 

15.  GRISWOLD  and  DINWIDDIE,  Ind.  Eng.  Chem.,  34,  1188  (1942). 

16.  HARRISON  and  BERG,  Ind.  Eng.  Chem.,  38,  117  (1946). 

17.  HIRSHBERG,  Bull.  soc.  Mm.  Beiges,  41,  163-195  (1932). 

18.  HOLDREN  and  HDCON,  Ind.  Eng.  Chem.,  38,  1061  (1946). 

19.  " International  Critical  Tables,"   McGraw-Hill  Book  Company,  Inc.,  New 
York,  1926-1930. 

20.  LACKER  and  HUNT,  J.  Am.  Chem.  Soc.,  63,  1752  (1941). 

21.  LANGDON  and  KEYES,  Ind.  Eng.  Chem.,  34,  938  (1942). 

22.  LEE,  /.  Phys.  Chem.,  36,  3569  (1931). 

23.  LEWIS  and  CAREY,  Ind.  Eng.  Chem.,  24,  882  (1932). 

24.  MARTIN  and  COLLE,  J.  Chem.  Soc.,  1932,  2658  (1932). 

25.  MARTIN  and  GEORGE,  /.  Chem.  Soc.,  1933. 

26.  MCDONALD  and  McMiLLAN,  Ind.  Eng.  Chem.,  36,  1175  (1944). 

27.  OTHMER,  Ind.  Eng.  Chem.,  Anal.  Ed.,  4,  232  (1932). 

28.  OTHMER,  Ind.  Eng.  Chem.,  20,  743  (1928). 

29.  OTHMER  and  BENENATI,  Ind.  Eng.  Chem.,  37,  299  (1945). 

30.  OTHMER  and  JOSEFOWITZ,  Ind.  Eng.  Chem.,  39,  1175  (1947). 

31.  OTHMER  and  MORLEY,  Ind.  Eng.  Chem.,  38,  751  (1946). 

32.  PARKS  and  CHAFFEE,  /.  Phys.  Chem.,  31,  439  (1927). 

33.  PERRY  and  FUGUITT,  Ind.  End.  Chem.,  39,  782  (1947). 

34.  QUINN,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1940. 

35.  RHODES  and  BOSCOM,  Ind.  Eng.  Chem.,  19,  480  (1927). 

36.  RHODES,  WELLS,  and  MURRAY,  Ind.  Eng.  Chem.,  17,  1199  (1925). 

37.  ROSANOFF,  BACON,  and  WHITE,  J.  Am.  Chem.  Soc.,  36,  1803  (1914). 

38.  ROSANOFF  and  EASLEY,  /.  Am.  Chem.  Soc.,  31,  970  (1909). 

39.  SAMESHIMA,  /.  Am.  Chem.  Soc.,  40,  1482  (1918). 

40.  SCATCHARD,  WOOD,  and  MOCHEL,  /.  Am.  Chem.  Soc.,  62,  712  (1940). 

41.  SCHEELINE  and  GILLILAND,  Ind.  Eng.  Chem.,  31,  1050  (1939). 

42.  SCHUMACHER  and  HUNT,  Ind.  Eng.  Chem.,  34,  701  (1942). 

43.  SIMS,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1933. 

44.  SNYDER  and  GILBERT,  Ind.  Eng.  Chem.,  34,  1533  (1942). 

45.  STOCKHARDT  and  HULL,  Ind.  Eng.  Chem.,  23,  1438  (1931). 

46.  TONGBERG  and  JOHNSTON,  Ind.  Eng.  Chem.,  26,  733  (1933). 

47.  TREYBAL,  WEBER,  and  DALEY,  Ind.  Eng.  Chem.,  38,  815  (1946). 

48.  TYRER,  J.  Chem.  Soc.,  101,  81,  1104  (1912). 

49.  VAUGHAN  and  COLLINS,  Ind.  Eng.  Chem.,  34,  885  (1942). 

50.  WADE  and  MERRIMAN,  /.  Chem.  Soc.,  99,  997  (1911). 

51.  WASHBURN  and  HANDORF,  J.  Am.  Chem.  Soc.,  67,  441  (1935). 

52.  WEISSENBERGER,  HENKE,  and  KAWENSKI,  /.  prakt.  Chem.,  113,  171  (1926). 

53.  WILLIAMS,  ROSENBERG,  and  ROTHENBERG,  Ind.  Eng.  Chem.,  40, 1273  (1948). 

54.  WREWSKI,  Z.  physik.  Chem.,  81,  1  (1912). 

55.  YORK  and  HOLMES,  Ind.  Eng.  Chem.,  34,  345  (1942). 

56.  ZAWIDSKI,  Z.  physik.  Chem.,  36,  129  (1900). 

57.  ZEISBERG,  Chem.  Met.  Eng.,  27,  22  (1922). 


CHAPTER  3 
CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 

The  calculation  of  vapor-liquid  equilibria  is  important  because  of 
the  difficulty  of  obtaining  experimental  values  and  because  it  gives  a 
picture  of  the  general  behavior  of  liquid-vapor  mixtures.  The  basic 
thermodynamic  relationships  for  such  equilibria  are  complex,  and  in 
most  cases  they  involve  unknown  factors  or  quantities  which  limit 
their  usefulness  until  simplifying  assumptions  are  made.  It  is  the 
limitations  of  the  simplifying  assumptions  that  restrict  the  applicabil- 
ity of  the  thermodynamic  relations.  However,  even  for  the  cases  in 
which  such  approximations  do  not  give  satisfactory  quantitative  rela- 
tionships, they  do  serve  as  valuable  criteria  for  estimating  the  behavior 
of  a  distillation  process  and  they  help  to  clarify  and  explain  the 
divergencies  that  are  frequently  noted. 

VAPOR  PRESSURES   OF  COMPLETELY   MISCIBLE  LIQUID   MIXTURES 
AT  CONSTANT  TEMPERATURE 


Raoult's  Law.  Whe^<m&  liquid  is  dissolved  in  another,  the  partial 
pressure  of  each  is  decreased.  Assume  two  liquids,  the  molecules  of 
which  are  the  same  size  and  which  mix  without  the  complicating  effects 
of  molecular  association,  chemical  combination,  and  the  like.  In  an 
equimolecular  mixture  of  two  such  liquids,  each  unit  of  surface  area  of 
the  liquid  mixture  will  have  in  its  surface  half  as  many  molecules  of 
each  component  as  exist  in  the  liquid  surface  of  that  component  in  the 
pure  state.  Hence  the  escaping  tendency  or  partial  pressure  of  each 
component  in  the  mixture  will  be  half  that  of  the  same  component  in 
the  pure  state.  Similarly,  in  a  mixture  containing  25  mol  per  cent  of 
the  first  component  and  75  mol  per  cent  of  the  second,  the  first  will 
exert  a  partial  pressure  25  per  cent  of  that  of  this  component  in  the 
pure  state.  In  more  general  terms,  for  any  such  mixture  the  partial 
pressure  of  any  component  will  equal  the  vapor  pressure  of  that  com- 
ponent in  the  pure  state  times  its  mol  fraction  in  the  liquid  mixture. 
This  generalization  is  known  as  Raoult's  law  (Ref.  19).  It  is  expressed 
in  the  relationship,  pa  =  Paxa,  where  pa  is  the  partial  pressure  of  the 
component  A  in  the  solution,  xa  is  its  mol  fraction  in  the  solution  and 

26 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  27 

Pa  is  the  vapor  pressure  of  the  component  A  in  its  pure  state,  If  P6  is 
the  vapor  pressure  of  pure  B,  and  pb  the  partial  pressure  in  the  mixture, 
then  pi  =  P&a;&.  This  relationship  is  shown  graphically  in  Fig.  3-1, 
where  the  abscissas  are  the  mol  per  cent  of  the  two  components,  A  and 
By  in  the  liquid  portion.  The  ordinates  are  pressures,  C  being  the 
vapor  pressure  of  pure  A,  and  D  that  of  pure  B.  The  lines  AD  and 
BC  represent  the  partial  pressures  of  the  components  over  any  mixture, 
while  the  line  CD  is  the  total  pressure  of  the  mixture. 

Deviations  from  Raoult's  Law.  In  view  of  the  above  assumptions 
as  to  equal  molecular  size,  absence  of  association,  etc.,  it  is  not  sur- 
prising to  find  Raoult's  law  honored  more  in  the  breach  than  in  the 
observance.  Nonetheless  mixtures  of  some  organic  liquids,  such  as 
benzene-toluene,  deviate  from  it 
but  little.  The  deviations  of  mix- 
tures of  hydrocarbons  of  the  same 
series  can  usually  be  neglected  for  a 
great  deal  of  engineering  work,  and 
even  for  mixtures  of  a  number  of 
series  this  is  often  true.  For  mix-  0  Mol  ent  A 

tures  of  aromatic  and  aliphatic  com-     FlG<    3_L    gchematic    diagram     for 
pounds,    however,   the   deviations     Raoult's  law. 
are  often  large,  though  never  of  the 

order  of  magnitude  of  such  mixtures  as  hydrochloric  acid  and  water, 
and  the  like.  Organic  stereoisomers  obey  it  very  closely  as  would  be 
expected  from  the  considerations  upon  which  it  is  based.  However, 
the  data  for  the  great  majority  of  other  liquids,  when  plotted  as  shown 
in  Fig.  3-1,  deviate  largely  from  the  lines  BC  and  AD.  When  very 
near  points  C  and  D,  the  deviation  for  any  component  is  slight  if  that 
component  is  present  in  very  large  amount.  This  ordinarily  is 
expressed  by  saying  that  in  dilute  solution  Raoult's  law  applies  to  the 
solvent.  Since  the  deviation  from  Raoult's  law  may  be  either  positive 
or  negative,  great  or  small,  this  graphical  generalization  serves  as  a 
convenient  standard  of  comparison. 

Henry's  Law.  This  relation  is  a  modification  of  Raoult's  law  which 
applies  to  the  vapor  pressure  of  the  solute  in  dilute  solutions,  just  as 
Raoult's  law  applies  to  that  of  the  solvent.  Henry's  law  states  that  the 
partial  pressure  of  the  solute  is  proportional  to  its  concentration  in  the 
solution.  In  analogy  with  Raoult's  law  it  may  be  expressed  by  the 
equation 

Pa,  =  kxa  (3-1) 


28 


FRACTIONAL  DISTILLATION 


where  pa  =  partial  pressure  of  the  solute 
xa  —  its  mol  fraction 
k  =  an  experimentally  determined  constant 

Comparison  with  Raoult's  law,  pa  =  Paxa,  shows  that  they  differ  only 

in  the  constant  that  determines  the  slope  of  the  line.    This  constant  is 

the  vapor  pressure  of  the  pure  com- 
ponent in  the  one  case,  while  it  must 
be  experimentally  determined  in 
the  other.  A  typical  partial  pres- 
sure curve  for  one  component  of  a 
liquid  mixture  is  shown  in  Fig.  3-2 
where  BD  is  the  range  over  which 
Henry's  law  applies,  while  Raoult's 
law  holds  over  the  section  EC,  where 
C  is  the  vapor  pressure  of  pure  A. 

Dalton's  Law.  The  most  com- 
monly used  rule  for  relating  the 
composition  of  the  vapor  phase  to 
the  pressure  and  temperature  is 
Dalton's  law  (Ref.  8).  It  states 

that  the  total  pressure  is  equal  to  the  sum  of  the  partial  pressures  of 

the  components  present,  i.e., 


B 


Mol  percent  A. 

Fia.      3-2.     Schematic      diagram 
Henry's  law. 


100 
for 


Pi  +  P2  +  Pa  + 


(3-2) 


where  pi,  p2,  pa  =  partial  pressures  of  components  1,  2,  and  3 

IT  =  total  pressure 

For  Dalton's  law  partial  pressure  is  defined  as  the  pressure  that 
would  be  exerted  by  a  component  alone  at  the  same  molal  concentra- 
tion that  it  has  in  the  mixture.  If  the  perfect-gas  laws  apply  to  each 
of  the  components  individually  and  to  the  mixture,  it  can  be  shown 
that  the  partial  pressure  of  any  component  is  equal  to  the  mol  fraction 
times  the  total  pressure,  i.e., 


pi  = 


(3-3) 


where  y\  is  the  mol  fraction  of  component  1  in  the  vapor. 

This  is  the  most  commonly  used  form  of  Dalton's  law.  For  the  pre- 
diction of  vapor-liquid  equilibria,  it  is  usually  combined  with  Raoult's 
law  to  give 

2/iTT  »  xfi  (3-4) 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  29 

This  combination  gives  the  composition  of  vapor  as  a  function  of  the 
composition  in  the  liquid  with  the  total  pressure  and  vapor  pressure  as 
proportionality  constants.  Thus,  fixing  the  temperature  and  the 
total  pressure  defines  TT,  PI,  and  the  relationship  between  the  vapor  and 
the  liquid  composition.  It  is  to  be  noted  that  the  assumption  of 
Raoult's  and  Dalton's  laws  leads  to  the  conclusion  that  the  relationship 
between  the  vapor  and  liquid  composition  of  a  given  component  is  a 
function  of  the  temperature  and  pressure  only  and  is  independent  of 
the  other  components  present.  The  only  influence  of  the  other  com- 
ponents is  the  fact  that  they  may  be  instrumental  in  determining  the 
relationship  between  the  temperature  and  the  total  pressure. 

Deviations  from  Dalton's  Law.  Both  forms  of  Dalton's  law  are 
satisfactory  for  engineering  uses  for  most  gas  mixtures  at  pressures  of 
1  atm.  or  lower  because  deviations  from  the  perfect-gas  law  are  usually 
small  in  this  region.  However,  when  higher  pressures  are  encountered 
and  appreciable  deviations  from  the  perfect-gas  law  are  found,  Dal- 
ton's law  becomes  unsatisfactory.  A  later  section  of  this  chapter  will 
consider  methods  of  handling  these  deviations,  but  as  a  general  rule 
Dalton's  law  should  not  be  employed  for  cases  in  which  the  deviations 
from  the  perfect-gas  law  are  large. 

•  Volatility.  The  term  " volatility"  is  loosely  used  in  the  literature, 
generally  as  equivalent  to  vapor  pressure  when  applied  to  a  pure  sub- 
stance; as  applied  to  mixtures,  its  significance  is  very  indefinite. 
Because  of  the  convenience  of  the  term,  the  volatility  of  any  substance 
in  a  homogeneous  liquid  will  be^defined  as  its  partial  pressure  in  the 
vapor  in^  equilibrium  with  tHat  liquid,  divided  by  its  mol  fraction  in  the 
liquid nr  the  substance  is  in  the  pure  state,  its  mol  fraction  is  unity 
and  its  volatility  is  identical  with  its  vapor  pressure.  If  the  substance 
exists  in  a  liquid  mixture  that  follows  Raoult's  law,  its  volatility  as 
thus  defined  is  still  obviously  equal  to  its  vapor  pressure  in  the  pure 
state;  i.e.,  its  volatility  is  normal.  If  the  partial  pressure  of  the  sub- 
stance is  lower  than  that  corresponding  to  Raoult's  law,  e.g.,  that  of 
hydrochloric  acid  in  dilute  aqueous  solutions,  the  volatility  according 
to  this  definition  is  less  than  that  of  the  pure  substance,  i.e.,  is  abnor- 
mally low.  Similarly,  if  the  partial  pressure  is  greater  than  that 
indicated  by  Raoult's  law,  e.g.,  that  of  aniline  dissolved  in  water, 
the  volatility  is  abnormally  high.  The  volatility  of  a  substance 
in  mixtures  is  therefore  not  necessarily  constant  even  at  con- 
stant temperature  but  depends  on  the  character  and  amount  of  the 
components. 


30  FRACTIONAL  DISTILLATION 

Relative  volatility  is  the  volatility  of  one  component  divided  by  that 
of  another.  Since  the  volatility  of  the  first  component  of  a  mixture, 
0a,  is  its  partial  pressure,  pa  divided  by  its  mol  fraction  xa,  and  that  of 
the  second  /?&  =  pb/x*,  the  volatility  of  the  first  relative  to  the  second 
is  fta/ffb  =  paVb/pbXa.  When  Dalton's  law  applies,  the  relative  amount 
of  any  two  components  in  the  vapor  (expressed  in  mols)  is  ya/yb  =*  Pa/p** 

&  =  »-•£  =  a  (3-5) 

&      yb  xa  ^     J 

Owing  to  the  utility  of  the  mol  fraction  ratio  given  in  Eq.  (3-5)  for 
distillation  calculations,  the  group  will  be  used  as  the  definition  of  the 
relative  volatility  for  all  cases  whether  or  not  Dalton's  law  applies. 
Thus,  in  this  text  the  relative  volatilities  of  any  two  components  in  a 
mixture  will  be  defined  as  the  ratio  of  y/x  values  for  the  two  components. 

In  any  constant-boiling  homogeneous  liquid  mixture  the  composi- 
tion of  the  liquid  is  identical  with  that  of  the  vapor  in  equilibrium  with 
it,  i.e.,  x  =  y]  hence,  the  relative  volatility  a  is  unity. 

Volatility,  like  vapor  pressure,  increases  rapidly  with  rise  in  tem- 
perature. The  ratio  of  the  pressures  of  pure  substances  does  not 
change  rapidly  with  change  in  temperature,  and  the  same  is  true  of 
relative  volatilities;  whereas  vapor  pressures  always  increase  with 
temperature,  relative  volatility  may,  in  a  given  case,  either  rise  or  fall, 
depending  on  the  nature  of  the  components.  At  constant  temperature 
the  relative  volatility  is  independent  of  the  liquid  composition  for 
systems  that  obey  Raoult's  law;  however,  for  most  systems  a  is  a  func- 
tion of  the  liquid  composition  and  frequently  is  greater  than  unity  for 
one  range  of  concentrations  and  less  than  unity  for  another  range. 
Relative  volatility  is  the  most  important  factor  in  determining  ease  of 
separation  of  components  by  distillation. 

Raoidt-Dalton  Laws;  Example.  The  data  of  Rhodes,  Wells,  and  Murray 
(Ref .  20)  indicate  that  the  system  phenol-o-cresol  obeys  Raoult's  law.  Using  the 
vapor  pressure  data  tabulated  at  the  top  of  page  31  together  with  Raoult's  and 
Dalton's  laws,  construct  the  following  curves  for  75  mm.  Hg  abs.  pressure: 

1.  Temperature-mol  fraction  diagram  giving  both  the  vapor  and  the  liquid 
curves. 

2.  Mol  fraction  ,of  phenol  in  vapor  phase,  y,  vs.  the  mol  fraction  of  phenol  in 
liquid,  x. 

3.  Relative  Volatility  vs.  the  mol  fraction  of  phenol  in  the  liquid. 

4.  Mol  fraction  of  phenol  in  the  vapor  vs.  the  mol  fraction  of  phenol  in  the  liquid, 
using  as  an  average  value  of  the  relative  volatility,  a,  the  arithmetic  mean  value 
of  a  at  x  »»  0,  and  x  «  1.0. 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 
DATA  ON  VAPOR  PRESSTJRES 


31 


Tt  °C. 

Phenol,  pres- 
sure, mm. 

o-Cresol,  pres- 
sure, mm. 

113.7 

75.0 

57  8 

114.6 

78.0 

59,6 

115.4 

81.0 

61  6 

116  3 

84.0 

63  8 

117.0 

87.0 

65.7 

117.8 

90.0 

68  0 

118  6 

93.0 

70.5 

1919.4 

96.0 

73.0 

120  0 

98  6 

75.0 

Solution 

1.  Let  mol  fraction  of  phenol  in  liquid  •»  xp. 

2.  Choose  a  temperature,  T. 

3.  Read  vapor  pressures  at  T:PP,  Poc. 

4.  Employ  the  combined  form  of  Raoult's  and  Dalton's  laws. 

XpPp  "f  XocPoe   «  75 


_ 

Pp    -  Poc 


TABLE  3-1 


Tt  °C. 

Poo 

PP 

xp 

yp 

Of 

yp  (calc.  on 

a  -  1.313) 

113.7 

57.8 

75 

1.0 

1.0 

1.295 

1.0 

114.6 

59.6 

78 

0.837 

0.87 

1  31 

0.87 

115.4 

61.6 

81 

0.691 

0.746 

1.315 

0.748 

116.3 

63.8 

84 

0.555 

0.622 

1.315 

0.622 

117.0 

65.7 

87 

0.437 

0.507 

1.32 

0.506 

117.8 

68 

90 

0.318 

0.383 

1.323 

0.382 

118.6 

70.5 

93 

0.2 

0  248 

1.32 

0.247 

119.4 

73.0 

96.7 

0.084 

0.109 

1.325 

0.108 

120.0 

75 

99.8 

0 

0 

1.33 

0 

ypxoc 


•  ___-_--~ 
'  1  4-  («  -  l)xp 


32 


FRACTIONAL  DISTILLATION 


These  values  are  plotted  in  Fig.  3-3.  An  a  of  1.313  was  chosen  to  calculate 
equilibrium  curve.  Within  the  accuracy  of  the  calculations,  this  curve  was  coin- 
cident with  the  curve  obtained  in  Part  2. 


0.2  0.4  06  08  1.0 

Mol  fraction  phenol  in  liquid,  %p 

FIG.  3-3.     Calculated  vapor-liquid  equilibria  for  system  phenol  — o-cresol. 

Basic  Thermodynamic  Relations.  Most  mixtures  are  not  ideal, 
and  the  deviations  from  Raoult's  law  are  large  in  all  but  a  limited  num- 
ber of  cases.  For  such  mixtures,  the  basic  thermodynamic  relations 
can  be  employed  to  formulate  expressions  for  the  equilibria  involved. 
It  is  not  the  purpose  of  this  text  to  consider  in  detail  the  thermody- 
namics of  solutions,  but  an  appreciation  of  this  subject  is  necessary  for 
an  understanding  of  the  following  sections  on  vapor-liquid  equilibria. 
For  background  in  this  field  the  reader  is  referred  to  any  modern  text 
on  chemical  thermodynamics. 

In  the  study  of  vapor-liquid  equilibria,  one  of  the  most  important 
of  the  basic  relations  states  that  at  equilibrium  the  fugacity  of  a  given 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  33 

component  is  the  same  in  all  phases.    Thus, 

fy  =  /*  (3-6) 

where  /  is  the  f ugacity ,  F  refers  to  the  vapor  phase,  and  L  refers  to  the 
liquid  phase.  This  same  equation  applies  to  each  component  in  the 
mixture  at  equilibrium.  To  be  useful  for  distillation  calculations, 
the  fugacity  terms  in  Eq.  (3-6)  must  be  related  to  the  compositions  of 
the  phases.  For  isothermal  changes  in  conditions  at  constant  composi- 
tion, the  change  in  fugacity  is  evaluated  by  the  following  equation : 

RT  d  In  /  =  v  dp  (3-7) 

where  R  =  gas  law  constant 

T  =  absolute  temperature 
/  =  fugacity 

v  =  molal  volume,  i.e.,  volume  of  1  mol  of  substance  under 
consideration  =  partial  molal  volume  for  a  component  in  a 
mixture 
p  =  pressure 

For  mixtures  that  obey  the  perfect-gas  law  this  relationship  makes 
the  fugacity  of  a  component  in  the  vapor  proportional  to  the  partial 
pressure  of  the  component,  i.e.,  the  mol  fraction  of  the  component 
times  the  total  pressure.  It  is  customary  to  define  the  standard  state 
for  fugacity  such  that  it  is  numerically  equal  to  the  pressure  for  a  per- 
fect gas;  i.e.,  the  proportionality  factor  is  made  equal  to  unity. 

The  fugacity  of  a  component  in  an  ideal  liquid  is  defined  as  propor- 
tional to  the  mol  fraction  of  that  component  in  the  liquid  times  the 
fugacity  of  the  pure  component  under  the  same  temperature  and  pres- 
sure as  the  mixture.  Thus  for  a  perfect  gas  and  an  ideal  solution, 

fv  =  yi*  =  JL  =  foxi  (3-8) 

where  y\  =  mol  fraction  of  component  1  in  vapor 
xi  =  mol  fraction  of  component  1  in  liquid 
TT  =  total  pressure  of  mixture 
/£  =  fugacity  exerted  by  the  pure  liquid  at  temperature  T  under 

total  pressure,  w 

This  equation  is  not  identical  with  Raoult's  law,  and  /£  is  not  in 
general  equal  to  PI,  the  vapor  pressure  of  pure  liquid  1  at  temperature 
r,  because  the  liquid  is  under  a  pressure  different  from  Pi.  This 
difference  can  be  evaluated  by  Eq.  (3-7). 

(3'9) 


34  FRACTIONAL  DISTILLATION 

where  v  =  partial  molal  volume  of  liquid  1 

PI  «  vapor  pressure  of  pure  component  1  at  temperature  T 
R  =  gas-law  constant 

fpi  =  fugacity  of  component  1  at  temperature  T  and  pressure  PI 
Over  the  pressure  range  usually  involved,  v  is  essentially  constant  and 

f*i  =  /pi«'(r-pl)/*r  (3-10) 

Equations  (3-8)  and  (3-10)  differ  from  Raoult's  law  by  the  expo- 
nential term,  and  fpi  instead  of  PI.  These  corrections  are  frequently 
small  and  can  be  neglected.  However,  in  high-pressure  equilibria  the 
corrections  become  large,  and  even  an  ideal  solution  would  not  be 
expected  to  obey  Raoult's  law  because  (1)  the  .components  in  the  liquid 
mixture  at  a  given  temperature  are  under  a  different  total  pressure 
than  they  would  be  as  pure  components  and  (2)  the  fugacity  of  the  pure 
liquid  is  not  equal  to  its  vapor  pressure.  An  equation  similar  to  (3-8) 
applies  to  each  of  the  components  in  an  ideal  mixture. 

Most  mixtures  do  not  obey  Raoult's  law  or  the  corrected  Raoult's 
law  given  by  Eqs.  (3-8)  and  (3-10).  The  deviations  from  the  ideal 
solution  laws  can  be  due  to  the  vapor  phase,  the  liquid  phase,  or  both. 
These  deviations  are  both  chemical  and  physical  in  nature.  The  most 
important  factors  involved  in  these  deviations  are  believed  to  be  (1) 
the  fact  that  the  molecules  have  volume  and  (2)  the  fact  that  the  mole- 
cules exert  forces  on  each  other  that  may  be  attractions,  or  repulsions, 
or  actual  chemical  effects. 

The  problem  for  such  mixtures  becomes  one  of  relating  the  fugacities 
of  the  components  in  the  vapor  and  the  liquid  to  the  composition  and 
the  physical  properties  of  the  components,  and  it  has  been  found  desir- 
able to  consider  separately  the  deviations  in  the  liquid  and  the  vapor 
phase. 

VAPOR  PHASE 

Pure  gases  or  gaseous  mixtures  approach  agreement  with  the  perfect- 
gas  law  if  the  pressures  are  low  enough. l  Under  these  conditions,  the 
volume  of  the  gaseous  mixture  is  so  large  that  the  volume  of  the  mole-, 
cules  is  a  negligible  percentage  of  the  total,  and  the  molecules  on  the 

1  The  reduction  to  zero  pressure  may  cause  a  change  in  species.  Thus,  if  the 
pressure  on  pure  CU  gas  is  reduced,  the  gas  may  never  obey  the  perfect-gas  law 
since,  before  exact  agreement  is  reached,  the  CU  will  dissociate  to  atomic  chlorine. 
The  atomic  chlorine  should  agree  with  the  perfect-gas  law  at  zero  pressure,  and  for 
all  practical  engineering  purposes  diatomic  chlorine  gas  agrees  with  the  gas  laws  at  a 
pressure  less  than  one-tenth  of  an  atmosphere,  and  under  these  conditions,  the 
dissociation  to  atomic  chlorine  is  negligible  at  moderate  temperatures. 


CALCULATION  OF  VAPOR-LIQUID  EQVlLt&RtA  35 

average  are  so  far  apart  that  the  forces  between  them  are  small.  As 
the  pressure  increases,  the  effect  of  the  volume  of,  and  the  attraction 
between,  the  molecules  become  so  great  that  deviations  from  the  ideal 
solution  laws  become  large  and  it  is  necessary  to  employ  the  fugacity 
instead  of  the  pressure  in  vapor-liquid  relationships.  The  isothermal 
change  in  fugacity  of  a  mixture  of  given  composition  can  be  calculated 
by  integrating  Eq.  (3-7).  In  this  integration  the  absolute  value  of  the 
fugacity  at  any  given  state  can  be  arbitrarily  chosen.  Defining  the 
fugacity  equal  to  the  pressure  for  a  perfect  gas,  Eq.  (3-7)  can  be 
rearranged  as  follows: 

RTdlnf=vdp  (3-11) 

RTdlnt^  (va-v^dp  (3-12) 

where  va  =  actual  molal  volume 

vt  =  perfect-gas  law  molal  volume 

Fugacity  of  Pure  Gases.  In  order  to  utilize  Eq.  (3-12)  it  is  necessary 
to  have  information  on  the  actual  molal  volume  as  a  function  of  the 
pressure  at  the  temperature  in  question.  The  lack  of  these  data  limits 
the  utility  of  this  equation.  However,  it  has  been  found  possible  to 
develop  correction  factors  to  the  perfect-gas  law  that  will  apply  to 
almost  all  gases  and  gaseous  mixtures.  One  method  of  representing 
these  deviation  factors  is  to  plot  the  compressibility  factor  ju,  which  is 
equal  to  PV/RT,  as  a  function  of  the  reduced  pressure  at  constant 
reduced  temperature,  where  reduced  pressure  PR  is  the  pressure  divided 
by  the  critical  pressure,  and  reduced  temperature  TR  is  the  absolute 
temperature  divided  by  the  absolute  critical  temperature.  Such  a 
plot  is  given  in  Fig.  3-4;  similar  plots  for  higher  temperatures  and 
pressures  are  available. 

These  plots  give  good  agreement  with  the  experimental  data  for 
most  vapors  with  the  exception  of  hydrogen  and  helium,  both  of  which 
have  low  critical  temperatures  and  pressures.  It  has  been  found  possi- 
ble to  use  the  plot  for  these  two  gases  by  using  modified  critical  con- 
stants. However,  neither  gas  is  particularly  important  in  vapor- 
liquid  equilibria. 

Using  the  p,  factor  correction,  the  actual  molal  volume  va  becomes 
When  this  is  substituted  in  Eq,  (3-12),  it  gives 


(3-13) 
or 

rfln£-(f.-l)&  (3-14) 

P  P 


36 


FRACTIONAL  DISTILLATION 


If  this  equation  is  integrated  from  zero  pressure  up  to  the  pressure  in 
question,  the  limits  on  the  left-hand  side  are  from  1.0  to  the  ratio  of 
fugacity  to  pressure  at  the  pressure  in  question.  Thus  In  (f/p)  equals 
the  integral  from  zero  to  p  of  (M  —  1)  dp  /p. 

Equation  (3-14)  can  be  modified  and  dPR/PR  used  instead  of  dp/p. 
The  limits  of  integration  then  become  from  zero  to  PR.  At  constant 
temperature,  it  is  therefore  obvious  that  the  integral  is  a  function  only 
of  PR,  T/z,  and  the  variables  determining  ju.  If  the  M  values  are  a  func- 


P-V-T  RELATIONS  FOR  VAPORS 
BELOW  THE  CRITICAL 
PR=  P/  Pc=  Reduced  Pressure  - 
TR =T/VReduced  Temperature 
R= 6ois  Constant 


065  TR=07TR=015TR*08TR-085  TR-0.9     TR 


02      0.3      0.4      0.5      06      07 


FIG.  3-4. 

tion  of  PR  and  TR  only,  the  ratio  of  fugacity  to  the  pressure  is  also  a 
unique  function  of  the  reduced  pressure  and  the  reduced  temperature. 
Integration  using  the  ju  plots  have  been  made,  and  one  method  of 
presentation  is  given  in  Fig.  (3-5). 

With  an  accuracy  suitable  for  engineering  purposes  these  plots  make 
it  possible  to  calculate  the  fugacity  of  a  pure  gas  at  any  temperature 
and  pressure,  assuming  that  the  critical  constants  of  the  gas  are  known. 
Even  in  cases  where  the  critical  constants  are  not  known,  highly  satis- 
factory methods  have  been  developed  for  estimating  these  constants. 

Fugacity  of  Mixtures.  When  applying  the  n  plot  and  the  fugacity 
plot  to  mixtures,  the  question  arises  as  to  the  proper  values  to  be 
employed  for  the  critical  temperature  and  the  critical  pressure.  Mix- 
tures have  critical  temperatures  and  critical  pressures,  but  it  has  been 
found  that  these  values  do  not  give  satisfactory  results  when  used  for 
calculating  reduced  temperatures  and  pressures  to  be  used  with  the 
charts,  but  pseudocritical  constants  can  be  calculated  which  give  better 
agreement.  For  these  pseudo  constants  one  of  the  best  methods  of 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


37 


calculation  is  the  mol  fraction  average;  i.e.,  the  calculated  pseudo- 
critical  temperature  is  equal  to  the  sum  of  the  products  of  the  mol  frac- 
tion times  the  critical  temperature  for  each  of  the  pure  components. 
The  pseudocritical  pressure  is  calculated  in  an  analogous  manner. 
When  these  values  are  employed  for  calculating  the  reduced  tempera- 


0.2      0.4      06      08      10 


1.2      14      16      1.8      2.0 
Pfl  *  Reduced  pressure 

FIG.   3-5      Fugacities  of  hydrocarbon    vapors. 


26 20 50 


ture  and  pressure,  satisfactory  agreement  is  attained  with  /*  charts. 
It  should  be  emphasized  that  these  calculated  values  are  not  true 
critical  constants  for  the  mixture. 

Fugacity  of  a  Component  in  a  Mixture.  In  vapor-liquid  equilibria, 
vapor  mixtures  are  generally  involved,  and  Eq.  (3-6)  requires,  not  the 
fugacity  of  the  mixture,  but  the  fugacity  of  the  components  in  the 
mixture.  In  order  to  estimate  the  fugacity  of  the  components  in  the 
mixture,  Lewis  and  Randall  (Ref.  16)  suggested  that  this  fugacity 
was  equal  to  the  mol  fraction  of  the  component  in  the  mixture  times 
the  fugacity  of  the  pure  component  at  the  same  temperature  and  total 
pressure  as  the  mixture.  Thus, 

fvi  «  yi/n  (3-15) 

where  fvi  =  fugacity  of  component  in  vapor 

y  =  mol  fraction  in  vapor 

ffl  =  fugacity  of  pure  component  at  same  temperature  and 
pressure  as  mixture 


38  FRACTIONAL  DISTILLATION 

The  value  of  f*  can  be  estimated  from  the  fugacity  plots,  and  this 
fugacity  rule  has  been  widely  used  in  vapor-liquid  and  other  equilib- 
rium calculations.  It  can  be  shown  that  this  is  true  if  the  vapor  mix- 
ture obeys  Amagat's  law,  which  states  that  the  volume  of  a  mixture  is 
equal  to  the  sum  of  the  volumes  of  the  pure  components  when  these 
are  measured  at  the  same  temperature  and  total  pressure  as  the  mix- 
ture. Experimental  data  have  shown  this  rule  to  be  a  reasonable 
approximation  if  the  pressures  are  not  too  high;  at  high  pressures,  large 
deviations  are  found  to  occur.  However,  in  essentially  all  cases  it  is 
better  than  Dalton's  law. 

A  large  number  of  additional  methods  of  predicting  the  fugacity  of  a 
component  in  a  gaseous  mixture  have  been  proposed.  In  general  these 
have  been  based  on  different  rules  for  the  PVT  relations  of  mixtures 
used  to  evaluate  the  fugacity.  None  of  them  has  been  satisfactory  in 
all  cases  although  some  of  them  are  better  than  the  Lewis  and  Randall 
rule,  but  they  are  more  difficult  to  employ,  and  a  number  of  them 
require  more  experimental  information  than  is  usually  available.  This 
complexity  has  greatly  limited  their  use. 

The  Lewis  and  Randall  fugacity  rule  can  generally  be  used  with 
reasonable  accuracy  for  vapor  mixtures  up  to  pressures  approximately 
one-half  of  the  critical  pressure;  i.e.,  300  to  500  p.s.i.g.  in  the  case  of 
hydrocarbons.  This  is  only  an  approximate  limit,  but  large  errors  will 
not  generally  be  encountered  at  values  of  PR,  based  on  the  pseudo- 
critical  pressure  of  the  mixture,  less  than  one-half.  If  the  components 
in  the  mixture  are  of  similar  types,  i.e.,  02  and  Ns,  ethylene  and  ethane, 
benzene  and  toluene,  etc.,  satisfactory  results  can  often  be  obtained  for 
values  of  PR  as  high  as  0.9.  The  rule  is  not  satisfactory  at  the  critical 
condition.  Figure  3-5  gives  values  of  the  fugacity  for  conditions  where 
a  pure  component  cannot  exist  as  a  vapor.  These  values  were  obtained 
by  using  the  Lewis  and  Randall  rule  with  actual  vapor-liquid  equilibria 
to  calculate  the  fugacity  values. 

Instead  of  using  the  reduced  correlations,  it  is  thermodynamically 
possible  to  calculate  the r  true  fugacity  of  a  component  in  a  gaseous 
mixture,  if  sufficient  PVT  data  are  available.  The  methods  of  cal- 
culation are  very  laborious  and  require  PVT  data  of  high  pre- 
cision. This  method  has  not  had  any  significant  engineering  use 
owing  to  the  great  difficulty  in  obtaining  the  extensive  PVT  data 
necessary  and  the  work  involved  in  making  the  calculations  from 
such  data.  Its  chief  value  is  furnishing  an  exact  basis  to  evaluate  the 
empirical  rules. 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  39 

LIQUID  PHASE 

Deviations  from  ideal  solution  laws  are  more  important  for  the 
liquid  phase  than  those  for  the  vapor  phase  because  they  are  encoun- 
tered even  at  low  pressures,  and  in  general  their  magnitudes  are 
greater.  The  densities  of  the  liquids  are  such  that  the  volume  of  the 
molecules  and  the  forces  between  them  are  always  significant.  Devi- 
ations for  the  liquid  phase  are  of  at  least  two  main  types:  (1)  Those  due 
to  the  fact  that  the  vapor  does  not  obey  the  perfect-gas  law.  Thus,  if 
one  tries  to  define  the  fugacity  of  a  component  in  the  liquid  phase  as 
mol  fraction  times  the  vapor  pressure  of  the  pure  component,  this  can 
be  satisfactory  only  when  the  vapor  at  a  pressure  equal  to  the  vapor 
pressure  is  essentially  a  perfect  gas.  (2)  The  deviations  that  are  due 
to  special  phenomena  associated  with  the  liquid  phase  such  as  associa- 
tions or  chemical  combinations. 

Gas  Law  Deviation.  In  the  ideal  solution  the  partial  pressure,  or 
activity,  of  a  component  was  equal  to  the  mol  fraction  times  the  vapor 
pressure  of  the  pure  component  at  the  temperature  in  question.  If 
the  pressures  are  such  that  the  vapor  under  these  conditions  does  not 
obey  the  perfect-gas  law,  then  a  fugacity  correction  should  be  applied. 
For  such  a  case,  the  Lewis  and  Randall  rule  would  be 

/LI  =  *i/*i  (3-16) 

where  fLi  =  fugacity  of  component  1  in  liquid  phase 

Xi  =  mol  fraction  of  component  1  in  liquid  phase 
/*!  =  fugacity  of  pure  liquid  component  1  at  temperature  and 

pressure  of  mixture 

In  general  the  total  pressure  is  different  from  the  vapor  pressure  of 
the  pure  components  at  the  same  temperature,  and  /*t  is  not  equal  to 
the  fugacity  of  the  pure  liquid  under  its  own  vapor  pressure,  fpi. 
They  are  related: 


or,  assuming  v  =  constant, 

&  -  fPie't*-™'RT  (3-17) 

where  ir  =  total  pressure 

v  =  partial  molal  volume  of  component 
Pi  =  vapor  pressure  of  component  at  temperature  of  mixture 
T  «  absolute  temperature 
R  =  gas  law  constant 


40  FRACTIONAL  DISTILLATION 

From  equilibrium  consideration,  the  fugacity  of  the  pure  liquid 
under  its  own  vapor  pressure,  fpi,  is  equal  to  the  fugacity  of  the  satu- 
rated vapor  at  the  same  temperature  and  pressure.  Thus,  Fig.  3-5 
can  be  used  to  evaluate  fp,  but  it  should  be  emphasized  that  the  reduced 
pressure  is  calculated  at  the  vapor  pressure  of  the  pure  component 
instead  of  at  the  total  pressure.  This  application  of  the  fugacity  to 
the  liquid  phase  corrects  for  the  fact  that  the  vapor  is  not  a  perfect  gas, 
but  it  does  not  correct  for  the  special  phenomena  associated  with  the 
liquid  phase. 

Combining  Eqs.  (3-15),  (3-16),  and  (3-17)  gives 

yiAi  =  xi/pie**-™**  (3-18) 

In  using  Eq.  (3-18)  certain  difficulties  are  encountered.  Considering 
the  vapor  phase,  the  temperature  and  pressure  of  the  mixture  usually 
are  intermediate  between  those  of  the  pure  components,  and  it  is  often 
found  that  the  temperature  and  pressure  are  such  that  one  of  the  pure 
components  at  these  conditions  would  be  superheated  vapor  and  the 
other  a  supersaturated  vapor.  The  superheated  vapor  offers  no  diffi- 
culty since  it  is  easily  possible  to  obtain  PVT  data  for  superheated 
vapors,  and  this  type  of  information  was  used  to  develop  the  /*  plots. 
However,  it  is  essentially  impossible  to  obtain  PVT  data  on  super- 
saturated vapors  since  they  tend  to  condense  easily.  In  the  liquid 
phase  such  difficulties  are  not  encountered  at  moderate  temperatures. 
If  the  temperature  is  not  too  high,  the  vapor  pressure  of  the  compo- 
nents can  be  obtained  and  the  fugacity  can  be  determined.  However, 
at  higher  temperatures,  it  is  possible  for  the  temperature  of  the  mix- 
ture to  be  greater  than  the  critical  temperature  of  one  of  the  com- 
ponents and  still  have  this  component  present  in  the  liquid  phase  in 
large  amounts.  The  calculation  of  fp  for  such  conditions  is  complicated 
because  data  are  not  available  on  the  vapor  pressure  of  a  liquid  above 
the  critical  temperature. 

Empirical  rules  have  been  developed  to  handle  these  fugacity  diffi- 
culties for  the  vapor  and  liquid.  In  the  case  of  the  liquid  phase,  it  has 
been  customary  to  plot  the  logarithm  of  the  vapor  pressure  vs.  the 
reciprocal  of  the  absolute  temperature.  It  is  well  known  that  such 
plots  are  remarkably  straight,  and  for  fugacity  calculations  the  straight 
line  has  been  extrapolated  past  the  critical  to  higher  temperatures. 
Such  extrapolations  have  given  satisfactory  results.  In  the  case  of 
the  vapor,  the  problem  is  more  complex,  and  the  main  solution  has 
been  to  calculate  /*  by  Eq.  (3-18),  using  the  experimental  data  and  the 
estimated  value  of  /p.  This  has  been  done  for  a  number  of  mixtures, 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 
TABLE  3-2.     K  VALUES 


41 


Temp., 
°F. 

Absolute  pressures,  atm. 

1 

2 

3 

4 

5 

6 

8 

10 

20 

30 

40 

50 

CIU 


-90 

41 

21 

14' 

10  5 

8  5 

7  1 

5  4 

4  3 

2  3 

1  7 

1  3 

1  2 

-20 

83 

42 

28 

21 

17 

14 

10  6 

8  6 

4  4 

3  0 

2  3 

2  0 

20 

115 

57 

38 

29 

23 

19 

14  5 

11  5 

6  0 

4  1 

3  1 

2  6 

100 

190 

96 

64 

48 

38 

32 

24 

19 

9  7 

6  6 

5  0 

i.l 

200 

300 

150 

100 

76 

61 

51 

39 

31 

15  5 

10  3 

7.8 

6.2 

400 

580 

290 

195 

145 

116 

98 

73 

58 

29 

19.5 

14.5 

12.0 

-60 

7.7 

3  9 

2  6 

2  0 

1  6 

1  4 

1.1 

0  9 

0  51 

0  42 

0  38 

0  3( 

20 

24  5 

12  3 

8  3 

6  3 

5  1 

4  2 

3  2 

2  6 

1  44 

1  08 

0  92 

0  & 

100 

51 

25  5 

17  2 

13  0 

10  2 

8  8 

6  5 

5  3 

2  8 

2.0 

1  6 

1  31 

200 

103 

51 

34  5 

26 

20  5 

17  2 

13  0 

10  4 

5  3 

3  7 

2  8 

2  31 

300 

172 

87 

68 

44 

35 

29 

22 

17  5 

8  9 

6  0 

4  7 

3  9 

400 

265 

132 

88 

66 

54 

45 

34 

27 

14 

9  2 

7  0 

5  6 

C2Hft 


*-60 

4  8 

2  4 

1.6 

1  3 

1  0 

0  88 

0  69 

0  58 

0  36 

0  29 

0  26 

0.20 

20 

16  0 

8  0 

6  4 

4  1 

3  3 

2  8 

2  1 

1  75 

1  0 

0  79 

0  69 

0  66 

100 

34 

17 

11  7 

8  8 

7  1 

5  9 

4  5 

3  7 

2  0 

1  45 

1  18 

1.0 

200 

71 

36 

24 

18 

14  5 

12  1 

9  2 

7  4 

3  8 

2  7 

2.1 

1  75 

300 

123 

62 

42 

32 

25  5 

21 

16 

13 

6  6 

4  5 

3  4 

2  8 

400 

195 

98 

66 

50 

40 

33 

25 

20 

10 

6  9 

5  2 

4  3 

-60 
20 

'0  79 
4  5 

0.35 
2  3 

0  27 
1  55 

0  21 
1  2 

0  17 
0.95 

0  15 
0  82 

0  12 
0  64 

0  10 
0  54 

0  07 
0  34 

0  06 
0  28 

0  05 
0  26 

0  05 
0  26 

100 

12  2 

6  3 

4  3 

3  3 

2  G 

2  2 

1  7 

1  4 

0  82 

0  06 

0  59 

0.68 

200 

30 

15 

10 

7  7 

6  2 

5  2 

4  0 

3  2 

1  75 

1  3 

1  1 

0  98 

300 

59 

30 

20 

15 

12 

10 

7  7 

6  2 

3  2 

2  3 

1  8 

1  5 

400 

100 

50 

33 

25 

20 

16 

12  5 

10  0 

5  2 

3.6 

2  8 

2.3 

-60 
20 
100 
200 
300 
400 

0  67 
3  7 
11 
25 
51 
85 

0  34 
1  9 
5  5 
13 
25 
43 

0  23 
1  3 
3  7 
8  6 
17 
29 

0  18 
0  98 
2  8 
6  5 
13 
22 

0  15 
0  8 
2  3 
5  2 
10 
17 

0  13 
0  69 
1.9 
4  4 
8.7 
14 

0  10 
0.54 
1  5 
3.4 
6.5 
11 

0  09 
0  45 
1.2 
2  8 
5  3 
8  8 

0  06 
0  29 
0  75 
1  5 
2  8 
4  5 

0.05 
0  24 
0,6 
1  1 
2.0 
3.1 

0  05 
0  23 
0  54 
0  95 

1  6 
2  4 

0  06 
0  24 
0  54 
0.85 
1  3 
2.0 

t-CJIic 


-60 
20 
100 
200 
400 
600 

0  18 
1  2 
4  5 
13 
51 
125 

0  09 
0  63 
2  3 
6  7 
25 
63 

0  07 
0  43 
1  6 
4  5 
17 
42 

0.05 
0  34 
1  2 
3  5 
13 
32 

0  04 
0  28 
1.0 
2  8 
10  5 
25.5 

0  04 
0  24 
0.86 
2.4 
8  8 
21  5 

0  03 
0  19 
0  67 
1  8 
6.6 
16 

0  03 
0  16 
0  57 
1  5 
5  3 
13 

0  02 
0  11 
0  36 
0  91 
2  9 
6.6 

0.02 
0.10 
0,31 
0  74 
2  0 
4.5 

0  02 
0  09 
0  30 
0  68 
1  6 
3.4 

0  02 
0  11 
0  32 
0  67 
1  4 
2  9 

42 


FRACTIONAL  DISTILLATION 
TABLE  3-2.     K  VALUES     (Continued) 


Temp,, 

oy 

Absolute  pressures,  atm. 

1 

2 

3 

4 

5 

6 

8 

10 

20 

30 

40 

60 

-60 

0  10 

0  05 

0  03 

0  03 

0  02 

0  02 

0  02 

0  01 

0  01 

0  01 

0  01 

0  01 

20 

0  81 

0  43 

0  29 

0  23 

0  19 

0  16 

0  13 

0  11 

0  07 

0  07 

0  07 

0  08 

100 

3  4 

1  7 

1  2 

0  9 

0  73 

0.63 

0  5 

0  43 

0  28 

0  24 

0  24 

0  26 

200 

10  5 

5.4 

3  7 

2  8 

2  3 

1.9 

1.5 

1  3 

0  77 

0  64 

0  61 

0  61 

400 

43 

21 

14 

11 

8  8 

7.3 

5  5 

4.5 

2  4 

1  7 

1  4 

1  3 

600 

no 

54 

36 

27 

22 

18 

14 

11 

5.6 

3  9 

3  0 

2  5 

-20 

0  09 

0.05 

0  03 

0  02 

0  02 

0  02 

0  01 

0  01 

0  01 

0  01 

0  01 

0  01 

60 

0  63 

0.32 

0  22 

0.17 

0  14 

0  13 

0.10 

0  08 

0  06 

0  06 

0.06 

0  07 

100 

1  3 

0  67 

0  47 

0  36 

0  30 

0  26 

0  21 

0  18 

0  12 

0  11 

0  12 

0  14 

200 

5.2 

2  7 

1  8 

1  4 

1  2 

1  0 

0  77 

0  65 

0  43 

0  37 

0.37 

0  40 

400 

25.5 

13 

8.8 

6  7 

5  4 

4  5 

3  4 

2  8 

1  6 

1  2 

1  0 

0  94 

600 

69 

34 

23 

17 

14 

12 

8  7 

7.0 

3.7 

2  6 

2  1 

1  72 

-20 

0  05 

0  03 

0  02 

0  02 

0.01 

0  01 

0.01 

0  01 

60 

0.47 

0  24 

0  17 

0  13 

0,11 

0  09 

0  07 

0  06 

0  05 

0  04 

0  04 

0  05 

100 

1.0 

0.53 

0  36 

0  28 

0  23 

0  20 

0  16 

0  14 

0  10 

0  09 

0  09 

0  10 

200 

4.3 

2.2 

1  5 

1  2 

0  96 

0  83 

0  65 

0  55 

0  36 

0  32 

0  32 

0  35 

400 

22  5 

12 

7.8 

5  9 

4  8 

4  0 

3  1 

2  5 

1  4 

1.1 

0  93 

0  89 

600 

65 

33 

22 

17 

13  4 

11  2 

8  5 

6.9 

3  6 

2  6 

2  0 

1  7 

n-Cellu 


60 

0  13 

0  07 

0  05 

0  04 

0  03 

0  02 

0  02 

0  02 

0  01 

0  01 

0  02 

0  02 

100 

0.35 

0.18 

0  13 

0  10 

0  08 

0  07 

0  06 

0  05 

0  04 

0  04 

0  04 

0  05 

200 

1.9 

0.97 

0  67 

0  51 

0  43 

0  37 

0  30 

0  26 

0  18 

0  17 

0  18 

0  22 

400 

13  6 

7  0 

4  7 

3.6 

2  9 

2  5 

1  9 

1  6 

0  98 

0  80 

0  77 

0  77 

600 

42 

21 

14 

11 

8.6 

7  2 

5.5 

4.4 

2.4 

1.8 

1  5 

1  3 

100 

0  11 

0  05 

0  04 

0.03 

0  03 

0.02 

0  02 

0.02 

0.01 

0  01 

0  02 

0.02 

200 

0.83 

0.43 

0.29 

0  23 

0  20 

0  17 

•0.14 

0.12 

0  09 

0,09 

0.10 

0  14 

400 

8.4 

4  3 

2  9 

2  2 

1  8 

1.6 

1  2 

1  0 

0  67 

0.60 

0.60 

0  67 

600 

29 

15 

10 

7.6 

6.1 

5.1 

3.9 

3.2 

1.75 

1.3 

1.1 

1.05 

n-CsHis 


100 

0.04 

0  02 

0  01 

0  01 

0  01 

200 

0  39 

0  20 

0.14 

0  11 

0.09 

0  08 

0.07 

0,06 

0.04 

0.05 

0,06 

0.09 

400 

5,3 

2.6 

1.8 

1.4 

1.15 

1.0 

0.78 

0  67 

0  47 

0  45 

0  49 

0.57 

600 

19.5 

10 

6.7 

5.1 

4.2 

3.5 

2.7 

2.25 

1  3 

1.1 

0.95 

0.91 

CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  43 

and  these  calculated  values  for  the  supersaturated  region  have  been 
plotted  on  the  fugacity  plot  and  used  as  a  basis  for  extrapolating  the 
curves  based  on  PVT  relations.  In  most  cases,  the  extrapolation  is 
not  large,  and  the  deviations  should  not  be  too  great. 

When  equilibrium  calculations  are  to  be  made  a  number  of  times  for 
the  same  components,  it  is  convenient  to  simplify  the  equation  and 
express  it  the  form  of  ^i  =  Kx\  where 

K  ~  fjp  (3-19) 

J* 

K  values  based  on  fugacity  calculations  have  been  prepared  for  the 
common  lower  molecular  weight  petroleum  hydrocarbons  and  are 
presented  in  Table  3-2.  On  the  whole,  they  are  equivalent  to  results 
obtained  with  Eq.  (3-18)  and  the  fugacity  plots,  but  they  are  simpler 
to  use  in  vapor-liquid  calculations.  If  such  values  have  not  been  cal- 
culated for  the  components  in  question,  it  is  necessary  to  employ  the 
fugacity  plot. 

Equation  (3-18)  includes  the  exponential  correction  term,  and  this 
term  is  appreciable  at  high  pressure.  However,  it  has  been  found  that 
in  hydrocarbon  systems  up  to  about  GOO  p.s.i.  the  agreement  is  about 
the  same  whether  or  not  this  factor  is  included.  This  is  probably 
because  the  values  of  ff  in  Fig.  3-5  for  the  supersaturated  region  were 
evaluated  from  experimental  data  without  the  use  of  the  exponential 
term.  If  an  exact  method  of  computing  the  fugacity  of  the  com- 
ponents were  available,  the  exponential  correction  should  be  applied. 
However,  in  using  values  from  this  figure  it  is  suggested  that  it  be  neg- 
lected for  distillation  calculations. 

Equation  (3-18)  has  been  found  to  give  good  agreement  with  the 
experimental  data  on  mixtures  of  similar  compounds  up  to  pressures 
equivalent  to  a  reduced  pressure  of  about  0.5.  It  has  been  extremely 
useful  in  high-pressure  calculations  associated  with  petroleum  mix- 
tures. The  vapor-liquid  equilibrium  data  for  mixtures  of  oxygen  and 
nitrogen  are  well  correlated  by  the  equation.  Basically,  the  equation 
should  be  useful  for  any  mixtures  that  agree  reasonably  well  with 
Raoult's  law  at  low  pressure.  At  pressures  greater  than  PR  «  0.5,  the 
agreement  is  not  so  good,  but  it  is  still  a  useful  approximation.  How- 
ever, this  method  of  calculation  breaks  down  completely  in  the 
neighborhood  of  the  critical  region  of  the  mixtures.  This  condition 
will  be  considered  in  a  later  section. 

The  fugacity  equation  can  be  applied  to  each  component  in  the  mix- 


44 


FRACTIONAL  DISTILLATION 


ture.     It  is  suitable  for  multicomponent  mixtures  as  well  as  for  binary 
mixtures. 

Example  Illustrating  the  Use  of  Fugacity  Corrections.  Scheeline  (Ref.  23)  has 
studied  the  vapor-liquid  equilibrium  of  propane-isobutylene  mixtures  at  high 
pressures.  Using  the  data  and  notes  given  below,  make  a  y,x  plot  for  a  total 
pressure  of  400  p.s.i.a.  showing 

1.  Experimental  data. 

2.  Curve  calculated  by  Raoult's  law. 

3.  Curve  calculated  by  using  fugacity  corrections. 

CRITICAL  CONSTANTS 


PC,  p.s.i.a. 

Tc,  °F, 

Cz 

632 

209 

c, 

580 

291 

1.  Experimental  data: 


Absolute 
pressure, 

Temp., 

xct 

yc9 

p.s.i.a. 

400 

242 

0  086 

0  140 

400 

222 

0  286 

0  413 

400 

206 

0.438 

0.572 

400 

190 

0.648 

0.754 

400 

184 

0  734 

0.827 

400 

175 

0.847 

0.895 

Solution.     Calculate  the  y,x  values  at  400  p.s.i.a.  and  200°F. 

2.  Raoult's  and  Dalton's  laws: 

ZsPa  4-  £4P4  =  400 

400j-^P4 

At  200°F.  from  Table  3-3  P3  -  569  p.s.i.a.,  P4  -  232  p.s.i.a.,  x*  «  0.498,  and 
yz  **  0.498  (569/400)  -  0.709. 

3.  Fugacity  corrections : 

2/s/7r3  ••  Xzfpt        x\  —  \  —  Xz 

2/4/7T4    =   ^4/p4  2/4    —    1—2/3 

Combining  and  solving, 

/7T4   —  /P4 


.  0.88; 
«  0.986; 


=»  0.403;        Km  **  40%so  «*  0.69 
-  0.90;       -  *•*»  -  40%32  -  0.633 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 
From  Fig.  3-5, 

(fs\    =  0.665  (&\    =  0.755 

VP/3  V   /S 

(&}    -0.78  (^    -0.6 

VP/4  \7T   /4 


45 


.64 


As 

/P4 


0.665  X  569  -  379 
0.78  X  232  -  181 


181 


-  0.755  X  400  -  302 
=  0.64  X  400  -  256 

0.53 


2/3     =    — 


302 
0.53(379) 


302 


-  181 
-  0.665 


In  a  similar  manner,  the  data  in  Tables  3-3  and  3-4  were  calculated. 

TABLE  3-3.     RAOULT'S  LAW 
(400  p.s.i.a.) 


T,  °F. 

P4,  p.s.i.a. 

PS,  p.s.i.a. 

#3 

2/3 

255 

399 



0 

0 

237 

305 

730 

0.224 

0.409 

220 

287 

688 

0.282 

0.485 

200 

232 

569 

0.498 

0  709 

180 

187.5 

470 

0.753 

0.885 

163 

150 

400 

1.0 

1.0 

TABLE    3-4    FUGACITY  CALCULATION 


T,  °R. 

697 

680 

660 

640 

623 

TRS 

1.041 

1.017 

0.986 

0.957 

0.931 

P3 

730 

688 

569 

470 

400 

PR3 

1.16 

1.09 

0.90 

0.744 

0.633 

(A/p)3 

0  64 

0.65 

0.665 

0.68 

0.71 

AS 

466 

447 

379 

319 

284 

(U  M  3 

0.8 

0.78 

0  755 

0.74 

0.71 

/,3 

320 

312 

302 

296 

284 

TRI 

0.929 

0.906 

0.88 

0.853 

0  830 

P4 

305 

287 

232 

187.5 

150 

P*4 

0.526 

0.495 

0.40 

0.324 

0.258 

(fp/P)< 

0.76 

0.77 

0.78 

0.79 

0.82 

/P4 

232 

221 

181 

148 

123 

(ArA)4 

0.68 

0.655 

0.64 

0.605 

0.59 

/r4 

272 

262 

256 

242 

236 

a;3 

0.244 

0.266 

0  53 

0.832 

1.0 

2/3 

0.356 

0.381 

0.665 

0.896 

1.0 

46 


FRACTIONAL  DISTILLATION 


The  experimental  data  and  the  calculated  results  are  shown  in  Fig.  3-6.  It  will 
be  noted  that  the  fugacity  calculations  for  the  ytx  values  are  in  excellent  agreement 
with  the  experimental  results,  but  that  Raoult's  and  Dalton's  laws  give  values  of 
the  vapor  composition  that  are  much  too  high.  For  average  relative  volatility 
these  laws  give  2.45;  fugacity,  1.77;  and  experimental,  1.71. 

Solution  Deviations.  The  corrections  so  far  considered  have  been 
limited  to  those  associated  with  the  fact  that  the  vapor  does  not  obey 
the  perfect-gas  law.  A  large  number  of  mixtures,  in  fact  most  of 
them,  do  not  obey  the  ideal  solution  laws  even  at  very  low  pressure, 
and  the  deviations  cannot  be  predicted  by  the  use  of  gas-phase  fugacity 
corrections.  Thejleyj^c^jt^ejbhe  result  of  the  forces  between  the 
molecules  in  the  liquid  phase,  and  these  forces  can  Ee^veryTargely^ue 


IJU 

0.6 
0.6 
0.4 
02 

0 

^ 

£ 

/\ 

^/ 

y 

/ 

/ 

/  / 

/ 

/ 

// 

/ 

/ 

/, 

/ 

/ 

/ 

X 

/ 

——  Experimental 
Raoult's  and 
Dalton's  law 
X        Fugacity 

// 

/ 

// 

y 

y/ 

3              0.2              0.4 

06             0.8             1. 

Mol  fraction  propane  in  liquid, 
Fia.  3-6.     Vapor-liquid  curves  for  system  propane-isobutylene. 

to  the  close  packing  in  the  dense  phase.  The  theoretical  method  of 
attack  for  the  liquid  phase  is  not  so  simple  as  for  the  vapor  phase. 
For  the  vapor-phase  calculations  a  convenient  basis  was  possible 
because  at  low  pressure  all  vapor  mixtures  obey  the  perfect-gas  laws. 
Thus  the  deviations  could  be  calculated  on  the  basis  of  the  differences 
between  the  mixture  at  low  pressure  and  at  high  pressure.  In  the  case 
of  the  liquid  phase  no  such  convenient  basis  is  possible.  Thus  a  mix- 
ture of  ethyl  alcohol  and  water  does  not  agree  with  the  ideal  solution 
rules  under  any  practical  conditions  of  temperature  and  pressure. 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  47 

Basic  thermodynamic  relations  are  available  for  the  liquid  phase, 
but  their  practical  application  has  not  been  so  well  developed  as  those 
for  the  vapor  phase.  They  are  helpful  in  formulating  general  con- 
cepts and  are  directly  applicable  in  certain  special  cases.  One  of  the 
most  useful  relations  follows: 

+.-.=0(3-20) 


For  a  binary  mixture  dxi  —  ~rf#2,  Eq.  (3-20)  reduces  to 

(3-21) 


If  the  pressure  is  such  that  the  vapor  satisfies  the  perfect-gas  law,  then 
the  equation  can  be  modified  as  follows: 


This  equation  is  called  the  Duhem  equation  (Ref.  9).  Equation 
(3-20)  is  applicable  to  any  system  of  any  number  of  components,  but 
the  Duhem  equation  is  limited  to  a  binary  mixture  under  conditions 
such  that  the  perfect-gas  law  applies  to  the  vapor. 

Theoretically  these  equations  apply  only  to  a  process  carried  out  at 
constant  temperature  and  constant  total  pressure  on  the  liquid  phase.  In 
most  mixtures  encountered  in  distillation,  if  one  varies  the  composition 
at  constant  temperature,  the  total  pressure  also  varies  and  Eqs.  (3-20) 
to  (3-22)  are  not  strictly  applicable.  The  equations  would  apply  for 
this  constant-temperature  case  if  some  method  other  than  the  vapor 
pressure  were  employed  to  exert  pressure  on  the  liquid  which  was 
adjusted  to  keep  the  total  pressure  constant;  e.g.,  a  gas  insoluble  in  the 
liquid  could  be  added  to  the  vapor  to  maintain  constant  total  pressure. 
Actually,  these  equations  apply  satisfactorily  for  most  engineering 
purposes  if  they  are  employed  at  constant  temperature  and  a  variable 
total  pressure  equal  to  the  vapor  pressure.  The  error  introduced  is 
that  due  to  the  change  in  the  fugacity  of  the  liquid  with  the  total 
pressure  which  can  be  calculated  by  Eq.  (3-7).  A  more  exact  rela- 
tionship for  binary  mixtures  at  constant  temperature  is 

rainwpnl         Faincp^g)] 

L         0X1         Jr  L          ^2         Jr      t 

where  d  In  pf  is  the  fractional  change  in  effective  pressure  of  component 


48  FRACTIONAL  DISTILLATION 

1  due  to  the  change  in  total  pressure  on  the  liquid.  It  is  calculated  by 
Eq.  (3-7). 

d  ln  Pi  -  JLL  (3-24) 

rfT      "  RT  (6  M) 

where  TT  =  total  pressure  on  liquid  phase 

v\  =  partial  molal  volume  of  component  1 

R  =  gas  law  constant 

T  =  absolute  temperature 

The  value  of  d  In  p$  is  calculated  in  the  same  manner  with  v2  instead  of 
v\.  In  most  cases  these  corrections  at  constant  temperature  are  small 
in  comparison  to  the  change  in  fugacity  due  to  the  change  in  composi- 
tion; for  this  reason  Eq.  (3-22)  is  frequently  accepted  as  applying  to 
mixtures  under  their  own  vapor  pressure.  It  can  be  in  serious  error  at 
high  pressure,  and  Eq.  (3-23)  would  be  more  exact. 

For  constant  pressure,  variable  temperature  conditions,  Eq.  (3-22) 
generally  is  unsatisfactory  owing  to  the  rapid  change  in  fugacity  or 
vapor  pressure  of  a  liquid  with  changes  in  temperature.  Equation 
(3-22)  can  be  modified  to  compensate  for  this  effect  as  follows: 


f 

L 


where  d  In  P(  is  the  fractional  change  in  the  partial  pressure  of  com- 
ponent 1  at  a  composition  xi,  with  the  change  in  temperature.  It  is 
evaluated  by 

'  '  (3.26) 

^        } 


dT          RT* 

where  A£T'  is  the  heat  of  vaporizing  1  mol  of  the  component  from  the 
solution  into  a  vacuum. 

Equations  (3-23)  and  (3-25)  can  be  combined  and  expressed  in  the 
more  exact  fugacity  form, 


where  d  In  /°  is  the  fractional  change  of  the  fugacity  of  the  component 
in  the  liquid  phase  at  the  given  composition  and  is  calculated  by  Eq. 
(3-26)  for  constant  pressure  changes  and  by  Eq.  (3-24)  for  constant 
temperature  changes. 

The  Duhem  equation  as  such  cannot  be  integrated,  but  if  a  relation- 
ship between  the  pressure  of  one  of  the  components  and  the  mol  frac- 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  49 

tion  is  available,  it  is  possible  to  calculate  the  relationship  between  the 
partial  pressure  of  the  other  component  and  the  mol  fraction.  This  is 
not  of  any  real  engineering  utility  for  predicting  vapor-liquid  equilibria 
since  in  general  when  the  partial  pressure  of  one  component  is  known 
that  of  the  other  component  is  also  known.  However,  the  Duhem 
equation  is  useful  in  checking  experimental  data  and  also  in  guiding 
the  development  of  correlations.  For  example,  Eqs.  (3-22)  to  (3-27) 
can  be  used  to  evaluate  the  accuracy  of  vapor-liquid  equilibrium  data. 
Consider  the  case  of  a  binary  mixture  at  constant  temperature  for 
which  the  liquid- vapor  data  are  available.  By  Eq.  (3-23), 


j  1       Pi   ^2          i  *       PZ 

PI  ~~     i  -  x2    n  pf 

p^   i  M      r  **  ^  P« 

^*  )    ~  ln  I  -^  i    =  —    I     ^ a  In  —;: 

J>*/6  VPl/a  Ja    l-a;2  p? 


(3-28) 


If  one  experimental  value  of  Pi/p*  is  taken  as  a  base,  then  the  value 
of  this  ratio  can  be  determined  at  other  compositions  by  the  integration 
of  the  right-hand  side  of  the  equation.  This  integration  must  usu- 
ally be  performed  graphically,  and  a  convenient  method  is  plotting 
#2/(l  —  #2)  vs.  In  (pz/p*)  and  determining  the  area  under  the  curve. 
The  experimental  data  for  p2  as  a  function  of  x2  and  one  value  of  p\ 
allow  the  value  of  p\  to  be  predicted  at  other  compositions.  If  correc- 
tions for  total  pressure  on  p*  and  p*  are  to  be  included,  they  can  be 
evaluated  from  the  experimental  total  pressure  data  or  by  summing 
the  calculated  value  of  pi  and  p2;  in  the  latter  case,  the  integration 
becomes  trial  and  error. 

Constant  pressure  conditions  are  more  important  in  distillation  cal- 
culations than  constant  temperature,  and  for  this  condition  Eq.  (3-25) 
can  be  modified  to 


The  evaluation  of  the  Pf  terms  involves  the  heat  of  vaporizing  1  mol 
of  the  component  from  the  mixture.  At  moderate  pressure  this  is 
equal  to  the  latent  heat  of  vaporization  at  the  same  temperature  plus 
the  heat  effects  of  mixing  the  liquids  at  this  temperature  and  bringing 
them  to  the  total  pressure.  As  an  approximation  the  latent  heat  of 
vaporization  of  the  pure  liquids  can  be  employed  in  Eq.  (3-26),  but  it 
will  not  give  satisfactory  results  if  (1)  the  heat  of  mixing  is  large  or  (2) 


50  FRACTIONAL  DISTILLATION 

the  total  pressure  is  so  high  that  the  pressure-enthalpy  corrections  for 
the  vapor  are  large. 

One  of  the  cases  in  which  the  Duhem  equation  is  of  real  utility  in 
determining  vapor-liquid  equilibria  is  where  the  analysis  of  the  com- 
position of  the  two  phases  offers  serious  difficulties.  In  such  a  case,  if 
it  is  possible  to  prepare  known  mixtures  of  the  two  components  and  to 
determine  their  equilibrium  total  pressure  at  a  given  temperature, 
these  data  can  be  used  with  the  Duhem  equation  to  calculate  the  com- 
position of  the  equilibrium  vapor. 

Activity  Coefficient.  Because  there  are  no  convenient  conditions  for 
the  liquid  mixture  upon  which  to  base  calculations,  it  is  customary  to 
use  the  pure  liquids  before  they  are  mixed  as  the  basis,  and  then  calcu- 
late the  deviations  that  result  from  the  mixing  operation.  The  devia- 
tions in  the  liquid  phase  are  summed  up  in  what  is  termed  an  activity 
coefficient.  •  Thus  Raoult's  law  and  the  idealized  fugacity  law  are 
modified  by  the  insertion  of  a  factor  on  the  right-hand  side. 


pi  =  yi*  =  yfixi  (3-30) 

/i  =  yi/n  =  yifp&i  (3-31) 


The  value  of  the  activity  coefficient,  7,  is  the  factor  that  will  make 
these  equations  correct  for  the  case  in  question.  All  deviations  other 
than  those  associated  with  the  gas  law  are  lumped  into  the  one  value, 
and  the  problem  is  thus  made  into  one  of  predicting  the  activity  coeffi- 
cient. These  factors  will  be  different  for  each  component,  but  they 
are  interrelated  by  Eq.  (3-20)  and,  for  a  binary  mixture,  Eqs.  (3-21), 
(3-22),  (3-30),  and  (3-31)  become 


(3-32) 
V>T  v 


The  methods  that  have  been  used  for  predicting  the  activity  coeffi- 
cient are  either  empirical  or  semitheoretical,  the  theoretical  part  being 
the  use  of  thermodynamic  equations  to  direct  the  development  of 
empirical  rules.  A  number  of  rules  have  been  proposed  (Refs.  3,  5,  13, 
15,  28),  but  the  two  most  commonly  used  methods  of  estimating 
the  activity  coefficients  for  solutions  of  the  type  employed  in  distilla- 
tion are  the  Margules  and  Van  Laar  equations. 

Example  Illustrating  the  Use  of  Duhem  Equation.  Data  on  the  vapor-liquid 
equilibria  of  benzene-rc-propanol  and  ethanol-  water  are  given  in  the  accompanying 
tables.  Using  the  Duhem  equation,  check  the  consistency  of  the  data. 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


51 


EQUILIBRIUM  DATA  FOR  THE  SYSTEM  BENZENE-n-PROPANOL 
[From  Lee  (Ref.  14).     Temp.  -  40°C.] 


Mol  fraction  in  liquid 

Partial  pressure,  mm.  Hg 

Benzene 

n-Propanol 

Benzene 

w-Propanol 

0 

"      1.0 

0 

50.2 

0.099 

0.901 

59.6 

42.4 

0.209 

0.791 

95.0 

39.6 

0.291 

0  709 

118.6 

37.4 

0.360 

0.640 

132.2 

36,2 

0.416 

0.584 

139.1 

35.9 

0.508 

0.492 

149.2 

34.3 

0.700 

0.300 

161.6 

31.4 

0.820 

0.180 

167  4 

28.6 

0  961 

0.039 

175  6 

15.7 

1  0 

0 

183  5 

0 

VAPOR  PRESSURE  DATA  FOR  SYSTEM  KTHANOL- WATER 
(Ref.  12) 


Vapor  pro  HSU  re,  mm. 


J.  ,  1^. 

Ethanol 

Water 

76 

693 

301.4 

78 

750 

327.3 

80 

812 

355.1 

82 

877 

384  9 

84 

950 

416.8 

86 

1,026 

450.9 

88 

1,102 

487.1 

90 

1,187 

525.76 

92 

1,280 

566.99 

94 

1,373 

610.9 

96 

1,473 

657.6 

98 

1,581 

707.3 

100 

1,693 

760.0 

52 


FRACTIONAL  DISTILLATION 


ETHANOL-WATER  EQUILIBRIUM  DATA  AT  NORMAL  BAROMETRIC  PRESSURE 

(Ref.  4) 


T,  °C. 

Mol  fraction  of  ethanol 

Liquid  x 

Vapor  y 

95.7 

0.0190 

0.1700 

90.0 

0.0600 

0.3560 

86.4 

0.1000 

0  4400 

84.3 

0.1600 

0.5040 

83.3 

0.2000 

0  5285 

82.3 

0.2600 

0.5570 

81.8 

0.3000 

0  5725 

81.2 

0.3600 

0.5965 

80.7 

0.4000 

0  6125 

80.2 

0.4600 

0.6365 

79.8 

0.5000 

0.6520 

79.4 

0.5600 

0.6775 

79.13 

0.6000 

0.6965 

— 

0.6600 

0.7290 

78.6 

0.7000 

0.7525 

— 

0.7600 

0.7905 

78.3 

0.8000 

0  8175 

— 

0.8600 

0.8640 

78.17 

0  8943 

0.8943 

Solution  for  Benzene-n-Propanol  System  at  Constant  Temperature.     Using  Eq. 
(3-28)  given  on  page  49, 


Pi 


The  right-hand  side  of  this  equation  requires  graphical  integration.  At  the  low 
pressures  involved  the  variation  in  the  values  of  p*  and  p*  will  be  neglected.  In 
order  to  reduce  the  variation  in  the  groups  involved,  the  equation  was  rearranged  to 


In 


Pi* 


The  values  utilized  for  preparing  the  graphical  integration  plot  are  given  in  the 
first  five  columns  of  Table  3-5.  As  a  basis  for  calculations,  propanol  was  taken  as 
component  1  and  benzene  as  component  2.  As  the  check  point,  the  value  for  the 
partial  pressure  of  propanol  of  28.6  mm.  Hg  at  a  mol  fraction  of  propanol  of  0.180 
was  chosen. 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 
TABLE  3-5. 


53 


X2 

1  -*, 

Pz 

io-V2 

104s2 

In   Pl 

Plcalo 

plexp 

2/2calo 

2/2  oxp 

100  Ay 

(1  ~  xz)pl 

28.6 

y  -x 

0 

1.0 

0 

0 





. 

50.2 

t 

0 

_ 

0.099 

0.901 

59.6 

0  356 

0,309 

0.554 

49.8 

42.4 

0.545 

0  585 

9.2 

0.209 

0.791 

95.0 

0  903^ 

0  293 

0.472 

45.8 

39.6 

0.675 

0.706 

6.2 

0.291 

0.709 

118.6 

1.406 

0.292 

0.398 

42.6 

37.4 

0.736 

0.760 

5.1 

0.360 

0.640 

132.2 

1.747 

0.322 

0.346 

40.4 

36.2 

0.767 

0.783 

3.8 

0.461 

0  584 

139.1 

1.933 

0.408 

0.312 

39.1 

35.9 

0.781 

0.795 

4  2 

0.508 

0.492 

149.2 

2.225 

0  464 

0.249 

36.7 

34.3 

0.803 

0.814 

3.6 

0  700 

0.300 

161.6 

2.612 

0.894 

0.118 

32.2 

31.4 

0.833 

0.838 

3.6 

0  820 

0.180 

167  4 

2  800 

1.625 

0.0 

28  6 

28.6 

0.855 

0  855 

0.0 

0  961 

0.039 

175.6 

3  080 

8.0 

-0.673 

14.6 

15.7 

0  923 

0.917 

11.4 

1.0 

0 

183.5 

3  367 

0 

— 

— 

0 

— 

1  0 

— 

The  sixth  column  of  the  table  gives  the  area  on  each  side  of  the  check  point,  and 
these  are  of  course  equal  to  the  left-hand  side  of  the  equation  with  the  base  value, 
pib,  equal  to  28.6.  The  seventh  column  lists  the  calculated  values  for  the  partial 
pressure  of  propanol,  and  the  eighth  column  gives  the  experimental  values.  The 
agreement  is  not  so  good  as  would  be  desired  but  is  a  fair  correlation.  When  the 
partial  pressure  values  are  converted  to  mol  fractions  in  the  vapor,  the  agreement 
appears  somewhat  better.  The  ninth  and  tenth  columns  of  the  table  give  calcu- 
lated and  experimental  values  for  mol  fractions  of  benzene  in  the  vapor  phase. 
The  last  column  of  the  table  gives  the  difference  in  these  vapor  compositions  as  a 
percentage  of  the  difference  between  the  experimental  vapor  and  liquid  composi- 
tions. Most  of  the  calculated  points  are  within  5  per  cent  of  the  experimental 
points  on  this  basis,  and  this  is  probably  reasonably  good  accuracy  for  vapor- 
liquid  equilibrium  measurement. 

Solution  for  Ethanol-W ater  at  Constant  Pressure.     For  this  case,  it  is  convenient 
to  use  the  mol  fraction  form  of  Eq.  (3-29)  instead  of  partial  pressures. 


1  -  ; 


"p; 

/b 
,  r^r/In 

For  convenience  in  plotting,  this  equation  was  rearranged  to 


1/t 

p' 

** 


The  integration  of  this  equation  requires  a  knowledge  of  the  variations  of  P'  with 
temperature.  Equation  (3-26)  indicates  that  the  variation  of  the  pressure  with 
the  temperature  could  be  calculated  if  the  heat  of  vaporization  of  1  mol  of  the 
component  from  the  solution  into  a  vacuum  were  known.  This  heat  of  vaporiza- 


54 


FRACTIONAL  DISTILLATION 


tion  differs  from  the  true  heat  of  vaporization  of  the  liquid  by  two  main  effects : 
(1)  enthalpy  effect  due  to  mixing  of  the  liquid  phase  and  (2)  enthalpy  effects  result- 
ing from  nonideal  behavior  of  the  vapor  phase.  In  the  present  case  the  pressure  is 
low  enough  that  the  vapor-phase  effects  should  be  very  small,  and  for  illustration 
purposes  the  enthalpy  of  the  mixing  in  the  liquid  phase  will  be  neglected.  On  the 
basis  of  such  an  assumption  d  In  P'  becomes  equal  to  d  In  P,  where  P  is  the  vapor 
pressure  of  the  pure  component  at  the  temperature  in  question.  With  this  modifi- 
cation and  the  data  given  in  the  table,  it  is  possible  to  carry  out  the  integration. 
The  results  of  such  calculations  are  given  in  Table  3-6.  The  first  five  columns  of 

TABLE  3-6.     ETHANOL- WATER 


X'l 

2/2 

Pa 

«*(£)' 

10-8^2 

In  Vl/Pl 

103  Vi 

Pi 

Vi 

2/2  oalc 

100  Ay 

[(1  ~  32)(WP*)P 

m  0.0003  18 

10  Pi 

V  -  x 

0  019 

0  170 

1450 

0  138 

14  0 

.428 

1.325 

647 

0  856 

0  144 

17.2 

0  060 

0  356 

1180 

0  926 

6  9 

387 

1  272 

527 

0  670 

0  330 

-  8.8 

0  100 

0  440 

1035 

1  808 

6.1 

.359 

1  237 

459 

0  568 

0  432 

2  4 

0  2 

0.528 

916 

3  325 

7  5 

308 

1.177 

404 

0  475 

0  525 

0  9 

0.3 

0.572 

865 

4  40 

9.7 

262 

1  122 

380 

0.426 

0.574 

0.7 

0.4 

0.612 

827 

5  48 

12.2 

1  203 

1  058 

363 

0.384 

0  616 

1.9 

0.5 

0  652 

802 

6  60 

15  2 

1.129 

0.984 

351 

0  345 

0  655 

2.0 

0.6 

0  696 

782 

7.90 

19  0 

1  019 

0.880 

343 

0.304 

0  694 

2  1 

0.7 

0  752 

769 

9  57 

24  4 

0  839 

0  736 

337 

0.248 

0  752 

0  0 

0.8 

0.817 

760 

11.54 

34.7 

0.555 

0  554 

333 

0  184 

0  816 

5  9 

0.8943 

0  8943 

757 

13.96 

60  6 

0 

0.318 

332 

0.1057 

0  8943 

0 

this  table  give  the  values  of  the  groups  used  for  preparing  the  graphical  integration 
plot.  Ethanol  was  taken  as  component  2,  and  the  check  point  for  the  system  was 
taken  at  the  azeotrope.  The  sixth  column  gives  the  measured  areas  from  the 
azeotrope  and  is  therefore  equal  to  the  left-hand  side  of  the  equation.  The  seventh 
column  gives  calculated  mol  fraction  in  the  vapor  divided  by  vapor  pressure  for 
the  water  component.  The  ninth  column  lists  the  calculated  mol  fractions  of 
water  vapor.  The  tenth  column  gives  the  calculated  values  of  alcohol  in  the 
vapor  and  is  obtained  by  subtracting  the  ninth  column  from  1.  This  column  is 
directly  comparable  to  the  second  column  and  indicates  that,  except  for  the  first 
two  points,  the  agreement  is  excellent.  These  first  two  points  are  at  low  composi- 
tion and  may  be  somewhat  in  error.  The  last  column  in  the  table  gives  the 
percentage  deviations. 

Margules  Equation.  Margules  (Ref.  17)  developed  an  expression 
for  the  activity  coefficient  of  the  components  in  a  binary  mixture  by 
taking  empirical  expressions  for  these  coefficients  as  follows: 


In  71  = 
In  72  - 


+  bx%  +  cx\ 
i  +  Vx\  +  c'x\ 


When  the  Margules  equations  for  the  activity  coefficients  are  sub- 
stituted in  Eq.  (3-32),  the  following  relations  are  obtained: 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  55 

Xl  — ~ as    — *Xl  — -= =     —  (O^l  +  2bXiXz  4 

#2 — — —  =  ~#2 — - — *  =  —  (a'#2  ~|-  2b'xiXz  +  Sc'xzxl) 

By  Eq.  (3-32)  the  right-hand  side  of  the  two  preceding  equations 
must  be  equal  for  all  compositions,  and  this  relation  will  be  satis- 
fied when  the  coefficients  of  corresponding  terms  are  equal ;  thus,  using 
xl  =  1  —  x<t  and  equating  the  components  of  the  corresponding  terms 
gives  the  following  solution : 

a  =  0 
c'  =  -c 
a'  =  0 

26  +  3c 

0   -     ~  2 
In  71 
In  72 

It  will  be  noted  that  the  equations  have  two  independent  constants. 
However,  only  one  experimental  point  is  needed  to  evaluate  both  of 
these,  since  one  determination  of  the  vapor-liquid  equilibrium  for  a 
binary  mixture  gives  both  71  and  72,  and  these  values  can  be  used  to 
evaluate  the  constant.  If  more  data  are  available,  it  is  convenient  to 
plot  the  logarithm  of  the  activity  coefficient  divided  by  the  mol  fraction 
squared  for  the  other  component  vs.  the  mol  fraction  of  one  of  the 
components.  If  the  Margules  equation  agrees  with  the  data,  both  of 
the  activity  coefficients  should  give  straight  lines  which  are  parallel 
when  plotted  in  this  manner,  and  the  slope  should  be  equal  to  the  value 
of  the  constant  c.  The  intercepts  will  be  different  but  can  be  used  for 
evaluating  the  constant  b.  The  two  activity  coefficient  equations  can 
be  arranged  to  plot  as  a  single  line. 

7i 

=*   U  -p  1^2 

(3-34) 


^__Lf  «  5  +  C(Q.5  +  x, 

Xi 

Thus  if  (In  71) /x\  is  plotted  vs.  #2,  and  (In  72)/£?  is  plotted  against 
0.5  +  xZj  they  will  both  give  a  line  of  slope  c  and  intercept  b.  Such  a 
plot  facilitates  the  evaluation  of  the  constants. 

The  constants  in  the  Margules  equation  are  a  function  of  tempera- 
ture; thus,  if  one  experimental  point  is  used  to  evaluate  them,  then 


56  FRACTIONAL  DISTILLATION 

assuming  the  equation  to  apply,  they  should  be  suitable  for  other  com- 
positions at  the  same  temperature.  A  generalized  relationship  for  the 
effect  of  temperature  on  the  activity  coefficient  is  not  available,  but  as 
an  approximation  it  is  suggested  that  the  constants  be  taken  as 
inversely  proportional  to  the  one-fourth  power  of  the  absolute  tem- 
perature, Thus, 

T°-25  In 


T2 
#2 


= 


In  72 


Van  Laar  Equation.  Van  Laar  (Ref .  25)  attempted  to  follow  a  more 
theoretical  approach  than  did  Margules.  He  based  his  derivation  on 
the  thermodynamic  changes  occurring  on  the  mixing  of  pure  liquids. 
Two  of  the  basic  thermodynamic  equations  relating  to  isothermal 

processes  are 

AF  =  A#  -  T  AS  (3-35) 

d&F  =  RTdlnf  (3-36) 

where  AF  =  partial  molal  change  in  free  energy 
&B  =  partial  molal  change  in  enthalpy 

T  =  absolute  temperature 
A$  =  partial  molal  change  in  entropy 

/  =  fugacity 

For  an  ideal  mixture,  there  is  no  change  in  volume  when  the  mixing 
is  carried  out  at  constant  temperature  and  under  a  constant  total 
pressure,  and  there  are  no  heat  effects.  Thus  A#  is  equal  to  zero,  and 
the  partial  molal  change  in  free  energy  is  equal  to  —  T  A/S  for  each  com- 
ponent. For  such  an  ideal  solution,  the  partial  molal  entropy  of  mix- 
ing is  —  R  In  x,  and  the  fugacity  is  proportional  to  the  mol  fraction 
(Lewis  and  Randall  rule).  However,  in  most  actual  solutions  there 
are  changes  of  volume  on  mixing,  heat  effects,  and  the  entropy  of  mix- 
ing differs  from  that  for  an  ideal  solution. 

The  combination  of  Eqs.  (3-30)  and  (3-31)  with  Eq.  (3-36)  gives 

RT  d  In  7  =  d  AFa  -  d  AF»  =  d  AFe  (3-37) 

and 

AFe  =  A#e  -  T  A&  (3-38) 

where  A^a  =  actual  partial  molal  free  energy 

AFi  =  partial  molal  free  energy  of  ideal  solution  of  same  com- 
position =  RT  In  x 

Ape  =  excess  partial  molal  free  energy  =  AFa  —  APi 
A5C  =  excess  partial  molal  enthalpy  =  AJ?« 
A&  «  excess  partial  molal  entropy  =  AS«  —  R  In  x 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  57 

Thus  in  calculating  the  activity  coefficient,  7,  the  value  of  the  excess 
partial  molal  free  energy  is  desired.  If  methods  were  available  for 
calculating  the  partial  molal  entropy  and  enthalpy  effects  upon  mixing, 
the  deviations  from  the  ideal  solution  law  could  oe  determined.  As  a 
first  approximation  for  systems  containing  molecules  not  too  dissimi- 
lar, several  writers  (Refs.  11,  22,  25)  have  recommended  the  assump- 
tion of  a  solution  in  which  only  the  internal  energy  change  on  mixing  is 
different  from  an  ideal  solution.  In  such  solutions,  which  Hildebrand 
terms  "regular,"  the  partial  molal  volume  of  each  component  remains 
constant,  and  the  change  in  entropy  is  equal  to  that  of  an  ideal  solu- 
tion; i.e.,  complete  " randomness"  exists.  Thus  the  only  deviation  of 
a  "  regular  "  solution  from  an  ideal  solution  is  due  to  the  fact  that  there 
is  an  excess  partial  molal  internal  energy  of  mixing.  In  addition  to 
these  assumptions,  Van  Laar  attempted  to  calculate  the  internal 
energy  of  mixing  by  the  use  of  the  van  der  Waals  equation  of  state. 
The  assumptions  used  by  Van  Laar  in  deriving  his  equations  are 

1.  A£  =  Q',i.e.,  ASa  =  A£. 

2.  No  volume  change  on  mixing. 

3.  The  van  der  Waals  equation  applies  to  each  of  the  components 
and  to  the  mixture,  both  as  liquids  and  as  vapors. 

4.  The  van  der  Waals  constants  of  the  mixture  can  be  calculated 
from  the  constants  of  the  pure  constituents. 

For  a  van  der  Waals  fluid,  i.e.,  one  that  satisfies  the  relationship 

„         ET          a    .,         ,      ,          .,. 
P  =       _  ,  —  •=£,  it  can  be  shown  that 


(*l\  _± 

\dV/T    ~    V* 


where  E  =  internal  energy 

V  =  volume 

a  =  van  der  Waals  constant 

T  =  absolute  temperature 

When  this  relationship  is  integrated  from  a  vapor  at  zero  pressure  to  the 
liquid  state, 

(3"39) 


Van  Laar  substituted  the  volume  constant,  6,  of  the  van  der  Waals 
equation,  for  the  molal  volume  of  the  liquid  and  used  the  values  devel- 
oped by  other  investigators  (Refs.  2,  10)  for  evaluating  the  constants 
of  a  mixture. 


Otnix  =  (xi  Voi  +  £2  Vosi  +  x*  Vos  +  '  '  '  )2  1         (3-40) 

femix    =  Xibi  +  Xj)2  +  Xzbz  +    '    '    •  J 


58  FRACTIONAL  DISTILLATION 

where  o»t*,  &«**  «  van  der  Waals  constants  of  mixture 
ai,«2,a«,  bifbzjb9  =  van  der  Waals  constants  of  pure  components 
For  a  binary  mixture  the  internal  energy  of  mixing  the  pure  liquid 
components  at  constant  temperature  per  mol  of  mixture  is 

A.EL  =  ELM  —  X\E^  --  X%ELZ 

where  AEL  =  internal  energy  change  in  mixing  per  mol  of  mixture 
ELM  =  molal  internal  energy  of  liquid  mixture 
ELI  =  molal  internal  energy  of  pure  component  1  as  a  liquid 

before  mixing 
EL»  =  molal  internal  energy  of  pure  component  2  as  a  liquid 

before  mixing 
By  Eq.  (3-39)  and  using  6  =  F, 


T H  (  Emi*  oo    —   X\Ei^    —   #2/?2oe    I 


6mix  6l 

The  last  term  is  the  internal  energy  of  mixing  of  the  vapors  at  zero 
pressure  and  is,  therefore,  equal  to  zero.     The  values  of  Eq.  (3-40)  give 


From  Eq.  (3-41)  the  partial  molal  change  in  internal  energy  is 

,»  ^ 

(3-42) 


_  .      l6l  + 

and  on  the  basis  of  Van  Laar's  assumptions  : 


lnTl  «  ^l'L       "*   ^  bl 


combining  constants, 

In 

where  A  *  61/62 

61 

62 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  59 

By  combining  Eq.  (3-44)  with  Eq.  (3-32), 

'          ln^=-      AB/T 


Scatchard  and  Hildebrand  obtained  similar  expressions  without  the 
use  of  a  van  der  Waals  fluid.  They  recommend  evaluating  the  expres- 
sion for  (A£I)L  as 


where  w  =  internal  energy  of  vaporization 

V  =  molal  volume 

This  leads  to  equations  of  the  same  form  as  Eqs.  (3-44)  and  (3-45), 
except  that 


It  has  been  pointed  out  by  Cooper  (Ref  .  7)  that  the  same  relation- 
ship can  be  obtained  more  simply  than  the  method  employed  by  Van 
Laar.  What  is  desired  in  calculating  the  activity  coefficient  is  the 
difference  between  the  partial  molal  free  energy  of  mixing  of  an  actual 
solution  and  that  of  an  ideal  solution.  If  the  excess  free  energy  of 
mixing  per  unit  volume  of  mixture  is  assumed  proportional  to  the 
product  of  the  volume  fractions  of  the  two  components,  an  expression 
identical  in  form  to  the  Van  Laar  equation  is  obtained.  Thus, 

(3.47) 


where  AFe  =  excess  free-energy  change  of  mixing  =  actual  free-energy 
change  minus  ideal  free-energy  change 

m  =  mols  of  component  1 

n2  =  mols  of  component  2 

Vi  =  molal  volume  of  component  1 

F2  =  molal  volume  of  component  2 
But 

(MB 


60  FRACTIONAL  DISTILLATION 

and 


In  -n  =  S  7-^ x-2       •  (3-49) 


In  71  =  y  v2  (3-50) 


where  J5  =  K'Vi/R 
A  «  7i/Fi 

These  equations  are  identical  with  those  given  by  Van  Laar,  but  the 
assumptions  are  somewhat  different.  In  the  case  of  the  Van  Laar, 
Scatchard,  and  Hildebrand  derivations,  both  A  and  B  should  be  posi- 
tive, while  in  Cooper's  equation  B  could  be  either  positive  or  negative. 

On  the  basis  of  the  derivations,  the  two  constants  of  the  Van  Laar 
equations  and  the  modifications  of  it  are  related  to  the  physical  proper- 
ties of  the  pure  components.  When  the  best  values  of  the  constants 
are  chosen  to  fit  the  data,  they  usually  do  not  agree  with  the  predicted 
values,  although  the  trends  are  approximately  the  same.  Generally 
the  constants  are  chosen  to  agree  with  the  data,  and  the  equations  are 
used  empirically. 

As  was  the  case  with  the  Margules  equation,  two  constants  are 
involved.  For  a  binary  mixture  one  vapor-liquid  equilibrium  point 
will  give  the  activity  coefficients  of  both  components  and  thus  define  the 
whole  equation.  The  form  of  the  equations  are  significantly  different 
from  that  of  the  Margules.  It  includes  a  temperature  correction,  and 
the  value  of  the  constants  should  be  independent  of  temperature. 
Thus  an  experimental  determination  at  one  temperature  should  allow 
equilibrium  data  to  be  calculated  at  other  temperatures.  There  are 
several  ways  in  which  Van  Laar  equations  can  be  rearranged  to  plot 
as  a  straight  line  in  order  that  the  data  for  more  than  one  determination 
can  be  easily  correlated.  One  of  the  most  convenient  methods  of 
making  such  a  plot  is  to  use  the  reciprocal  of  the  square  root  of  the 
temperature  times  the  logarithm  of  the  activity  of  component  1, 
l/(r  In  7i)H,  vs.  the  ratio,  x\/x*.  A  similar  plot  can  be  made  for  the 
other  component.  The  plots  should  both  be  straight  lines;  the  slopes 
and  intercepts  will  be  different  but  related  because  they  are  based  on 
the  same  constants. 

Clark  Equation.  For  the  vapor-liquid  equilibria  of  a  binary  mixture 
at  either  constant  pressure  or  constant  temperature,  Clark  (Ref .  6)  has 
suggested  that  the  ratio  of  the  mol  fractions  in  one  phase  is  a  linear 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  61 

function  of  the  ratio  of  the  mol  fractions  in  the  other  phase,  when  the 
ratios  are  utilized  such  that  the  component  in  largest  amount  appears 
in  the  numerator. 

Thus,  when  component  1  is  present  in  largest  amount, 

8i-0£!  +  6  (3-51) 

2/2          o:2  v        ' 

and,  when  component  2  is  present  in  largest  amount, 

»!«a'£!  +  6'  (3-52) 

y\        xi 

Clark  uses  the  value  xi/xz  =  ^a'b/aV  as  the  change-over  point 
between  the  two  equations.  The  use  of  these  equations  requires  three 
experimental  points.  This  greatly  limits  its  utility. 

Evaluation  of  Empirical  Equations.  Tucker  (Ref.  24)  and  Mason 
(Ref.  18)  have  studied  the  agreement  of  the  various  equations  with 
published  experimental  data.  The  data  available  were  screened,  and 
only  those  that  gave  good  agreement  with  the  Duhem-type  equations 
were  selected  for  the  evaluation.  The  experimental  data  were  plotted 
to  evaluate  the  equation  constants,  and  the  average  constants  so 
obtained  were  used  to  recalculate  the  y,x  curve.  In  distillation  calcu- 
lations the  difference  between  the  vapor  and  the  liquid  compositions 
gives  a  better  indication  of  the  ease  of  separation  than  the  absolute 
value  of  the  vapor  composition.  Mason  therefore  made  the  compari- 
son on  the  basis  of 

Per  cent  deviation  =  \y^°  ~  ^exp|  100 
L  (V  -  z)e*pj 

As  a  qualitative  standard,  he  classified  average  deviation  of  0  to  5 
per  cent  as  good,  5  to  11  per  cent  as  fair,  and  greater  than  11  per  cent 
as  poor.  Some  of  the  results  are  given  in  Table  3-7.  The  table  gives 
the  average  deviation  and  the  maximum  deviation.  In  certain  cases, 
the  agreement  between  the  experimental  and  the  calculated  values  is 
very  good  but  very  poor  in  other  cases.  It  is  difficult  to  determine  any 
definite  types  of  mixtures  that  give  good  or  poor  agreement.  However, 
in  all  cases,  mixtures  approaching  immiscibility,  i.e.,  y>x  curves  that 
are  nearly  horizontal  over  an  appreciable  concentration  region,  gave 
poor  agreement  with  the  Margules  and  Van  Laar  equations.  Good 
agreement  was  obtained  with  all  the  maximum  boiling  mixtures 
studied.  These  latter  mixtures  give  negative  values  of  the  Van  Laar 


62 


FRACTIONAL  DISTILLATION 
TABUS  3-7 


Margulea 

Scatchard 

Van  Laar 

Clark 

System 

Max. 

Av. 

Corre- 

Max. 

Av. 

Corre- 

Max. 

Av. 

Corre- 

Max. 

Av. 

Corre- 

% 
dev. 

% 
dev. 

lation 

% 
dev. 

% 
dev. 

lation 

% 
dev. 

dev. 

lation 

dev. 

% 
dev. 

lation 

Ethanol-water 

14  6 

4.5 

Good 

5  8 

3  1 

Good 

5  7 

4  4 

Good 

10  6 

2.4 

Good 

Methanol  -water  . 

6  9 

4  9 

Good 

6  2 

5  1 

Good 

14  7 

5  8 

Good 

2  3 

1.0 

Good 

Kthylenebromide-  1- 

23  8 

6  8 

Good 

Vi  **  V*, 

Good 

8  2 

2  9 

Good 

nitropropano 

see  Mar- 

gules 

Carbon  disulfide  -ben- 

zene.                

13.9 

5  3 

Good 

6  7 

3  0 

Good 

11   1 

4  3 

Good 

Benzene—  aniline  

1.5 

0.5 

Good 

Vi  - 

V,. 

Good 

2.0 

0.5 

Good 

see  Mar- 

gules 

Carbondisulfido-  nitro- 

benzene*   

2  5 

1.8 

Good 

1  3 

0  7 

Good 

7  5 

2  7 

Good 

Chloroform—  nitroben- 

zene* ... 

1  4 

0  5 

Good 

0  8 

0  2 

Good 

6  8 

2  2 

Good 

Ethyl  ether-nitroben- 

0.7 

0.3 

Good 

Vi  ->i 

Good 

11  8 

4  2 

Good 

zene* 

see  Mar- 

gules 

Methanol-nitroben- 

fsene*   .      .          . 

4.1 

1.7 

Good 

19  9 

5  9 

Good 

16  2 

5  5 

Good 

Carbon  tetrachloride~ 

benzene  

6.6 

3.6 

Good 

6  6 

3  6 

Good 

15  0 

6  2 

Fair 

18  2 

8  5 

Fair 

Acetone-chloroform  .  . 

12  5 

4  3 

Good 

17.6 

4  8 

Good 

52  9 

13  5 

Fair 

55.2 

13  5 

Fair 

Ethyl  ether-acetone.  . 

21.7 

6.0 

Good 

12  1 

4  7 

Good 

21  0 

8.5 

Fair 

Benzene~phenol  .  .  ,  ,  . 

0.8 

0  5 

Good 

Vi  - 

V» 

Good 

50  0 

16  5 

Poor 

see  Mar- 

gules 

Benzene-nitroben- 

1.8 

0.7 

Good 

1.8 

0.8 

Good 

No  corre- 

Poor 

zene* 

lation 

CS*—  isobutylene-  chlo- 

ride           

14  7 

7  0 

Fair 

16  4 

6  8 

Fair 

14  2 

6  7 

Fair 

Isopentane~CS$ 

23.3 

9  0 

Fair 

26  8 

9.1 

Fair 

10  9 

6  5 

Fair 

CCli-ethyl  acetate.  .  . 

40.0 

11.8 

Fair 

40  0 

11.1 

Fair 

54  5 

32  8 

Poor 

40  0 

10.1 

Fair 

n-  Heptane-toluene  .  .  . 

53  5 

9.6 

Fair 

53  4 

9  8 

Fair 

15  9 

8  9 

Fair 

8  7 

4  2 

Good 

Benzene-n-propanol  .  . 

53.3 

8.6 

Fair 

68  4 

11  4 

Poor 

18.8 

17  8 

Poor 

16  3 

4  7 

Good 

Methanol-benzene  .  .  . 

23  0 

8  5 

Fair 

65  2 

18  4 

Poor 

76  8 

28  8 

Poor 

Water—  pyridine   .  . 

54.4 

21  .  1 

Poor 

32  0 

12   1 

Poor 

47  7 

18.4 

Poor 

3  9 

1.5 

Good 

GSj-chlorof  orm  .    .    .  . 

23  2 

11.2 

Poor 

14  7 

13.9 

Poor 

23  5 

16  3 

Poor 

12  3 

4.5 

Good 

Ethyl  ether-CSa  

65.0 

32.1 

Poor 

50.0 

21.3 

Poor 

40  0 

15.5 

Poor 

CSt—  acetone  

29.1 

16.0 

Poor 

33  7 

18  8 

Poor 

94.4 

59.2 

Poor 

CSr-cyolohexane.    .  .  . 

51  0 

24.4 

Poor 

26.6 

14.6 

Poor 

73  3 

74.9 

Poor 

Cyoiohexane-ethanol  . 

48.2 

11.8 

Poor 

50.0 

12.8 

Poor 

29  1 

11  0 

Fair 

18.8 

3.8 

Good 

The  per  cent  deviations  are  based  on  [(yealo  —  y«p)/(v  — 
These  systems  marked  *  are  baaed  on  [(po*io  —  p«xp)/Pe*pl  100. 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  63 

constant,  B,  which  is  not  consistent  with  the  derivation,  but  the  use  of 
the  negative  value  gave  satisfactory  predictions.  It  was  also  found 
that  the  use  of  one  point  for  the  estimation  of  the  constants  of  the 
Van  Laar  equation  was  not  very  satisfactory.  In  some  cases  the 
single-point  method  worked  well  and  in  other  cases  very  poorly,  and 
it  depends  upon  the  accuracy  of  the  particular  point  in  question. 
However,  if  only  one  point  is  available,  such  as  is  frequently  the  case 
when  the  azeotrope  only  is  known,  the  Van  Laar  and  Margules  equa- 
tions probably  offer  the  best  method  of  estimating  the  vapor-liquid 
equilibria  for  other  conditions.  When  data  on  the  azeotropic  condi- 
tion are  not  available,  they  can  be  estimated  from  Figs.  3  to  5  of  the 
Appendix. 

The  Van  Laar  equation  is  also  very  useful  for  transferring  data 
obtained  at  constant  temperature  to  constant  pressure  and  the 
reverse.  It  is  also  useful  for  transferring  data  from  one  temperature 
to  another.  The  unsatisfactory  agreement  in  the  cases  of  solutions 
approaching  immiscibility  is  not  surprising.  In  these  cases,  the 
entropy  of  mixing  cannot  be  equal  to  that  for  an  ideal  solution,  and 
for  complete  immiscibility  the  entropy  of  mixing  would  be  zero.  It 
is  therefore  not  surprising  that  solutions  approaching  immiscibility 
deviate  from  the  Van  Laar  equation. 

Other  Applications  of  the  Van  Laar  Equation.  The  Van  Laar  equa- 
tion can  be  used  to  indicate  qualitatively  the  type  of  phenomena 
encountered  in  liquid  mixtures.  Thus,  it  can  be  arranged  as  follows: 


This  equation  gives  the  logarithm  of  the  ratio  of  the  vapor  pressures 
divided  by  the  relative  volatility  as  a  function  of  the  Van  Laar  con- 
stants and  the  concentrations.  If  the  solution  were  ideal,  the  loga- 
rithm term  would  be  zero.  Thus  the  real  fact  determining  deviation 
from  ideal  solution  is  the  constant  B.  If  B  is  equal  to  zero,  the  relative 
volatility  will  be  equal  to  the  ratio  of  vapor  pressures,  and  the  y,x 
values  will  be  the  same  as  those  calculated  by  Raoult's  and  Dalton's 
laws.  If  B  is  not  equal  to  zero,  the  system  is  not  ideal.  The  terms 
involving  A  and  the  concentrations  would  have  about  the  same  varia- 
tion independent  of  the  value  of  B. 

It  is  interesting  to  consider  this  equation  for  various  limits.    For 
example,  for  Xi  equal  to  zero,  the  logarithm  term  equals  —  B/T.    For 


64  FRACTIONAL  DISTILLATION 

most  mixtures  encountered,  B  is  positive.  (B  is  negative  for  solutions 
having  maximum  boiling  azeotropes  and  for  certain  other  solutions.) 
Thus  the  logarithm  is  negative,  and  the  relative  volatility  is  greater 
than  the  ratio  of  the  vapor  pressures.  In  these  cases,  it  is  easier  to 
remove  the  component  in  low  concentration  than  would  be  expected 
from  the  ideal  solution  law.  At  the  other  extreme,  i.e.,  #2  equals  zero, 
then  Xi  over  x2  is  infinity,  and  the  logarithm  becomes  B/AT,  and  with 
B  positive  (in  all  cases  so  far  encountered  A  is  positive)  the  relative 
volatility  is  less  than  the  ideal  relative  volatility.  These  conditions 
are  found  in  most  common  mixtures;  i.e.,  the  relative  volatility  at  low 
concentration  is  greater  than  that  of  an  ideal  solution  and  the  relative 
volatility  at  high  concentration  is  lower.  It  is  often  expressed  by 
saying  that  the  components  in  small  amount  are  squeezed  out.  Thus, 
at  the  low  concentration  of  the  volatile  component,  it  is  squeezed  out 
and  the  relative  volatility  is  high.  At  high  concentration  of  the 
volatile  component,  the  nonvolatile  component  is  squeezed  out,  and 
the  relative  volatility  is  low.  For  the  mixture  in  which  B  is  negative, 
the  reverse  phenomena  are  true. 

The  Van  Laar  equation  also  would  state  that  a  mixture  would  agree 
with  Raoult's  law,  independent  of  the  value  of  By  when  the  ratio  of  the 
mol  fractions  equals  l/-\/~A.  For  most  mixtures,  the  value  of  A  is 
somewhere  between  0.5  and  2,  which  would  indicate  that  the  vapor- 
liquid  equilibria  at  some  concentration  in  the  middle  range  would 
agree  with  Raoult's  law.  Thus,  the  Van  Laar  equation  would  imply 
that  the  assumption  of  Raoult's  law  for  the  region  around  a  mol 
fraction  of  0.5  would  be  more  satisfactory  than  at  the  two  ends  of  the 
curve.  It  should  be  emphasized  that  this  relation  does  not  state  that 
Raoult's  law  is  valid  at  this  condition.  It  indicates  that  the  relative 
volatility  is  equal  to  the  ideal  relative  volatility  at  this  concentration, 
but  the  temperature-total  pressure  relationships  may  be  far  from  those 
indicated  by  Raoult's  law.  Thus,  in  the  system  ethyl  alcohol  and 
water,  the  total  pressure  is  always  higher  than  would  be  indicated  by 
Raoult's  law,  but  the  vapor-liquid  equilibrium  curve  crosses  the 
Raoult's  law  curve.  Below  this  intersection,  ethyl  alcohol  is  more 
volatile  than  would  be  indicated  by  Raoult's  law;  above  this  concen- 
tration, it  is  less  volatile"than  would  be  indicated  by  Raoult's  law. 

The  Van  Laar  equation  would  indicate  that  the  mixtures  would 
become  more  ideal  as  the  temperature  increased.  In  other  words,  the 
ratio  of  B/T  becomes  smaller  and  nearer  to  the  value  for  an  ideal 
solution. 

The  Van  Laar  equation  gives  interesting  relationships  for  the  condi- 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  65 

tion  under  which  the  relative  volatility  becomes  unity,  i.e.,  the  forma- 
tion of  an  azeotrope.  For  this  condition,  the  equations  can  be 
arranged  as  follows: 


7T 

t  =  5 


and  for  azeotropic  conditions: 


A  -  !n  <'  (3-54) 

In  (T/P 


COMBINED  GAS  LAW  AND  SOLUTION  DEVIATION 

In  the  preceding  discussion  the  deviations  of  the  vapor-liquid 
equilibria  due  to  gas  law  and  solution  abnormalities  have  been  treated 
separately;  however,  such  deviations  frequently  occur  simultaneously. 
In  such  cases,  the  factors  are  combined  into  a  single  equation  as 
follows  : 

yf,  =  7/pS  (3-56) 


In  many  cases,  by  combining  the  Van  Laar  and  Margules  equations 
with  the  fugacity  corrections,  it  is  possible  to  correlate  vapor-liquid 
equilibrium  data  up  to  high  pressure. 

At  very  high  pressures,  it  is  probable  that  the  density  of  the  vapor 
phase  may  be  such  that  deviations  of  the  type  handled  by  the  activity 
coefficient  equations  also  occur  in  this  phase.  Thus, 

yvvf*  =  7L/p3  (3-57) 

The  deviations  encountered  in  yv  would  probably  be  of  a  lower  degree 
than  those  for  the  liquid  phase,  but  equations  of  the  type  of  the  Van 
Laar  should  be  applicable.  Two  experimental  points  should  be 
sufficient  to  calculate  the  constants  involved  for  the  activity  coeffi- 
cients of  the  liquid  and  vapor  phase. 

Van  Laar-Margules  Equations;  Example.  The  table  of  data  below  gives  the 
vapor-liquid  equilibria  for  the  system  acetone-chloroform  obtained  by  Rosanoff 
and  Easley  (21),  for  a  total  pressure  of  760  mm.  Hg. 

1.  Using  the  Van  Laar  equation,  calculate  the  y,x  curve, 

a.  Employing  the  azeotrope  data  only  to  evaluate  the  constants. 
6.  Using  all  of  Rosanoff'  s  points  to  obtain  best  average  constants. 

2.  Repeat  la  and  16  using  the  Margules  equation. 


66  FRACTIONAL  DISTILLATION 

VAPOR-LIQUID  EQUILIBRIUM  VAPOR  PRESSURE  DATA 


Temp.,  °C. 

Xi 

2/i 

Pi,  mm.  Hg 

P2,  mm.  Hg 

57  45 

0.9145 

0.9522 

780 

675 

58.34 

0  8590 

0.9165 

805 

695 

59.44 

0.7955 

0.8688 

835 

745 

60.42 

0  7388 

0.8235 

865 

725 

61.60 

0.6633 

0.7505 

900 

775 

62.84 

0.5750 

0.6480 

935 

805 

63.91 

0.4771 

0.5170 

970 

835 

64.6 

0  3350 

0.3350 

995 

855 

64.36 

0.2660 

0.2370 

990 

850 

63.84 

0.2108 

0.1760 

965 

830 

63.08 

0.1375 

0.100 

950 

810 

62  77 

0  1108 

0.0650 

930 

795 

xi  "•  mol  fraction  of  acetone  in  liquid 

y\  «»  mol  fraction  of  acetone  in  vapor 

Pi  ***  vapor  pressure  of  acetone 

P»  »  vapor  pressure  of  chloroform 

Solution  of  Part  la.     Employing  azeotrope  data  only, 
Constant  boiling  temperature  «  64.6°C.  or  337.6°K 


At  azeo  tropic  point,  x 


yir  =  yPx          or 
y,  therefore,  y  ~  ir/P. 


• 

Px 


-  76%i>5  -  0.764;         In  Ti  -  -0.269 

-  76%55  -  0.889;         In  ^2  -  -0.118 

,  B/T 

In  yi  - 


AB 

In  y*  —  y 

xt  -  0.335          #2  -  0.665 
Solving  the  Van  Laar  equations,  one  obtains 

A  -  1.74         and         B  -  -320 

The  following  is  the  procedure  used  to  calculate  y  at  various  values  of  x: 
Let  xi  «  0.796  and  xa  —  0.204, 

5i  *  3.90 


Assume  T  -  332°K. 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 

In  7i    =    M    rn/o  r^\ r~T75    **     —0.0159 

In  71 


67 


[1.74(3.90)  +  lp 
••  0.998 
1 


In  72 


In  72 


In  72  -  1.74(3.90)2( -0.0159)  -  -0.421 
72  -  0.685 


2/2 

+2/2 


:  0-998(835)  (0.796)  _  Q  g?1 

0.685(725)  (0.204) 

,  _ .  0.133 

•  1.004 


* 

The  fact  that  the  sum  of  the  mol  fractions  is  greater  than  1.0  indicates  the  assumed 
temperature  is  too  high,  but  the  relative  volatility  is  nearly  constant  with  small 
temperature  changes,  and 

"•-S-"-"" 

0.133  _ 
2/2  "  1T004  -°-WJ 

Other  results  are  tabulated  as  follows: 


X 

2/calo 

2/exp 

100  A?/ 

(y  -  z)exp 

0.9145 

0.955 

0  9522 

7  9 

0  7955 

0  868 

0.8688 

1  1 

0  6633 

0.758 

0.7505 

8  6 

0.4771 

0.521 

0.5170 

10  0 

0.3350 

0  3350 

0  3350 

0  0 

0.2108 

0  175 

0.1760 

2  9 

0.1108 

0  070 

0.0650 

10  9 

Solution  of  Part  Ib 


In  71 


B/T 


(Tin- 


Btt 


xt 


Also, 


i  AB/T 

h,  72  .  -^^^^ 


4- 


(a) 


(6) 


68 


FRACTIONAL  DISTILLATION 


By  plotting 
evaluated. 


(T  In 


vs.  ~i  the  constants  A  and  B  can  be 


TABLE  3-8 


Xi 

Xz 

T,  °K. 

7i 

1 

5l 

Xz 

(-rin-yi)* 

0.9145 

0.0855 

330 

1.015 





0.8590 

0.1410 

331 

1  007 

— 

— 

0,7955 

0.2045 

332 

0  991 

0.709 

3.89 

0.7388 

0  2612 

333 

0.980 

0.376 

2.83 

0.6633 

0.3367 

335 

0.956 

0.258 

1  965 

0.5750 

0.4250 

336 

0.916 

0.185 

1  35 

0.4771 

0.5229 

337 

0.849 

0.135 

0.913 

0  3350 

0.6650 

338 

0.762 

0.105 

0  504 

0.2660 

0.7340 

337 

0  684 

0.088 

0  363 

0.2108 

0.7892 

337 

0.657 

0.084 

0  268 

0.1375 

0.8625 

336 

0.582 

0.074 

0.1596 

0.1108 

0.8892 

336 

0.479 

0  063 

0.1246 

In  this  case  the  mixture  is  of  the  maximum  boiling  point  type,  and  the  activity 
coefficients  are  less  than  unity,  resulting  in  negative  values  for  In  7.  This  negative 
value  is  handled  in  the  square  roots  by  multiplying  Eq.  (a)  by  l/\/  —  1.  The 
intercept  is  then  l/\/—jB,  and  the  slope  is  A/^/  —  B-  A  similar  procedure  is 
employed  for  In  72.  The  first  two  values  of  71,  given  in  Table  3-8  are  greater  than 


0.4 
03 
02 
0.1 
0 

^ 

O 

/ 

^- 

/ 

^^ 

^ 

^ 

< 

I/ 

& 

^ 

«^ 

/ 

^ 

^ 

0         '            v 

.  |i 

,/ 

^^x 

f-^ 

^^ 

"  •TuT?,  ' 

2 

^ 

>" 

& 

31234567 
*i  „  *2 

FIG.  3-7.     Van  Laar  plot  for  system  acetone-chloroform. 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


69 


unity,  while  the  values  of  72  at  the  same  point  are  less  than  unity,  and  it  <;an  be 
shown  that  this  is  not  consistent  with  the  Duhem  equation.  These  two  values  of 
71  are  so  near  to  unity  that  a  very  small  error  would  account  for  the  discrepancy, 
and  the  values  were  not  used  in  the  calculations.  In  all  cases  the  values  of  the 
activity  coefficients  near  unity  tend  to  be  of  little  value  for  calculating  the  constants 
of  either  the  Van  Laar  or  Margules  equations  because  In  7  is  small  and  subject  to 
large  errors.  Unfortunately,  the  method  of  plotting  used  in  Fig.  3-7  emphasizes 
these  inaccurate  points  and  gives  less  weight  to  the  better  values.  The  plotting 
for  the  Margules  equation  is  better  in  this  respect  and  tends  to  weigh  all  the  values 
about  equally. 


Intercept  of  plot  =  0.056 
Slope  of  plot  -  0.097  = 


1 


Thus,  B  =  -319  and  A  -  1.73. 
given  in  Table  3-9. 


Similar  values  for  the  other  component  are 


TABLE  3-9 


Xi 

y* 

72 

1 

£2 

Xi 

(-7Tln7*)H 

0.9145 

0  0478 

0.610 

0.078 

0  093 

0  8590 

0  0835 

0  643 

0.083 

0  164 

0  7955 

0  1312 

0  675 

0.087 

0  294 

0.7388 

0  1765 

0.690 

0  090 

0  354 

0.6633 

0  2495 

0  724 

0.096 

0  410 

0  5750 

0  3520 

0.782 

0.098 

0.740 

0.4771 

0.4830 

0.841 

0.131 

0.095 

0.3350 

0.6650 

0.884 

0.155 

1.987 

0.2660 

0.7630 

0.925 

0.195 

2  76 

0  2108 

0.8240 

0  957 

0.257 

3  62 

0.1375 

0  900 

0  964 

0.294 

6  27 

0  1108 

0.9350 

0.985 

0  445 

8  02 

From  Fig.  3-7, 

Intercept  is  0.074  =  (  ~  g) 
Slope  is  0.0434  =  _AB% 

Thus  B  -  -310  and  A  -  1.72.  Average  A  =  1.725  and  average  B  m  -315. 
The  values  of  A  and  B  obtained  here  are,  within  the  accuracy  of  the  method,  the 
same  as  those  obtained  from  the  azeotrope  data  alone,  and  the  y,x  calculations  will 
not  be  repeated. 


70  FRACTIONAL  DISTILLATION 

Solution  of  Part  2a 

NOTE.     Some  values  used  below  are  taken  from  Part  1. 

In  71  «  bx\  -f-  cx\ 
In  72  -  6aJ  +  3Acxl  -  cx\ 
Tl  »  0.764,         72  =  0.889  (from  Part  1) 
Xi  «  0.335,         £2  «  0.665 
-0.269  -  6(0.665)2  +  c(0.665)3 
-0.118  -  6(0.335)2  +  ^c(0.335)2  -  c(0.335)« 

Thus  b  -  -0.0106  and  c  -  -0.894. 

In  Tl  „  -(0.0106  +  0.894£s)g| 
In  72  -  -(1.351  -  0.894zi)z? 

Let  xi  -  0.4771,  xz  -  0.5229, 

In  71  =  -(0.0106  +  0.894a?s)ajJ  -  -0.131 

Tl  -  0.878 
In  72  -  -(1.351  -  Q.894*iX  »  -0.210 

72  -  0.810 

Assuming  T  -  63.91°C., 

yi  .  0.878(970X0.4771)  ^  Q  53g 

0.810(835) (0.5229)  _  n  _- 
yi  »  _  U.4b5 

Sy  -  1.000.  If  the  sum  of  y  is  not  equal  to  1.000,  then  other  values  of  P  should 
be  used  until  y  =  1.0. 

Values  of  y  corresponding  to  various  other  values  of  x  may  be  similarly  calcu- 
lated. The  results  are  given  in  Table  3-10. 

TABLE  3-10 


Xi 

2/loalo 

2/lexp 

100  Ay 

(y   -  X)exp 

0.9145 

0.951 

0.9552 

3.2 

0.7955 

0.869 

0.8688 

0.3 

0.6633 

0.753 

0.7505 

2.9 

0.4771 

0.519 

0.5170 

5.0 

0  3350 

0.3350 

0.3350 

0.0 

0.2108 

0.173 

0.1760 

8.6 

0.1108 

0.072 

0.0650 

15.0 

Solution  of  Part  26. 


Since  In  71  —  bx% 
In  71 


x\ 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 
Also  In  78  ••  bx\  -H 


71 


or,  substituting, 


-cxl 
In  72 


In  72 


b  -f  -  c  - 


,6+ic 


-  cx* 
••  b  +  c(0.5  -f  X*] 


(d) 

(e) 

By  plotting  Eqs.  (c)  and  (e),  a  single  line  should  be  obtained  with  a  slope  —  c, 
and  intercept  at  xz  —  0  equal  to  6.  The  values  for  these  equations  are  given  in 
Table  3-11. 

TABLE  3-11 


Xi 

X2 

In  71 

In  72 

111  71 

In  72 

*S 

x\ 

0.9145 

0.0855 



-0.495 



-0  591 

0  8590 

0.1410 

— 

-0.441 

— 

-0.599 

0  7955 

0.2045 

-0.0091 

-0.394 

-0.217 

-0  621 

0.7388 

0  2612 

-0.0202 

-0  371 

-0.296 

-0.680 

0  6633 

0.3367 

-0  045 

-0  324 

-0.396 

-0.737 

0.5750 

0.4250 

-0.0878 

-0.246 

-0.486 

-0.744 

0.4771 

0.5229 

-0.164 

-0.173 

-0.600 

-0.760 

0  3350 

0.6650 

-0.271 

-0.123 

-0.613 

-1.094 

0.2660 

0.7340 

-0.380 

-0.0780 

-0.706 

-1.101 

0.2108 

0  7892 

-0  420 

-0.0440 

-0.675 

-0.986 

0.1375 

0.8625 

-0.540 

-0.0366 

-0.726 

-1.937 

0.1108 

0  8892 

-0.739 

-0.0152 

-0.936 

-1.239 

The  values  of  Table  3-11  are  plotted  in  Fig.  3-8. 

Slope  of  plot  =  -0.74  =  c 
Intercept  of  plot  «  -0.11  «  b 

Thus,  In  7j  -  -(0.11  +  0.74z2)z2  and  In  72  -  -(0.48  +  0.74x^x1 
Values  of  y  at  various  values  of  x  are  calculated  as  in  Part  2a.     The  results  are 
given  in  Table  3-12. 

TABLE  3-12 


Xi 

T/cfllc 

•^exp 

100  Ay 

(y  —  #)«p 

0.9145 

0.959 

0  9522 

18  0 

0.7955 

0.871 

0.8688 

3.0 

0.6633 

0.753 

0.7505 

2.9 

0.4771 

0.520 

0  5170 

7.5 

0  3350 

0.336 

0.3350 

0.0 

0.2108 

0.175 

0  .  1760 

2.9 

0.1108 

0.076 

0.0650 

15.0 

72 


FRACTIONAL  DISTILLATION 


An  appreciable  portion  of  the  variations  in  100  &y/(y  —  x)  is  probably  due  to 
the  errors  in  the  experimental  calculation  and  the  data.  The  agreement  with  both 
the  Van  Laar  and  the  Margules  equations  is  fairly  good  for  this  system.  The 
azeotrope  point  for  the  Margules  equation  gave  better  agreement  than  the  line 


u 
1.0 
0.3 
*06 
0.4 
0.2 

A 

/ 

A 

/ 

<T 

< 

) 

> 

/* 

/ 

/ 

o 

X 

A 

Qu 

t& 

S 

/^ 

x* 

y 

-/ 

y 

Inxj 
o  Values  of  £|~~     vs  *2 

A  Values  of  -  •—£*•     V3(x2  +  05) 

/ 

x3 

> 

/ 

V0             02            0.4            06            0.8            1.0             1.2             1,4            U 
FIG.  3-8.     Margules  plot  for  system  acetone-chloroform. 

drawn  through  the  points,  but  a  line  drawn  on  the  basis  of  the  azeotrope  constants 
would  agree  with  the  plotted  points  as  well  as  the  line  employed.  In  drawing  the 
plot  for  the  Margules  evaluation  no  attempt  was  made  to  place  the  line  to  give  the 
minimum  average  deviation  in  100  Ay/  (y  —  x).  Mason  studied  the  "  optimum 
line"  and  obtained  constants  that  gave  a  maximum  deviation  of  12.5  per  cent  and 
an  average  deviation  of  4.3  for  100  A^/(y  —  x). 


MULTICOMPONENT  SYSTEMS 

The  ideal  solution  laws  such  as  Raoult's  law  and  Raoult's  law  cor- 
rected for  gas  law  deviation  are  applicable  to  binary  or  multicompo- 
nent  systems.  They  treat  each  component  independently  of  any 
other  component  present;  i.e.,  the  relationship  between  the  mol  frac- 
tion in  the  vapor  and  in  the  liquid  for  a  given  component  depends  only 
on  the  temperature  and  total  pressure.  In  many  cases,  these  simpli- 
fied rules  are  not  applicable,  and  there  is  interreaction  between  the 
various  components.  It  would  be  particularly  desirable  to  have  a 
satisfactory  theoretical  approach  to  the  problem  of  multicomponent 
vapor-liquid  equilibria  since  the  experimental  determination  for  this 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  73 

case  is  an  order  of  magnitude  more  difficult  than  for  binary  mixtures. 
The  equations  as  originally  given  by  Margules  and  Van  Laar  were 
limited  to  binary  mixtures,  although  Van  Laar  did  indicate  the  method 
for  multicomponent  problems.  In  recent  years,  there  have  been  a 
number  of  derivations  of  multicomponent,  Van  Laar-type  equations 
(Refs.  1,  7,  22,  26,  27).  For  example,  Bonham  (Ref.  1)  obtained  a 
multicomponent  Van  Laar  equation  by  using  multicomponent  mixture 
constants  for  the  van  der  Waals  equation,  amix  and  Zw,  given  by 
Eq.  (3-40). 

A  similar  derivation  to  that  employed  for  the  binary  Van  Laar 
equation  leads  to 


(3-58) 


T  In  73  = 


In  the  above  equations,  the  values  of  A  and  B  are  similar  to  those 
given  for  a  binary,  but  in  this  case,  considerable  care  must  be  exer- 
cised with  respect  to  the  subscripts.  Thus  A  12  is  equal  to  61/62,  A  32  is 
equal  to  63/62,  and  similarly  for  other  subscripts.  It  should  be  pointed 
out  that  for  a  ternary  mixture,  there  are  only  two  independent  A 
terms.  Any  other  A  terms  can  be  calculated  from  these  two  by 
multiplication  or  division.  The  three  activity  coefficient  equations 
contain  only  the  values  of  A  and  B  associated  with  the  three  binary 
mixtures  possible  from  the  three  components.  If  the  Van  Laar  equa- 
tion for  multicomponent  mixtures  is  applicable,  the  only  information 
needed  is  the  vapor-liquid  equilibrium  data  for  the  binary  mixtures. 

In  the  case  of  the  value  of  5,  it  should  be  noted  that  it  occurs  in  the 
multicomponent  equation  as  a  square  root,  and  this  immediately 
raises  the  question  of  whether  the  value  is  positive  or  negative.  This 
question  can  be  answered  by  considering  the  relationships  given  on 
page  58.  On  the  basis  of  the  relations  employed  for  amix  and  6mi«, 
the  square  root  of  Bn  is  equal  to 


B 

and  the  value  is  negative  or  positive  depending  on  whether 

is  greater  or  less  than  vW&2.    There  are  at  least  two  ways  of  deter- 


74  FRACTIONAL  DISTILLATION 

mining  the  sign  of  the  \/B.  The  first  involves  having  sufficient  data 
on  the  multicomponent  mixture  in  order  that  the  three  equations  can 
be  forced  to  fit.  This  method  is  of  little  value  since  it  requires  a  large 
amount  of  data  on  the  mixture,  and  the  mathematical  procedure  is 
complicated.  It  is  useful,  however,  in  cases  where  data  are  available, 
and  it  is  desirable  to  interpolate  or  extrapolate.  The  other  method  is 
based  on  evaluating  the  sign  of  these  terms  independent  of  vapor- 
liquid  equilibria  data  and  determining  their  magnitude  from  binary 
data. 

Applying  the  Van  Laar  to  a  binary  mixture  does  not  indicate  the 
sign  of  the  -\/B.  The  basis  for  deciding  the  sign  is  the  relative  values 
of  the  Va/6?  which  corresponds  to  the  square  root  internal  pressure  of 
the  liquid.  Thus  polar  compounds  which  have  high  internal  pressures 
would  be  expected  to  have  high  values  of  this  group,  while  compounds 
of  low  polarity  would  be  expected  to  have  low  values.  If  the  com- 
pounds in  the  binary  mixture  are  of  widely  different  polarity,  it  is 
fairly  easy  to  determine  whether  the  square  root  should  be  positive  or 
negative.  For  example,  for  a  mixture  of  ethyl  alcohol  and  water, 
components  1  and  2,  respectively,  it  is  well  established  that  water  is 
the  more  polar;  therefore,  if  V^i  is  taken  to  be  positive,  \/#i2  would 
be  negative. 

If  the  square  roots  of  the  £12/61,  523/&2,  and  Bsi/fes  are  added,  it  will 
be  found  that,  on  the  basis  of  the  definition  given  on  page  58,  the 
sum  is  zero.  In  fact,  all  that  is  necessary  to  obtain  this  conclusion  is 
to  have  each  one  of  the  subscripts  appear  first  on  one  of  the  J5's  and 
last  on  another,  and  to  have  the  subscript  on  the  6  correspond  to  the 
first  subscript.  These  relationships  are  extremely  useful  since  any 
two  independent  values  of  B  together  with  the  corresponding  A's  are 
sufficient  to  evaluate  the  other.  Thus,  for  a  ternary  mixture,  it  is 
necessary  to  have  data  on  only  two  of  the  binaries.  This  is  useful  in 
several  ways,  e.g.,  predicting  the  three-component  data  from  data  on 
two  of  the  binaries  or  predicting  the  vapor-liquid  equilibria  for  a 
binary  for  which  there  are  no  data.  In  this  latter  case,  it  is  necessary 
to  find  data  on  two  binaries  which  have  a  common  component,  the 
other  components  being  the  ones  for  the  desired  mixture.  For  exam- 
ple, if  data  are  available  on  ethanol  and  water  and  methanol  and 
water,  it  is  possible  to  calculate  the  vapor-liquid  equilibria  for  methanol 
and  ethanol  on  the  basis  of  these  data,  assuming  that  the  systems 
agree  with  the  Van  Laar-type  equation. 

The  various  relations  of  the  A's  and  B's  of  the  Van  Laar  equations 
for  a  ternary  mixture  are  summarized  below: 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


75 


or,  alternatively, 


In  the  case  of  more  than  three  components,  the  relationship  is 
obtained  by  the  same  procedure,  but  it  is  relatively  easy  to  write  the 
activity  coefficient  equations  simply  by  inspection.  Returning  to  the 
three  component  equations,  Eq.  (3-58),  it  is  noted  that  the  denominator 
is  always  the  same  and  simply  involves  the  square  of  the  sum  of  the 
mol  fraction  times  the  A  for  the  term  in  question  relative  to  some  given 
base  component.  In  the  equations  given,  component  2  is  used  as  the 
base,  and  -4.22  was  omitted  since  it  is  obviously  1.  The  numerator 
involves  terms  for  the  components  other  than  the  one  under  considera- 
tion. These  terms  are  all  of  a  general  type  and  can  best  be  explained 
by  considering  the  specific  equations.  Thus,  for  the  T  In  71,  there 
are  terms  involving  #2  and  #3,  each  term  has  the  A  value  corresponding 
to  its  component  and  the  \/B  for  component  1  relative  to  it.  The 
same  rules  apply  to  T  In  72  and  Tin  73  with  appropriate  shifts  of  sub- 
scripts. For  n  components,  the  equations  become 


-f 


)2,  CtC. 


-f  OV 
\/Bn 


(3-61) 


Tlnyn 


and  the  following  relationships  apply: 


A  large  number  of  equations  of  this  latter  type  can  be  written  as 
long  as  the  number  for  each  component  appears  first  in  the  subscript 
of  one  B  and  last  in  another  subscript,  and  the  A  terms  are  made  to 
correspond  to  the  first  subscript  of  the  associated  B  term.  Depending 


76  FRACTIONAL  DISTILLATION 

on  what  particular  values  of  B  are  available,  there  may  be  a  preference 
for  some  particular  series.  The  other  relationships  for  the  A's  and 
B's  are  the  same'as  listed  for  the  ternary  system. 

To  use  these  equations  with  the  evaluation  of  the  constants  from 
binary  data,  it  is  necessary  to  have  information  on  n  —  1  binary 
mixtures. 

As  was  shown  in  a  previous  section,  the  application  of  Van  Laar's 
equation  to  a  binary  mixture  showing  negative  deviations  from 
Raoult's  law,  i.e.,  tendency  to  form  a  maximum  boiling  mixture,  gives 
negative  values  for  the  B  term.  If  some  of  the  binary  mixtures  give 
negative  values  and  some  give  positive  values,  they  cannot  be  used  in 
Eq.  (3-58).  If  all  the  B  terms  are  either  negative  or  positive,  the 
equation  can  be  applied.  If  any  of  the  binary  systems  involved  does 
not  agree  with  Van  Laar's  equation,  then  the  multicomponent  relation- 
ship should  not  be  applied  in  the  region  near  this  binary.  Multicom- 
ponent equations  of  the  Margules  type  have  been  presented,  but  they 
involve  so  many  independent  constants  that  their  engineering  utility 
is  small. 

Cooper's  relationship  (see  page  59)  can  be  applied  to  multicom- 
ponent mixtures.  Thus  for  a  three-component  mixture  Eq.  (3-47) 
becomes 

'  f  ,«  QON 


_ 

n\vi  +  UzVz  +  n&s  (H^VI  +  n&z  + 

and,  by  partial  differentiation  and  substituting  for 

T  1  =  xigl2 

71 
T  1         -  xlAltBti  +  xlAlzBw  +  xiXz(A*JBu  + 

T2 


+  x*  + 
,  x\A\iBzi  ' 

1     in  *Y3    == 


:  -  ; 

+  £2  + 

(3-64) 

In  these  equations  the  A  values  are  related  in  the  same  manner 
as  for  Eq.  (3-58),  but  no  assumption  has  been  made  relative  to  the 
values  of  the  B  terms.  To  be  consistent  with  the  binary  equations, 
Bn  =  AwBzi,  etc.,  and  this  leaves  three  independent  B  terms.  If  the 
values  of  these  terms  are  made  to  fit  Eqs.  (3-59)  and  (3-60),  then  Eq. 
(3-64)  reduces  to  Eq.  (3-58).  However,  the  values  can  be  used  inde- 
pendently, and  the  equation  then  is  more  general  than  Eq.  (3-58).  If 
binary  data  are  used  to  evaluate  the  constants,  information  on  all 
three  binaries  must  be  available,  and  frequently  this  limits  the  useful- 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  77 

ness  of  Eq.  (3-64)  as  compared  to  (3-58).  Negative  values  of  B  may 
lead  to  mathematical  difficulties  in  the  use  of  Eq.  (3-58),  but  they  can 
be  handled  by  Eq.  (3-64).  Because  Eq.  (3-58)  requires  less  experi- 
mental data  to  evaluate  the  constants  and  is  easier  to  use,  it  has  been 
more  widely  applied  than  Eq.  (3-64). 

Nomenclature 

A  ,B  «  constants  in  Van  Laar  equation 

E  —  internal  energy 

AF  =  partial  molal  change  in  free  energy 
AF0  =  actual  value  of  AF 
AFt-  =  ideal  value  of  AF 

AFe  =s  excess  partial  molal  free  energy  change  »  A/*0  —  &Pi 
AF€  =  excess  free-energy  change  on  mixing 

/  =  fugacity 

/L  =  fugacity  of  component  in  liquid  phase 
fp  =s  fugacity  of  pure  liquid  under  its  own  vapor  pressure 
/*  =  fugacity  of  pure  liquid  under  total  pressure  of  mixture 
fv  =  fugacity  of  component  in  vapor  phase 
fv  as  fugacity  of  pure  vapor  at  pressure,  TT 
Aj?  «  partial  molal  change  in  enthalpy 
AHe  —  excess  partial  molal  change  in  enthalpy 

A/T  =  enthalpy  of  vaporizating  1  mol  of  a  component  from  a  liquid  into  a  vacuum 
K  as  equilibrium  constant  =  y/x 
K'  =  proportionality  constant 
n  =*  mols 

p  =  vapor  pressure 
p  =  partial  pressure 
R  =  gas  law  constant 
AS  =  partial  molal  change  in  entropy 
ASe  —  excess  partial  molal  entropy  change 
T  —  absolute  temperature 
y  =  volume 

t;  =  molal  volume  or  partial  molal  volume 
x  =»  mol  fraction  in  liquid 
y  «  mol  fraction  in  vapor 
a  «  relative  volatility 
0  =*  volatility  ==  p/x 
y  «•  activity  coefficient 
TT  =  total  pressure 
fi  =  gas  law  correction  factor  »  PV/RT 

Subscripts: 

a,6,l,2,3  refer  to  components 
L  refers  to  liquid  phase 
R  refers  to  reduced  conditions 
V  refers  to  vapor  phase 


78  FRACTIONAL  DISTILLATION 


References 

1.  BONHAM,  M.S.  thesis  in  chemical  engineering,  M.I.T.,  1941. 

2.  BEBTHELOT,  Compt.  rend.,  126,  1703,  1857  (1898). 

3.  BBOWN,  J.  Chem.  Soc.,  39,  304,  517  (1881). 

4.  CABBY,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1929. 

5.  CAUBET,  Compt.  rend.,  130,  828  (1900). 

6.  CLABK,  Trans.  Faraday  Soc.,  41,  718  (1945);  42,  742  (1946). 

7.  COOPEB,  10.90  Report,  M.I.T.,  1941 

8.  DALTON,  Gilberts  Ann.  Physik,  12,  385  (1802). 

9.  DUHEM,  Compt.  rend.,  102,  1449  (1886). 

10.  GALITZINE,  Wied.  Ann.  Physik,  41,  770  (1890). 

11.  HILDEBBAND,  "Solubility  of  Non-electrolytes,"  2d  ed.,  Reinhold  Publishing 
Corporation,  New  York,  1936. 

12.  "International  Critical  Tables,"  Vol.  Ill,  pp.  212-217,  McGraw-Hill  Book 
Company,  Inc.,  New  York,  1928. 

13.  LECAT,  "L'Azeotropisme,"  1918. 

14.  LEE,  /.  Phys.  Chem.,  36,  3554  (1931). 

15.  LEHFELDT,  London,  Edinburgh  Dublin  Phil.  Mag.,  6,  246  (1895). 

16.  LEWIS  and  RANDALL,  "Thermodynamics,0  p.  226,  McGraw-Hill  Book  Com- 
pany, Inc.,  New  York,  1923. 

17.  MABQULES,   "  Sitzungsberichte  der  math-naturw,"   Classe  der  Kaiserlichen 
Akademie  der  Wissenschaften  (Vienna),  104,  1243  (1895). 

18.  MASON,  M.S.  thesis  in  chemical  engineering,  M.I.T.,  1948. 

19.  RAOITLT,  Compt.  rend.,  104,  1430  (1887). 

20.  RHODES,  WELLS,  and  MUBBAY,  Ind.  Eng.  Chem.,  17,  1200  (1925). 

21.  ROSANOFP  and  EASLEY,  /.  Am.  Chem.  Soc.,  31,  977  (1909). 

22.  SCATCHABD,  Trans.  Faraday  Soc.,  33,  160  (1937). 

23.  SCHEELINE,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1938. 

24.  TUCKEB,  M.S.  thesis  in  chemical  engineering,  M.I.T.,  1942. 

25.  VANLAAB,  Z.  physik.  Chem.,  72,  723  (1910);  83,  599  (1913). 

26.  WHITE,  Trans.  Am.  Inst.  Chem.  Engrs.,  41,  539  (1945). 

27.  WOHL,  Trans.  Am.  Inst.  Chem.  Engrs.,  42,  215  (1946). 

28.  ZAWIDSKJ,  Z.  physik  Chem.,  35,  129  (1900). 


CHAPTER  4 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 

(Continued) 

Critical  Regions.  At  very  high  pressures  special  phenomena  asso- 
ciated with  the  critical  region  are  encountered  in  vapor-liquid  equi- 
libria. If  the  vapor  pressure  of  a  pure  component  is  plotted  vs.  the 
temperature,  a  line  concave  upward  is  obtained.  This  line  terminates 
at  the  critical  point.  Conditions  below  the  line  in  region  A,  Fig.  4-1, 


35 


30 


I  25 


20 


95 


120 


220 


245 


145  170  195 

Temperature,  °£ 

FIG.  4-1.     Typical  pressure-temperature  curves. 

correspond  to  all  vapor  and  no  liquid.  The  line  represents  conditions 
under  which  the  vapor  and  liquid  are  in  equilibrium.  Conditions 
above  the  line  in  region  B  represent  all  liquid.  In  region  C  the  state 
of  the  substance  is  in  question  since  it  is  possible  to  obtain  either  vapor 
or  liquid  without  a  change  in  phase.  If  a  given  binary  mixture  is 
plotted  in  the  same  way,  similar  conditions  are  attained  except  that  a 
loop  region  is  obtained  for  a  mixture  of  given  composition  instead  of  a 
single  line,  Fig.  4-1.  The  upper  line  of  the  loop  represents  the  bubble- 
point  curve,  i.e.,  the  condition  under  which  the  mixture  first  forms 
vapor.  The  lower  side  of  the  loop  is  the  dew  point  curve,  i.e.,  the 
condition  under  which  the  mixture  begins  to  condense.  In  the  case 

79 


80 


FRACTIONAL  DISTILLATION 


of  the  pure  substance  these  two  lines  coincide,  and  the  critical  tem- 
perature and  pressure  are  the  maximum  for  both  variables  that  can 
exist  and  have  coexistence  of  liquid  and  vapor  phase.  In  the  case  of 
the  mixture  the  maximum  temperature  does  not  coincide  with  the 
maximum  pressure. 

For  a  pure  component  at  the  critical  condition,  the  properties  of  the 
vapor  and  the  liquid  phases  become  identical  in  all  respects.  In  the 
case  of  the  loop  curve  for  a  binary  mixture,  the  property  of  the  vapor 

40r 


95 


120 


220 


245 


145          170          195 

Temperature,  °C 
FIG.  4-2.     Pressure-temperature  curves  for  butane-hexane  system. 

and  the  liquid  are  in  general  different,  both  at  the  point  of  maximum 
temperature  and  at  the  point  of  maximum  pressure.  There  is  a  point 
on  the  loop,  usually  between  the  maximum  temperature  and  the  maxi- 
mum pressure  points,  at  which  the  properties  of  the  liquid  and  the 
vapor  are  identical.  This  is  taken  as  the  critical  of  the  mixture.  This 
critical  point  K  is  shown  on  Fig.  4-1.  Figure  4-2  shows  loop  curves 
for  three  different  mixtures  of  butane  and  hexane,  and  a  curve  indi- 
cating the  loci  of  critical  points  is  shown.  If  the  conditions  inside  a 
single  loop  curve  are  analyzed,  it  is  found  possible  to  plot  lines  of  con- 
stant fraction  vaporized;  such  curves  for  0,  20,  40,  60,  80,  and  100  per 
cent  vapor  have  been  drawn  in  Fig  4-3.  All  these  curves  converge  to 
a  common  point  at  the  critical. 

The  curves  of  Fig.  4-3  indicate  the  phenomena  of  retrograde  con- 
densation, i.e.,  conditions  under  which  an  increase  in  pressure  causes 
vaporization  instead  of  condensation,  or  in  which  a  decrease  in  tem- 
perature causes  vaporization  instead  of  condensation.  Thus,  start- 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


81 


ing  at  point  A,  Fig.  4-3,  and  increasing  the  temperature  at  constant 
pressure,  the  bubble-point  curve  is  first  contacted  and  vapor  begins 
to  form.  An  increase  in  temperature  causes  more  vaporization  up  to 
about  20  per  cent.  Higher  temperatures  then  cause  a  decrease  in 
vapor  until  the  mixture  is  again  all  liquid  at  point  C.  A  similar  type 
of  phenomenon  is  exhibited  by  curve  EL.  If  at  constant  temperature 
the  pressure  is  lowered  from  point  E,  the  conditions  first  reach  the 
dew-point  curve;  i.e.,  the  mixture  is  all  vapor.  On  lowering  the  pres- 


A 


Temperature 
FIG.  4-3.     Typical  pressure-temperature  loop  curve. 

sure,  the  mixture  becomes  more  and  more  liquid  until  about  25  per 
cent  is  condensed.  In  this  region  it  has  exhibited  the  retrograde 
phenomena.  At  still  lower  pressures,  the  mixture  behaves  normally 
and  vaporizes  with  decrease  in  pressure. 

In  a  binary  mixture,  if  two  of  the  loop  curves  intersect,  i.e.,  if  the 
vapor  curve  of  one  crosses  the  liquid  curve  of  the  other,  then  the  two 
compositions  determine  a  vapor-liquid  equilibrium  point.  This  is  due 
to  the  fact  that,  for  a  binary  system  of  two  phases,  the  phase  rule 
allows  two  degrees  of  freedom.  However,  the  value  may  not  be 
unique,  i.e.,  in  the  higher  pressure  region,  particularly  very  near  the 
critical,  it  is  possible  for  a  given  vapor  to  have  two  possible  equilibrium 
liquids  of  different  compositions.  These  two  conditions  can  be  at  the 


82 


FRACTIONAL  DISTILLATION 


0.2  0.4  06  08 

Mol  fraction  C02  in  liquid 

(a) 


0.2  0.4  06  OB  1.0 

Mol  fraction  C02  in  liquid 

W 

FIG,  4-4.    a,  Vapor-liquid  equilibria  for  C02-S02  at  constant  pressure.    6.  Vapor-liquid 
equilibria  for  COa-SQj  at  constant  temperature. 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


83 


same  temperature,  but  at  different  total  pressures.  This  is  shown  in 
Fig.  4-4  which  gives  the  vapor-liquid  equilibrium  data  for  the  system 
carbon  dioxide-sulfur  dioxide. 

In  general,  it  is  found  that  the  relative  volatility  decreases  as  the 
total  pressure  increases,  and  there  are  several  factors  that  combine  to 
give  this  effect.  It  is  generally  found  that  (1)  the  ratio  of  the  vapor 
pressures  of  two  components  becomes  nearer  to  unity  as  the  pressure 
increases  and  (2)  the  deviations  from  the  perfect-gas  laws  also  tend  to 
reduce  the  relative  volatility.  Actually  the  decrease  in  relative 


0.2  0.4  0.6  0.8  1.0 

x=  Mol  Fraction  €4  in  Liquid 
FIG.  4-5.     Effect  of  pressure  on  vapor-liquid  equilibria. 

volatility  is  even  larger  than  the  result  of  these  two  factors.  At  the 
critical  pressure  of  a  binary  mixture,  the  vapor  and  liquid  phases 
become  identical  in  all  respects,  and  the  relative  volatility  becomes 
unity.  No  separation  is  possible,  not  only  because  of  the  relative 
volatility  effects  but  because  differentiation  between  the  vapor  and 
liquid  is  no  longer  possible.  This  decrease  in  relative  volatility  to  the 
value  of  1  at  the  critical  is  a  progressive  effect,  although  a  large  part  of 
it  occurs  close  to  the  critical  condition.  The  Lewis  and  Randall 
fugacity  rule  does  not  show  the  convergence  of  the  relative  volatility 
to  unity  at  the  critical  For  example,  in  the  case  of  a  mixture  like 
ethane  and  propane,  the  rule  would  show  a  finite  relative  volatility  not 
only  at  the  critical  condition  but  at  pressures  much  higher. 

This  effect  of  total  pressure  on  the  vapor-liquid  equilibria  of  the 


84  FRACTIONAL  DISTILLATION 

system  rc-butane-ft-hexane  (Ref.  1)  is  shown  in  Fig.  4-5.  The  decrease 
in  the  relative  volatility  with  increasing  pressure  is  apparent,  and  the 
y,x  curves  become  discontinuous  at  pressures  above  30  atm.  Thus, 
at  33.5  atm.  the  y,x  curves  exist  only  for  liquids  containing  more  than 
24  mol  per  cent  butane.  Mixtures  containing  less  than  this  amount  of 
butane  are  at  pressures  above  the  envelope  curve  of  Fig.  4-2,  and  only 
a  single  phase  is  present.  At  higher  pressures  the  range  of  the  dis- 
continuity increases,  and  above  37.5  atm.  the  y,x  curve  becomes  dis- 
continuous at  both  ends  because  there  are  mixtures  of  butane  and 
hexane  which  have  higher  critical  pressures  than  either  of  the  pure  com- 
ponents. Above  38  atm.  only  one  phase  is  present  for  all  compositions. 

From  the  viewpoint  of  the  ease  of  separation,  it  is  almost  always 
disadvantageous  to  operate  at  high  pressure,  but  it  is  frequently 
necessary  to  accept  this  more  difficult  separation  in  order  to  obtain 
other  desirable  features  of  high  pressure,  such  as  higher  condensation 
temperatures  and  lower  volume  of  apparatus. 

The  relative  volatilities  of  all  vapor-liquid  systems  do  not  decrease 
with  increasing  pressure  in  all  regions.  Thus,  it  is  possible  for  abnor- 
mal mixtures  to  have  an  increase  in  relative  volatility  in  some  regions 
for  an  increase  in  pressure.  It  is  in  general  true  that  they  will  not 
increase  the  relative  volatility  at  all  compositions.  For  example,  in 
the  case  of  ethyl  alcohol  and  water  at  atmospheric  pressure,  the  con- 
stant-boiling mixture  is  about  89  mol  per  cent  alcohol.  As  the  pressure 
is  increased,  the  composition  of  the  constant-boiling  mixture  becomes 
lower  in  alcohol,  and  the  relative  volatility  of  water  to  alcohol  at  a 
composition  of  89  mol  per  cent  becomes  greater  than  1  as  the  pressure 
is  increased  above  1  atm.  and  then  decreases  at  still  higher  pressures. 

The  approach  of  the  relative  volatility  to  unity  at  the  critical  means 
that  the  compositions  of  the  vapor  and  liquid  are  identical.  Thus  the 
K  values  for  all  components  equal  1.  The  temperature  and  pressure 
at  which  these  values  become  unity  are  functions  of  the  other  com- 
ponents present.  Thus,  a  mixture  of  butane  and  ethane  would  have  a 
certain  critical  temperature  and  pressure,  while  a  mixture  of  butane 
with  hexane  would  have  different  critical  temperature  and  pressure, 
but  under  both  conditions  the  K  value  for  butane  would  have  to  be 
equal  to  unity.  Thus,  the  values  given  in  Table  3-2,  which  were  taken 
to  be  independent  of  the  character  of  the  other  components  and  a 
function  of  the  temperature  and  pressure  only,  cannot  apply  in  the 
critical  region.  In  most  cases,  these  effects  of  the  critical  region  are 
not  serious  at  total  pressures  less  than  0.5  to  0.7  of  the  critical  pressure. 
Modifications  of  the  method  of  utilizing  the  K  values  in  the  critical 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  85 

region  have  been  suggested  which  allow  for  the  effect  of  the  other 
components  present  (Ref.  2). 

Immiscible  Liquids.  Immiscible  liquids  are  not  an  important  case 
encountered  in  fractional  distillation.  It  is  much  simpler  to  separate 
two  liquids  which  are  insoluble  in  each  other  by  simple  decantation 
than  it  is  by  fractionation.  However,  the  physical-chemical  laws 
that  apply  to  such  cases  are  helpful  in  explaining  certain  of  the  phe- 
nomena involved  in  the  intermediate  case  of  partially  miscible  liquids. 

The  various  rules  developed  for  the  vapor  phase  for  miscible  liquids 
apply  equally  well  to  this  case.  In  the  case  of  the  liquid  phase,  if  the 
liquids  are  completely  immiscible,  at  equilibrium  each  component 
would  exert  its  own  vapor  pressure  independent  of  the  others  present. 
For  this  case,  the  vapor-liquid  equilibrium  expression  will  be 

pi  **Vi*  =Pi) 

p*  =  wr  -  P2J  (4'1} 

at  high  pressures, 

A  =  yiAi  =  /Pie"i< 


where  p  =  partial  pressure 

P  =  vapor  pressure 

TT  =  total  pressure 

y  =  mol  fraction  of  vapor 

/  =  fugacity 

/*.  =  fugacity  of  pure  component  at  total  pressure 
fp  =  fugacity  of  pure  liquid  under  its  own  vapor  pressure 

v  =  partial  molal  volume 

R  =  gas-law  constant 

T  =  absolute  temperature 

These  equations  are  similar  to  those  for  miscible  liquids  except  that 
the  mol  fraction  in  the  liquid  is  omitted.  It  is  to  be  expected  that  the 
fugacity  relationship  would  give  satisfactory  results  up  to  a  pressure 
of  approximately  one-half  of  the  critical  pressure.  At  higher  pres- 
sures the  Lewis  and  Randall  fugacity  rule  for  the  vapor  mixture  would 
tend  to  be  less  satisfactory.  Actually  it  is  doubtful  whether  absolute 
immiscibility  ever  occurs.  However,  there  are  cases  in  which  the 
miscibility  is  so  limited  that  each  phase  would  act  as  essentially  a 
pure  material,  e.g.,  mercury  and  water. 

Example  for  Immiscible  Liquids.     1.  A  two-phase  liquid  mixture  comprised  of 
30  mols  of  toluene  and  70  mols  of  water  existing  as  liquids  at  1  atm.  pressure  and 


86 


FRACTIONAL  DISTILLATION 


60°C.  is  heated  at  constant  pressure.  Assuming  that  water  and  toluene  arc 
completely  nonmiscible  and  that  all  the  vapors  formed  were  at  all  times  in  equilib 
rium  with  the  remaining  liquids,  and  using  the  vapor  pressure  data  given  below 
construct  the  following  curves: 

a.  Mol  per  cent  vaporized  vs.  temperature. 

6.  Mol  fraction  of  toluene  in  vapor  vs.  temperature. 

2.  A  two-phase  liquid  mixture  consisting  of  30  mols  of  toluene  and  70  mols  o 
water  is  vaporized  at  a  constant  temperature  of  85°C.  by  reducing  the  total  pres 


h 

> 
\  1000 

) 

„ 

' 
5 

IOCJ 

\ 

\ 

\ 

\ 

^ 

Watet 
s. 

x 

\ 

^\ 

\ 

\  \ 

\ 

\ 

Toluene  "' 

-•*\ 

\ 

\\ 

\ 

\ 

^ 

^ 

\ 

\ 

.0            2.2             2.4             26             28             30            3.< 
1000 
"W 

FIG.  4-6.     Vapor  pressures  of  toluene  and  water 

sure.     The  initial  pressure  is  2  atm.  absolute.     Assuming  that  the  vaporization 
carried  out  such  that  equilibrium  between  vapor  and  liquid  exists  at  all  times,  plol 

a.  Mol  per  cent  vaporized  vs.  pressure. 
6.  Composition  of  vapor  vs.  pressure. 

Solution  of  Part  1.  System:  30  mols  toluene,  70  mols  H20.  (Let  subscripts 
and  W  refer  to  toluene  and  water,  respectively.) 

Assuming  complete  immiscibility,  from  the  phase  rule  it  follows  that  for  tl 
three-phase-two-component  system  there  exists  only  one  degree  of  freedoi 
Hence,  vaporization  will  occur  at  constant  temperature  as  long  as  two  liquid  phas 
are  present, 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


87 


Vaporization  occurs   when  v  «  Pw  +  PT  «•  760  mm.     From   Fig.   4-6  this 
temperature  »  84.4°C. 


PT  «  336  mm.        yT  -  3a^feo  -  0.442 
Pw  «•  424  mm.          TT  «  760  mm. 

Both  water  and  toluene  will  vaporize,  but  the  composition  of  the  vapor  will 
remain  constant  at  yr  •»  0.442  as  long  as  two  phases  are  present.  Since  the  ratio 
of  the  vapor  pressures  of  toluene  to^water  is  greater  than  the  ratio  in  the  charge, 
the  toluene  liquid  phase  will  disappear  first.  When  all  of  the  toluene  has  just 
vaporized,  the  water  vaporized  will  be 

30  X  424-is6  -  37.9  mois 
Per  cent  vaporized  »  37.9  4-  30  «  67.9 

At  70  per  cent  vaporized,  the  vapor  will  contain  30  mols  of  toluene  and  40  mols  of 
water. 

Pw  -  760(4?<f  0)  -  434  mm.        t  -  85°C. 
»  0.425 


The  following  table  was  prepared  in  this  manner: 


r,  °c. 

Per  cent 
vaporized 

yT 

60 

0 



70 

0 

— 

80 

0 

— 

84.4 

0 

— 

84.4 

67.9 

0.442 

85 

70 

0.429 

87.4 

80 

0.375 

89 

90 

0  333 

90  3 

100 

0  300 

These  results  are  plotted  in  Fig.  4-7. 

Solution  of  Part  2.     Basis:  30  mols  toluene,  70  mols  HaO.     Vaporization  will  not 
occur  until  the  total  pressure  is  equal  to  the  sum  of  the  partial  pressures.    At  85°C., 

PT  »  341  mm. 

Pw  -  434mm. 

IT  «  775  mm. 

Reasoning  as  in  Part  1,  all  of  the  toluene  will  be  vaporized  at  this  pressure 
together  with  43^4i  X  30  =  38.2  mols  of  H20. 


Per  cent  vaporized 


30  +  38.2  -  68.2 
30 
68.2 


0.440 
In  order  to  vaporize  the  remaining  H20,  the  pressure  must  be  lowered.    Let 


88 


FRACTIONAL  DISTILLATION 

N  «•  per  cent  vaporized 
-  N  ~  3°.  _  30 

yw ~~  •      yT  _  _ 


Using  this  relationship, 


434JV 

AT  -  30      N  -  30 


AT  (%  vaporized) 

TT,  mm. 

VT 

70 

759 

0  429 

80 

694 

0.375 

90 

651 

0.333 

100 

620 

0.300 

The  results  of  these  calculations  are  plotted  in  Fig.  4-8. 


IUU 

80 

§ 

^  60 
c 

T3 

<u 

N 

'§_40 
e 

20 
°6 

J 

L 

/  %  vapor/zed 

o 

^ 

\ 

^ 

J 

0                 70                  80                  90                10 
Temperature,  °C 

Fio.  4-7. 

Partially  Miscible  Liquids.  There  are  a  large  number  of  systems  in 
which  the  components  are  miscible  only  over  limited  ranges  of  con- 
centration. These  mixtures  form  a  very  important  group  for  frac- 
tional distillation.  The  fact  that  the  liquids  are  partially  miscible 
greatly  alters  the  normal  type  of  vapor-liquid  relationships,  and  the 
fact  that  two  liquid  phases  are  present  requires  that  one  less  degree  of 
freedom  be  available  than  is  normal  for  a  system  of  a  given  number  of 
components.  In  general,  the  vapor-liquid  equilibria  of  these  systems 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


89 


are  very  abnormal  and,  by  properly  exploiting  these  abnormalities  and 
combining  distillation  and  decanting  operations,  separations  can  easily 
be  made. 

The  types  of  systems  in  this  category  range  from  those  which  are 
almost  immiscible,  such  as  water  and  benzene,  to  those  that  are  miscible 
except  for  very  limited  regions,  such  as  phenol  and  water.  Even 
though  benzene  and  water  are  essentially  immiscible,  distillation  is 
often  employed  for  the  purpose  of  drying  benzene.  Benzene  saturated 
with  water  contains  a  very  small  percentage  of*  the  latter.  By  the 


IUUS 

80 

»- 
>% 
o 

2  60 

73 

c 

O 
T5 

<D 

|40 

20 

0 
62 

X 

^v, 

^\ 

*^n 

%  vaporized 

^ 

^^, 

•^-^ 

^^  —  ^9 

^T 

—  " 

*-  

^  —  / 

\<  — 

—  -A—* 

' 

.  —  -A- 

>0    640    660    680    700     720    740    760    780    800   82 
Total  pressure  mmHg 

FIG.  4-8. 

proper  distillation  technique  this  water  can  be  completely  and  easily 
removed.  This  type  of  operation  is  employed  chiefly  for  drying'mate- 
rials  that  are  partially  miscible  with  water.  If  the  solubility  of  water 
in  the  material  is  low,  distillation  is  an  economical  method  of  drying 
the  liquid  and  is  effective  down  to  the  extremely  low  concentrations 
of  water.  As  an  example  of  the  method  of  estimating  the  vapor-liquid 
equilibria  for  systems  in  which  the  mutual  solubility  is  low,  the  mix- 
ture ethyl-ether  and  water  will  be  considered. 

In  Fig.  4-9  a  constant-temperature  diagram  for  this  system  has  been 
constructed  in  which  the  ordinate  is  the  partial  pressure  of  the  com- 
ponents, and  the  abcissa  is  the  mol  fraction  of  ether  in  the  mixture. 
This  mol  fraction  is  the  mols  of  the  ether  in  the  combined  liquid  phases 
divided  by  the  total  mols  of  ether  and  water  in  all  of  the  liquid  present. 


90 


FRACTIONAL  DISTILLATION 


It  is  a  "pseudo"  mol  fraction.  At  the  left-hand  side  of  the  diagram 
the  water  partial-pressure  curve  starts  at  the  vapor  pressure  of 
water,  and  the  ether  curve  starts  at  zero,  since  none  is  present.  As 
ether  is  added  to  the  system,  it  first  dissolves  in  the  water,  giving  only 
one  liquid  phase,  until  a  concentration  of  0.9  mol  per  cent  ether  is 
reached,  which  is  the  solubility  of  ether  in  water  at  60°C.  In  the 


Water  phase 
See  Fig.  4  -/O 
for  details 


-Two  phases 


Ether  phase  *-* 


2? 


2000 
1600 
1200 
800 
400 

0 

i 

i 

i 

Tota) 

pres. 

$ure^ 

* 

7 

Pa 

rtial 

press 

ure  o 

fetht 

•rS 

/ 

/ 

\ 

Raoult's  law 
for  ether*.  \      , 

/ 

1 

. 

^ 

67% 

wate 

r  — 

* 

| 

/ 

' 

1 

/ 

^ 

t 

/ 

PC 

rrf/a/ 

pres* 

wre 

of  wa 

fer\ 

IX 

h 

'enry 

'$  la 

y  fort  wah 

v  — 

_L 

)              02             0.4            0.6             08            t. 

XE  t  Mol  fraction  of  ether  in  liquid  * 
*  In  the  two-phase  region  mol  fraction  values  are  based  on  total  of  both  phases. 
FIG.  4-9.     Estimation  of  vapor-liquid  equilibria  for  system  ethyl  ether-water  at  60°C. 

region  below  0.9  mol  per  cent  ether,  the  water  partial  pressure  will 
agree  closely  with  that  predicted  by  Raoult's  law  since  there  is  not 
enough  ether  present  in  the  liquid  phase  to  alter  significantly  the  water 
properties.  If  one  component  of  a  binary  mixture  obeys  Raoult's 
law,  the  Duhem  equation  states  that  the  other  equation  must  obey 
Henry's  law,  p  =  Hx.  Thus  in  this  region  it  is  reasonable  to  assume 
Henry's  law  for  ether. 

When  the  amount  of  ether  present  in  the  liquid  exceeds  0.9  mol  per 
cent,  two  phases  form.  The  water  phase  will  consist  of  99.1  mol  per 
cent  water  and  0.9  mol  per  cent  ether.  The  ether  phase  will  consist  of 
93.3  mol  per  cent  ether  and  6.7  mol  per  cent  water.  These  are  the 
solubility  relationship  of  these  two  components  at  a  temperature  of 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


91 


60°C.  Thus,  at  a  mol  per  cent  of  ether  equal  to  0.9,  two  liquid  phases 
will  be  formed.  As  more  and  more  ether  is  added  to  the  system,  the 
amount  of  the  liquid  ether  phase  increases,  but  the  compositions  of  the 
ether  and  the  water  phases  remain  constant.  This  is  in  agreement 
with  the  phase  rule  which  states  that,  for  two-component  systems 
involving  three  phases,  only  one  degree  of  freedom  is  available,  and  in 


^er  phase 

<          "-               Wat 

T1 

1600 
1200 

xt- 

X 

Total  pressure 

^ 

// 

/ 

V, 

/ 

0.9% 
ether  -^^ 

^ 

-Par 
of  el 

fial  p 
herfa 

ressu 
?nr/'s 

re 

400 

x 

// 

x 

/ 

// 

n 

*ial  p 
(Rat 

ressu 
wit's 

<-e  of 
law) 

wate 

r  1  

/ 

's 

ran 

0 

)            OOOZ          0004         0.006         0.008          0.( 

xt  ,  Mol  fraction  of  ether  in  liquid  * 
*  In  the  two-phase  region  mol  fraction  values  are  based  on  total  of  both  phases. 
FIG.  4-10.     Estimation  of  vapor-liquid  equilibria  for  system  ethyl  ether-water  at  60°C, 

this  case  the  temperature  has  been  fixed.  Therefore,  the  composi- 
tions of  all  phases  are  fixed,  as  well  as  the  pressure,  as  long  as  the  three 
phases  are  present. 

On  adding  more  ether,  a  condition  is  finally  reached  at  which  suffi- 
cient ether  is  present  to  dissolve  all  of  the  water,  and  the  water  phase 
disappears.  The  concentration  of  the  phases  just  at  this  condition  is 
the  same  as  it  has  been  throughout  the  two-phase  region.  Thus,  at 
mol  per  cents  of  ether  from  0.9  to  93.3,  two  phases  are  present,  and  the 
partial  pressures  of  water  and  ether  in  the  vapor  are  constant.  Above 
a  mol  percentage  of  ether  of  93.3  only  one  liquid  phase  is  present,  and 
the  system  then  has  two  degrees  of  freedom;  i.e.,  the  vapor  composition 
then  becomes  a  function  of  the  liquid  composition.  For  the  ether 
phase  it  is  logical  to  assume  Raoult's  law  for  the  ether  and  Henry's  law 


92  FRACTIONAL  DISTILLATION 

for  the  water,  but  the  assumptions  are  probably  not  so  good  as  those 
made  for  the  water  phase  because  of  the  high  solubility  of  water  in 
ether.  With  these  assumptions,  the  pressure-composition  diagram 
can  be  completed. 

At  60°C.  the  vapor  pressures  of  ether  and  water  are  1,730  and  149.4 
mm.  Hg,  respectively.  Thus  for  mol  fractions  of  ether  of  0.9  or  less, 
the  partial  pressure  of  water  by  Raoult's  law  is  pw  =  149.4#TF,  and  the 
partial  pressure  of  ether  is  PE  =  HE%E.  For  mol  fractions  of  ether  of 
93.3  or  greater,  ps  =  l,730#j?,  and  pw  =  Hwxw,  where  HE  and  Hw  are 
the  Henry's  law  constants  for  ether  in  water  and  water  in  ether, 
respectively.  The  partial  pressures  of  a  component  must  be  the  same 
in  both  liquid  phases  in  the  two-liquid-phase  region  and  the  above 
equations  can  be  equated. 

For  water, 

149.4(0.991)  -  flV(0.067) 

Hw  =  2,210 
For  ether, 

(0.009)#*  =  1,730(0.933) 
HE  =  179,500 

It  is  interesting  to  note  that  ether  at  a  mol  fraction  of  0.009  in  water 
exerts  a  partial  pressure  equal  to  93.3  per  cent  of  that  of  pure  ether. 
Thus  its  volatility,  p/x,  is  extremely  high. 

The  results  of  such  calculations  are  shown  in  Figs.  4-9  and  4-10. 
The  latter  figure  is  an  expansion  of  the  left-hand  side  of  Fig.  4-9. 
Raoult's  law  for  ether  is  shown  as  a  straight  line  from  the  vapor  pres- 
sure of  ether  at  the  right-hand  side  of  the  diagram  to  zero  at  the  left- 
hand  side.  On  the  basis  of  the  above  assumptions,  this  line  is  used 
only  for  values  x*  from  0.933  to  1.0.  From  xs  =  0.009  to  0.933  the 
partial  pressure  of  ether  is  constant,  and  from  XE  =  0,009  the  partial 
pressure  drops  on  a  straight  line  to  0  at  XE  =  0.  A  similar  construction 
is  used  for  water.  The  sum  of  the  two  partial-pressure  curves  is  the 
total  pressure.  This  is  also  shown. 

If  these  data  are  replotted  as  mol  fraction  of  ether  in  the  vapor  vs. 
mol  fraction  of  ether  in  the  liquid,  assuming  that  the  vapors  obey  the 
perfect-gas  law,  one  obtains  the  results  given  in  Fig.  4-11.  The  value 
of  the  mol  fraction  of  ether  in  the  vapor  increases  very  rapidly  with  the 
mol  fraction  in  the  liquid  and  becomes  constant  at  0.915  in  the  two- 
phase-liquid  region.  The  vapor-liquid  curve  crosses  the  45°  line  at  a 
composition  of  0.915  mol  per  cent  ether.  The  mixture  of  this  compo- 
sition is  a  pseudo-azeotrope.  For  mol  fractions  of  ether  greater  than 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA 


93 


0.915,  water  is  more  volatile  than  ether  in  spite  of  the  fact  that  the 
vapor  pressure  of  ether  is  over  eleven  times  that  of  water.  If  a  liquid 
corresponding  to  a  composition  in  the  ether  phase  region  were  distilled 
at  60°C.,  water  would  tend  to  pass  off  in  the  vapor  leaving  ether  in  the 
still.  Thus,  ether  could  be  dried  by  fractionally  distilling  water 
overhead. 

This  simplified  analysis  is  probably  of  sufficient  accuracy  for  most 
distillation  calculations,  but  it  is  not  suitable  for  cases  of  partially 


Twopha 
' 


phi 


j  ITTV  fJIKJIVOO 

\*f**?*  (note  break  —  >k Ether  phase 

I     in  scale) 


001  092  094  096  098 

x-  Mol  fraction  ether  in  liquid 
*  In  the  two-phase  region  mol  fraction  values  are  based  on  total  of  both  phases. 

FIG.  4-11.     Estimated  vapor-liquid  equilibria  for  ethyl  ether-water  at  60°C. 

miscible  liquids  in  which  the  mutual  solubility  is  much  greater.  The 
experimental  data  on  such  systems  indicate  that  the  partial  pressure 
vs.  mol  fraction  curves  pass  from  the  one-phase  region  to  the  two- 
phase  region  in  a  smooth  type  of  curve;  i.e.,  the  corner  at  the  end  of 
the  horizontal  line  rounds  into  the  Henry  law  region.  The  construc- 
tion used  for  Fig.  4-9  indicated  a  sharp  corner.  This  rounding  effect 
tends  to  make  the  Henry  law  constant  larger.  If  one  component 
obeys  Raoult's  law,  the  Duhem  equation  indicates  that  the  other  com- 
ponent must  obey  Henry's  law,  which  in  the  special  case  may  also  be 
Raoult's  law.  It  follows  that,  in  the  region  where  the  curvature 


94  FRACTIONAL  DISTILLATION 

occurs  and  Henry's  law  does  not  apply,  the  other  components  cannot 
agree  with  Raoult's  law. 

These  effects  are  more  clearly  illustrated  by  mixtures  in  which  the 
immiscibility  is  limited  to  a  narrow  region.  In  such  cases  the  straight- 
line  construction  applied  in  Fig.  4-9  is  entirely  unsatisfactory.  The 
data  of  Sims  (Ref.  3)  for  the  system  phenol-water  for  the  constant- 
temperature  conditions  of  43.4°C.  are  presented  in  Fig.  4-12.  In  this 
ease  the  partial  pressure  of  each  component  divided  by  the  vapor  pres- 


gM 

0> 

|06 
I 

1" 

0.2 
0 

\> 

'Phenol 

S 

^ 

£ 

A 

\ 

/ 

/ 

\ 

/ 

l/vf/,;_  , 

r_^_.       nhr»*A     "I 

Raoulte  law 
for  phenol  -  ' 

^s 

N 

/ 

t 

,  - 

A 

/ 

^ 

^ 

U- 

phases 

/ 

/ 

\ 

\ 

Wat 

'7 

/*-R, 

loult 
>f  wa 

s  la* 
ter 

\ 

c 

0n 

1 

ii 

/ 

/ 

ft 

\ 

\ 

1 

7; 

-/- 

___ 

Phen 

ol  jo/ 

ase' 



- 

\ 

ij 

r 

\] 

0  02  0.4  06  0.8  t.O 

Mol  fraction  water  in  liquid* 

*  In  the  two-phase  region  mol  fraction  values  are  based  on  total  of  both  phases. 
FIG.  4-12.     System,  phenol-water  at  43.4°C. 

oire  of  the  pure  component  at  43.4°C.  is  plotted  vs.  the  mol  fraction 
n  the  liquid.  This  method  of  plotting  is  applied  since  the  vapor 
pressures  of  water  and  phenol  are  so  greatly  different  that  it  is  difficult 
,o  represent  both  of  them  on  the  same  graph. 

The  limit  of  solubility  of  water  in  phenol  corresponds  to  a  mol  frac- 
ion  of  water  of  0.74,  while  the  solubility  of  phenol  in  water  is  0.0225 
nol  fraction  phenol.  For  mol  fractions  of  water  between  0.74  and 
K9775,  the  components  exist  as  two  liquid  layers.  Throughout  this 
*ange  in  which  two  liquid  phases  are  present  and  where  the  eomposi- 
;ions  of  these  phases  are  therefore  constant,  the  partial  pressures  re- 
nain  constant  and  are  represented  by  the  horizontal  line  EF  and  CD. 
The  45°  lines  corresponding  to  Raoult's  law  are  drawn  for  both  com- 
)onents.  The  data  for  the  water  phase  cover  such  a  short  region  that 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  95 

deviations  from  Raoult's  law  for  the  water  component  are  not  obvious. 
In  the  phenol  phase  a  moderate  deviation  from  Raoult's  law  is  appar- 
ent. It  will  be  noted  in  the  figure  that  the  curves  tend  to  approach 
the  immiscible  region  with  rounded  corners  instead  of  with  sharp 
angles. 

The  two  points  of  importance  regarding  the  diagram  are  (1)  con- 
stancy of  partial  pressure  so  Iqng  as  two  liquid  phases  are  present  and 
(2)  the  character  and  extent  of  the  deviation  from  Raoult's  law.  The 
diagram  shows  that  the  partial  pressure  in  phenol,  when  dissolved  in 
water  is  abnormally  high;  i.e.,  it  is  much  greater  than  is  called  for  by 
Raoult's  law,  the  line  AB.  Limited  miscibility  of  two  liquids  implies 
that  the  molecules  of  one  find  it  difficult  to  force  their  way  into  the 
other.  Thus,  it  requires  a  relatively  high  pressure  for  phenol  to  force 
a  small  amount  of  itself  into  water.  This  is  equivalent  to  saying  that, 
when  phenol  has  been  dissolved  in  water,  the  volatility  of  phenol  is 
abnormally  high.  The  less  the  mutual  solubility,  the  more  abnormal 
the  partial  pressure;  hence  the  greater  the  volatility  of  the  dissolved 
component.  The  practical  results  of  these  relationships  are  shown  in 
the  following  example. 

Despite  the  fact  that  phenol  boils  almost  80°  higher  than  water,  in 
certain  regions,  i.e.,  low  concentrations  of  phenol,  the  volatility  of 
phenol  is  greater  than  that  of  water;  i.e.,  the  vapor  given  off  by  such  a 
solution  is  richer  in  phenol  than  the  solution  itself.  If  a  solution  in 
this  low  concentration  region  is  distilled,  the  water  is  discharged  from 
the  bottom  of  the  column  essentially  free  of  phenol,  which  is  found  in 
the  distillate. 

These  data  for  phenol  and  water  are  replotted  in  Fig.  4-13  as  the 
vapor-liquid  equilibrium,  i.e.,  the  mol  fraction  of  the  phenol  in  the 
vapor  as  a  function  of  the  mol  fraction  of  phenol  in  the  liquid.  The 
data  curve  is  labeled  ll experimental. "  These  data  indicate  that  this 
system  forms  a  minimum  boiling  azeotrope  at  a  concentration  of 
about  0.0073  mol  fraction  phenol.  At  concentrations  lower  than  this, 
phenol  is  more  volatile  than  water.  At  all  concentrations  greater  than 
this,  water  is  more  volatile  than  phenol.  This  y,x  curve  indicates  the 
constancy  of  vapor  composition  in  the  two-phase  region.  The  proce- 
dure involving  the  use  of  Raoult's  and  Henry's  laws  employed  for  the 
ethyl  ether-water  system  is  probably  not  suitable  for  the  phenol-water 
system  because  of  the  high  mutual  solubilities.  However,  such  calcu- 
lations were  made  for  illustrative  purposes  and  the  results  are  shown 
on  Figs.  4-13  and  4-14.  The  latter  figure  gives  the  relative  volatilities 
corresponding  to  the  vapor-liquid  curve  of  Fig.  4-13.  It  is  apparent 


96 


FRACTIONAL  DISTILLATION 


\ 


:l 

-if- 


:s 


J 


f 
JC 


r 


! 

t  & 


13* 


\\\ 


V 


"o     3 

s  1 

^.  -^ 


S    S 


\ 


QSDLjd  JodoA  u;  |ouei(d  uoipcjj.  | 


CALCULATION  OP  VAPOR-LIQUID  EQUILIBRIA 


97 


I 

5 

\ 

> 

\ 

// 

"^ 

^ 

1 

y 





jfe 

^ 

^  
/ 

—  "•" 

/ 

/ 

/  f 
/ 

- 

1 

*B                    * 

y  s 

*  r~ 

O  Q              P^ 

Vs. 

'/      / 

t^ 

O 

//      S 

qi 

'o 

-sj 

/ 

/     / 

§ 

•B    t     § 

1 

// 

s 

\ 

?      S      r^ 

S-5   °< 

\ 

^ 

/  / 

,1 

^-° 

/ 

Y  , 

s 

Hi 

S        O         +3 

/ 

V 

/ 

/ 

s 

t 

4  -a  2 

c  |    g 

T" 

\ 

.2   °    ^ 
•ft   H    <y 

/^ 

;* 

l> 

> 
\ 

101 
Xp,Mol  fra< 
s  are  based  01 
4,  Relativ 

J          1 

j* 

^* 

9             S      rH 

T  "~ 

? 

jf> 

1      ^ 

•J-: 

j$ 

^    6 

Q 

r§ 

0          M 

.2   & 

A 

•^ 

a 

1 

^ 

S 

<*« 

^ 

^ 

I- 

d 

^ 

! 

t 

a 

! 

o       c 

98  FRACTIONAL  DISTILLATION 

that  this  simplified  method  is  not  very  satisfactory.  In  fact  it  would 
indicate  that  water  was  more  volatile  than  phenol  at  all  compositions 
and  that  no  azeotrope  was  formed.  The  results  obtained  would  not 
be  of  much  utility  and  would  actually  be  very  misleading.  Thus,  if 
water  containing  a  low  concentration  of  phenol  were  to  be  distilled,  the 
calculations  based  on  Raoult's  and  Henry's  laws  would  indicate  that 
water  was  the  more  volatile  component,  while  the  actual  data  indicate 
that  phenol  is  the  more  volatile.  This  simplified  procedure  gives  only 
very  approximate  results  if  the  mutual  solubilities  are  over  a  few  mol 
per  cent. 

A  better  prediction  of  the  actual  vapor-liquid  equilibria  can  be 
obtained  by  the  use  of  the  Van  Laar  equation.  It  has  been  found  that 
in  general  this  equation  can  be  employed  empirically  to  give  satisfac- 
tory correlation  for  each  of  the  single-phase  regions  of  a  partially 
miscible  system.  Because  of  the  constancy  of  partial  pressure  over  the 
two-phase  region,  the  Van  Laar  constants  for  two  single  phases  should 
be  related  to  each  other  by  the  values  of  the  mutual  solubilities.  How- 
ever, in  most  cases  it  is  found  that  the  constants  obtained  for  the  two 
single-phase  regions  do  not  correlate  with  each  other  and  the  solubility 
limits  in  a  manner  theoretically  required.  This  is  undoubtedly 
because  the  assumptions  made  by  Van  Laar  are  not  satisfied  in  partially 
miscible  systems.  For  example,  it  is  obvious  that  the  entropy  of  mix- 
ing for  such  a  system  would  not  be  equal  to  that  for  the  ideal  case. 

To  apply  the  Van  Laar  equation  independently  to  each  of  the  two 
single-phase  regions  requires  experimental  vapor-liquid  data;  where 
such  data  are  available,  the  equation  is  useful  for  interpolating,  extra- 
polating, and  smoothing  the  results.  However,  if  such  experimental 
data  are  not  available,  the  Van  Laar  equation  can  be  used  with  the 
solubility  limits  to  predict  vapor-liquid  equilibria.  This  method  forces 
the  constants  for  the  two  single-phase  regions  to  be  identical  and,  as 
has  already  been  pointed  out,  the  experimental  data  for  a  number  of 
systems  give  different  constants  for  the  two  regions.  However,  in 
general,  this  method  is  a  better  approximation  of  the  vapor-liquid 
equilibria  than  calculations  based  on  Raoult's  and  Henry's  laws.  The 
only  information  required  is  the  solubility  limits  at  the  temperature 
in  question  and  the  vapor  pressure  of  the  two  pure  components  at  this 
temperature.  Thus  at  constant  temperature, 


CALCULATION  OF  VAPOR-LIQUID  EQUILIBRIA  99 

where  ww  =  water  in  water  phase 

wp  =  water  in  phenol  phase 

pw  =  phenol  in  water  phase 

pp  =  phenol  in  phenol  phase 
and  by  Van  Laar  relationships, 

T  In  yww  =  B 


T  In  ywp  = 
!T  In  ypu  = 


+  A  * 


:) 

R  A  (•*•      /„ 

T  In  ypp  = 


BA(xwp/xpp)2 


( 

V 


l+A 


Using  T  =  316.4°K,  a?wp  =  0.74,  and  xww  «  0.9775  gives  six  equa- 
tions with  six  unknown  quantities  which  can  be  solved  to  give  A  =  0.206 
and  B  =  238.  With  these  constants,  the  vapor-liquid  equilibria  were 
calculated  for  the  system  phenol  and  water  at  a  temperature  of  43.4°, 
and  the  calculated  results  are  given  in  Figs.  4-13  and  4-14,  labeled 
"Van  Laar."  It  will  be  noted  in  this  case  that  the  agreement  with  the 
experimental  results  is  satisfactory  and  would  be  of  great  utility  for 
actual  distillation  calculations.  The  water  phase,  i.e.,  the  low  phenol 
concentration,  shows  a  constant-boiling  mixture  very  close  to  that 
determined  experimentally. 

Where  it  is  necessary  to  estimate  vapor-liquid  equilibria  for  partially 
miscible  systems  for  which  such  data  are  not  available,  it  is  believed 
that  the  Van  Laar  equation  combined  with  the  solubility  limits  is  of  real 
utility.  This  method  of  calculation  becomes  equivalent  to  the  assump- 
tion of  Henry's  and  Raoult's  laws  if  the  mutual  solubility  of  the  two 
components  becomes  extremely  low. 

In  all  cases,  miscible,  partially  miscible,  or  immiscible,  fugacity 
should  be  used  instead  of  partial  pressure  and  vapor  pressure  if  the 
pressures  are  such  that  deviations  from  perfect-gas  law  are  significant. 

If  more  than  two  components  are  involved  and  are  partially  miscible, 
the  relationships  become  so  complicated  that  the  theoretical  method  of 
attack  in  its  present  form  is  not  particularly  helpful.  In  such  cases,  it 


100  FRACTIONAL  DISTILLATION 

is  necessary  to  determine  experimental  values.  However,  even  in  these 
cases  the  general  principles  that  have  been  developed  for  the  binary 
mixture  are  useful  in  attaining  a  picture  of  phenomena  to  be  expected. 

Nomenclature 

A,B  •«  constants  of  Van  Laar  equation 
/  =*  fugacity 

fp  «  fugacity  of  pure  liquid  under  its  own  vapor  pressure 
/*•  a*  fugacity  of  pure  vapor  under  total  pressure,  IT 
H  **  Henry's  law  constant 
P  as  vapor  pressure 
p  «•  partial  pressure 
T  *»  temperature  * 

x  »  mol  fraction  in  liquid 
y  «*  mol  fraction  in  vapor 
7  «s  activity  coefficient 
TT  •»  total  pressure 

References 

1.  CUMMINGS,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1933. 

2.  GILLILAND  and  SCHEELINE,  Ind.  Eng.  Chem.,  32,  48  (1940). 

3.  SIMS,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1933. 


CHAPTER  5 
GENERAL  METHODS  OF  FRACTIONATION 

There  are  several  methods  by  which  fractionation  can  be  obtained. 
The  more  important  among  them  are  (1)  successive  distillation  of  con- 
densed distillate,  (2)  fractional  condensation,  and  (3)  rectification. 

Successive  Distillation.  The  first  method,  successive  distillation  of 
the  condensed  distillate,  can  be  shown  best  by  referring  to  Fig.  2-1. 
Starting  with  a  large  amount  of  liquid  of  the  composition  #5  which  boils 
at  760  mm.  pressure,  temperature  U,  a  small  amount  of  vapor  of  the 
composition  x\  is  removed  from  the  apparatus  and  condensed,  giving 
a  liquid  of  the  composition  x\.  If  this  new  liquid  is  again  distilled, 
the  first  portion  of  the  distillate  will  have  a  composition  #2.  Continu- 
ing this  process,  successive  compositions  of  the  distillates  can  be  esti- 
mated by  following  a  series  of  steps  which  eventually  approach  the 
point  C,  pure  carbon  disulfide,  as  a  limit. 

Removal  of  any  vapor  of  the  composition  x\  from  the  liquid  of  the 
composition  rrs  will  change  the  composition  of  the  liquid  in  the  direction 
of  pure  carbon  tetrachloride.  Therefore,  if  the  distillation  of  the 
liquid  is  continued,  the  composition  will  approach  pure  carbon  tetra- 
chloride as  a  limit,  and  the  last  of  the  liquid  to  be  distilled  would  have 
this  composition. 

It  is  therefore  possible  by  a  systematic  series  of  distillations  to 
separate  any  mixture  of  carbon  disulfide  and  carbon  tetrachloride  into 
practically  pure  carbon  disulfide  and  pure  carbon  tetrachloride.  This 
systematic  fractionation  may  be  shown  diagramatically  as  in  Fig.  5-1, 
in  which  the  original  mixture  (1)  is  divided  into  a  distillate  (3)  and  a 
residue  (2).  (3)  and  (2)  are  then  distilled  separately  and  produce  dis- 
tillates and  residue.  The  distillate  from  (2)  and  the  residue  (3)  being 
combined  into  a  new  liquid  (5)  which  is  again  distilled  with  (4)  and  (6) 
to  continue  the  separation.  This  process  is  continued  until  practically 
complete  separation  is  obtained.  Such  a  process  is  sometimes  carried 
out  in  the  laboratory,  but  it  is  extremely  tedious  and  the  same  results 
usually  can  be  obtained  in  other  more  convenient  ways. 

The  procedure  outlined  in  Fig.  5-1  would  appear  to  result  in  a  num- 
ber of  intermediate  products  and  only  a  small  amount  of  the  desired 

101 


102 


FRACTIONAL  DISTILLATION 


fractions.  Actually  by  a  sufficient  number  of  distillations  the  original 
mixture  can  be  obtained  essentially  as  the  desired  products.  For 
example,  assume  that  fractions  11  and  15  represent  the  desired  separa- 
tion, then  fractions  12,  13,  and  14  can  be  redistilled  as  shown  in  Fig. 
5-2  to  give  fractions  11'  and  15'  which  can  be  obtained  as  the  same 


2 IRES i DUE  |  3  PISTI 

7\  / 

RESIDUE  |  5  |DIST.RE5|  6  [DISTILLA 

8   IDIST  RES  1  0  blSTREs!  tO  [DISTILLA 


FIG.  5-1.     Fractionation  diagram. 


Product 


Product 


FIG.  5-2.     Fractionation  diagram. 

composition  as  fractions  11  and  15.  Samples  20,  21,  and  22  can  be 
redistilled  in  the  same  manner,  and  by  a  repetition  of  the  procedure 
essentially  all  of  the  material  will  be  given  the  desired  separation. 

The  method  outlined  in  the  preceding  paragraph  can  be  made  con- 
tinuous. Thus  by  regulating  the  fractions  vaporized  in  the  various 
distillations  it  is  possible  to  have  fractions  5  and  13  of  the  same  com- 
position as  the  original  mixture,  and  fresh  feed  can  be  added  to  these 
fractions  before  they  are  distilled. 

Fractional  Condensation.  Instead  of  partially  distilling  a  liquid 
into  a  distillate  and  a  residue,  a  vapor  can  be  partly  condensed  into  a 
condensate  and  a  residual  vapor.  The  results  obtained  are  exactly 


GENERAL  METHODS  OF  FRACTION ATION 


103 


analogous  to  those  for  the  successive  distillation  and,  by  a  similar  series 
of  successive  vaporization  and  partial  condensation,  similar  separation 
can  be  effected.  In  fact,  successive  distillation  and  successive  frac- 
tional condensation  can  be  combined  to  increase  the  efficiency  of  the 
operation. 

Multiple  Distillation.  Suppose  an  apparatus  as  in  Fig.  5-3  consist- 
ing of  a  series  of  distilling  kettles  A,  J3,  C,  etc.,  each  kettle  containing 
a  heating  coil  and  necessary  connection  for  vapors  and  liquid.  Sup- 
pose that  kettle  A  contains  a  liquid  mixture  of  carbon  disulfide  and 
carbon  tetrachloride  of  the  composition  x&  as  in  Fig.  2-1;  the  kettle  J5, 


FIG.  5-3.     Diagram  of  multiple  distillation. 

a  liquid  of  composition  x\;  the  kettle  C,  a  composition  of  #2;  and  so  on. 
The  liquid  in  A  boils  at  t^  that  in  B  at  t^  and  that  in  C  at  U.  Since  the 
vapor  leaving  A  is  at  a  temperature  U  which  is  higher  than  the  boiling 
temperature  t%  in  5,  then,  if  the  vapor  from  A  is  led  into  the  heating 
coil  of  J5,  it  will  give  up  its  heat  to  the  contents  of  JS,  boiling  the  liquid 
and  itself  being  partly  condensed.  The  vapor  from  B,  if  led  into  the 
heating  coil  of  C  will  in  the  same  way  boil  the  liquid  in  C,  the  vapor 
being  itself  condensed  as  before.  The  condensed  vapors  in  the  coil 
may  be  drawn  off  into  receiver  D,  E,  and  F,  etc.  However,  since  the 
composition  of  the  liquid  in  B  was  selected  to  be  the  same  as  that  of 
the  vapor  coming  from  the  kettle  A,  from  Fig.  2-1,  the  condensed 
v$,por  in  the  coil  can  be  allowed  to  mix  with  the  contents  of  B  instead 
of  being  withdrawn  into  the  receiver  E.  Now  since  the  vapor  from  A 
is  being  mixed  with  the  liquid  in  B  and  since  there  is  a  heat  inter- 


104 


FRACTIONAL  DISTILLATION 


change  between  the  two,  it  is  much  simpler  to  blow  the  vapor  directly 
into  the  liquid  thus  dispensing  with  the  coil. 

The  vapor  leaving  still  B  will  be  richer  in  carbon  disulfide  than  the 
vapor  from  still  A,  and  the  liquid  in  B  will  therefore  tend  to  become 
poor  in  carbon  disulfide.  The  concentration  of  the  liquid  in  B  can 
be  maintained  constant  by  adding  liquid  from  C  which  is  rich  in  carbon 


-i- 


FIG.  5-4.     Schematic  diagram  of  rectifying  column. 

disulfide,  and  removing  liquid  from  B  and  adding  it  to  A.  By  a  similar 
procedure  the  operation  of  the  other  stills  can  be  maintained  at  a 
steady  state.  A  little  consideration  of  Fig.  2-1  will  show  that  to  make 
the  system  operate  at  the  compositions  indicated  would  require  that 
all  the  vapor  from  C  be  condensed  and  returned  through  line  L  to  C  and 
that  no  liquid  be  withdrawn. 

Rectification.  An  apparatus  in  which  this  direct  interchange  of 
heat,  condensation,  and  evaporation  can  take  place  is  called  a  rec- 
tifying tower,  and  the  process  carried  on  within  it  is  called  rectification. 

Such  a  system  is  shown  in  Fig.  5-4,  where  8  is  the  still  body,  or  ket- 
tle, Resting  on  the  outlet  of  the  still  is  a  column  divided  into  com- 


GENERAL  METHODS  OF  FRACTION ATION  105 

partments  by  plates  perforated  with  small  holes.  Each  plate  has  an 
overflow  pipe  discharging  into  a  pool  of  liquid  on  the  plate  below. 
The  layer  of  liquid  on  each  plate  is  prevented  from  passing  down 
through  the  holes  by  vapor  which  is  rising  up  through  these  holes  from 
the  compartment  next  below.  Any  excess  liquid  accumulated  on  the 
plate  flows  down  through  the  overflow  pipe.  The  letters  on  the 
apparatus  correspond  to  those  of  Fig  5-3.  The  vapor  from  the  still  at 
temperature  i*  and  composition  x\  passes  up  and  exchanges  heat  and 
molecules  with  the  liquid  in  compartment  B.  A  binary  vapor  of  the 
composition  x%  is  produced  which  bubbles  up  through  the  liquid  on  the 
next  higher  plate  which  is  richer  in  carbon  disulfide  with  the  composi- 
tion xz.  Here  again  exchange  between  the  vapor  and  liquid  takes 
place,  and  a  vapor  of  composition  x&  even  richer  in  the  carbon  disulfide 
is  produced.  This  can  be  repeated  any  number  of  times,  and  the  vapor 
finally  issuing  from  the  apparatus  at  the  top  and  into  the  condenser  is 
practically  pure  carbon  disulfide.  As  in  the  previous  case,  each  one  of 
the  compartments  in  the  column  may  be  considered  a  small  still,  in 
which  the  source  of  heat  is  the  hot  vapor  coming  from  below  and  the 
cooling  element  is  the  cooler  liquid  from  the  plate  above. 

The  quantitative  relationships  given  are  valid  only  in  case  the 
molal  ratio  of  liquid  overflowing  from  plate  to  plate  to  the  vapor  flow- 
ing through  the  plate  is  practically  unity;  i.e.,  the  ratio  of  distillates  to 
liquid  vaporized  is  exceedingly  small.  In  practice,  less  overflow  must 
be  employed  to  reduce  the  heat  consumption,  and  the  rate  of  enrich- 
ment is  less  rapid  than  that  indicated  in  the  explanation. 

The  analogy  between  this  fractionating  column  and  the  series  of 
kettles  would  be  better  if  the  vapor  leaving  the  liquid  on  the  plate 
had  the  equilibrium  composition  assumed.  But,  unfortunately,  no 
design  has  been  able  to  prevent  some  of  the  vapor  from  the  plate 
below  from  passing  through  the  liquid  on  the  plate  without  coming  into 
equilibrium  with  it.  The  vapor  above  any  plate,  therefore,  will  con- 
tain less  of  the  volatile  component  than  would  be  the  case  if  complete 
equilibrium  were  reached.  This  ideal  case  is  discussed  here  because  it 
brings  out  clearly  the  nature  of  the  underlying  phenomena,  The  more 
practical  cases  will  be  considered  in  Chap.  7  on  the  Rectification  of 
Binary  Mixtures. 

The  interchange  between  the  vapor  bubbles  and  the  liquid  o"n  the 
plate  is  a  result  of  the  fact  that  the  two  are  not  at  equilibrium.  Thus 
in  the  ideal  case  considered  a  vapor  of  composition  x\  was  bubbling 
through  a  liquid  of  the  same  composition.  The  vapor  in  equilibrium 
with  a  liquid  of  composition  x\  would  have  been  #2.  Thus,  as  the  sys- 


106  FRACTIONAL  DISTILLATION 

tern  tends  to  approach  equilibrium,  carbon  tetrachloride  molecules 
pass  from  the  vapor  to  the  liquid  and  carbon  disulfide  molecules  will 
pass  from  the  liquid  to  the  vapor.  The  number  of  molecules  passing 
in  the  two  directions  will  be  essentially  equal  since  in  most  cases  the 
energy  released  when  one  molecule  of  carbon  tetrachloride  goes  into 
the  liquid  phase  will  be  about  equal  to  that  required  for  vaporizing  one 
molecule  of  the  carbon  disulfide.  Thus  the  total  number  of  molecules 
in  the  vapor  tends  to  remain  about  constant.  This  interchange 
between  the  vapor  and  the  liquid  is  governed  by  the  usual  mass-trans- 
fer mechanism,  and  the  rate  of  exchange  increases  with  the  amount 
of  interfacial  area  and  the  turbulence  involved.  A  close  approach  to 
equilibrium  is  desired,  and  the  equipment  is  designed  to  give  intimate 
contact  between  the  two  phases.  Besides  the  bubbling  action  already 
described,  the  process  produces  a  considerable  amount  of  spray,  and 
there  is  also  an  interchange  between  the  vapor  and  the  liquid  droplets 
above  the  main  body  of  liquid  that  is  helpful  in  obtaining  a  closer 
approach  to  equilibrium. 

This  process  of  countercurrent  contact  of  a  vapor  with  a  liquid 
whicK  has  been  produced  bv  partial  condensation  of  the  vapor  is 
termed  rectification.  Its  result  is  equivalent  to  a  series  of  redistilla- 
tions with  the  consumption  of  no  additional  heat  and  is  analogous  in 
this" respect  to  multieffect  evaporation.  However,  it  is  only  the  result 
that  is  similar  and  not  the  mechanism  of  obtaining  it. 


CHAPTER  6 
-SIMPLE  DISTILLATION  AND  CONDENSATION 

Simple  Distillation.  Distillation  without  rectification  can  be 
carried  out  by  several  methods.  The  two  most  generally  considered 
cases  are  (1)  continuous  simple  distillation  and  (2)  differential  distilla- 
tion. In  continuous  distillation,  a  portion  of  the  liquid  is  vaporized 
under  conditions  such  that  all  the  vapor  produced  is  in  equilibrium 
with  the  unvaporized  liquid.  In  differential  vaporization,  the  liquid 
is  vaporized  progressively,  and  each  increment  of  vapor  is  removed 
from  contact  with  the  liquid  as  it  is  formed  and,  although  each  incre- 
ment of  vapor  can  be  in  equilibrium  with  the  liquid  as  it  is  formed,  the 
average  composition  of  all  of  the  vapor  produced  will  not  be  in  equi- 
librium with  the  remaining  liquid. 

Continuous  Simple  Distillation.  Distillations  that  approximate 
this  type  are  usually  carried  out  on  a  continuous  basis  such  that  the 
liquid  feed  is  added  continuously  to  a  well-mixed  still  in  which  a  definite 
fraction  is  vaporized  and  removed  and  the  excess  unvaporized  liquid  is 
withdrawn  from  the  still.  An  alternate  arrangement  is  to  preheat  the 
feed  and  add  it  to  a  flash  or  disengaging  section  where  vapor  and  liquid 
are  separated  and  removed  without  additional  heat  requirements.  In 
either  case,  assuming  that  the  vapor  and  liquid  leaving  are  in  equi- 
librium with  each  other,  the  two  fractions  are  related  to  each  other  by 
equilibrium  constants  and  material  balances.  Thus  for  each  com- 
ponent in  a  mixture  the  following  material  balance  can  be  written  : 

Vy  +  Lx  =  Fz  (6-1) 

where  F,  L,  F  =  mols  of  vapor,  unvaporized  liquid,  and  feed,  respec- 

tively 
y,  x,  z  =  mol    fractions    of    components    in    corresponding 

streams 
By  over-all  material  balance  V  +  L  =  F  and  the  fraction  vaporized, 


Thus  if  the  fraction  vaporized,  the  feed  composition,  and  the  relation 

107 


108  FRACTIONAL  DISTILLATION 

between  y  and  x  are  known,  both  the  vapor  and  the  liquid  compositions 
can  be  calculated.  In  order  to  determine  the  equilibrium  relation- 
ships either  the  temperature  or  the  total  pressure  must  be  known. 
Given  either  of  these,  the  other  can  be  determined  from  the  equilibrium 
relationship  for  all  the  components  involved. 

Differential  Distillation.  This  type  of  distillation  is  usually  carried 
out  as  a  batch  operation  although  continuous  units  may  also- operate 
in  this  manner.  Considering  first  a  batch  distillation,  if  a  mixture  of 
liquid  is  distilled,  the  distillate  contains  a  greater  portion  of  the  more 
volatile  material  than  the  residue,  and  as  distillation  proceeds  both  the 
distillate  and  the  residue  become  poorer  in  the  more  volatile  compo- 
nents. This  change  in  composition  may  be  estimated  quantitatively  if 
the  relation  of  the  composition  of  vapor  to  that  of  the  liquid  is  known. 
Consider  W  parts  of  original  mixture  containing  x0  fraction  of  com- 
ponent A.  Allow  a  differential  amount  —  dW  to  be  vaporized  of  a 
composition,  y,  under  such  conditions  that  the  vapor  is  continually 
removed  from  the  system. 

By  material  balance, 

-y  dW  =  ~-d(Wx) 

=  -W  dx  -  xdW 
W  dx 

—     ni     .,_-.      ff> 

~y 


fwdW=    /•'•    dx 

Jw.   W       Jx.y-; 


In    W  I"       dX  to  tt 

ln  W.  =  A,  ?=-*  (6"3) 

This  equation  was  developed  by  Rayleigh  (Ref.  2)  and  is  often 
termed  the  Rayleigh  equation.  It  can  be  used  with  W  as  weight  and 
x  as  weight  fraction,  or  with  W  as  mols  and  x  as  mol  fraction.  It  is 
usually  applied  on  the  basis  that,  at  any  given  instant,  y  is  in  equi- 
librium with  x,  but  the  derivation  does  not  require  this  condition.  A 
similar  equation  applies  to  each  component  in  a  mixture. 

The  use  of  Eq.  (6-3)  requires  the  relationship  between  y  and  x  and, 
even  if  they  are  assumed  to  be  in  equilibrium  with  each  other  as  the 
vapor  is  formed,  it  is  usually  difficult  to  express  the  equilibria  mathe- 
matically for  a  general  integration.  The  integration  can  be  performed 
graphically  if  the  relationship  between  y  and  x  is  available. 

Batch  Distillation  Example.  As  an  illustration  of  the  use  of  this  equation,  con- 
sider the  experiment  performed  by  Rayleigh.  1,010  g.  of  a  7.57  mol  per  cent  solu- 
tion of  acetic  acid  in  water  was  distilled  until  the  still  contained  254  g.  whose  com- 


SIMPLE  DISTILLATION  AND  CONDENSATION 


109 


position  was  11  mol  per  cent  acetic  acid.  Assuming  that  the  vapor  leaves  in 
equilibrium  with  the  liquid,  calculate  the  final  composition  to  be  expected  on  the 
basis  of  Eq.  (6-3).  Equilibrium  data  for  the  system  acetic  acid-water  are  given  in 
Table  6-1. 

There  are  several  approximate  methods  that  may  be  followed  in  integrating  Eq. 
(6-3).  First,  for  small  temperature  and  composition  ranges,  the  relation  between 
the  vapor  and  liquid  may  be  approximately  represented  by  a  straight  line,  or 
y  =  cXj  where  c  is  a  constant. 


w_  _  f*    dx    _  /•* 

JFo       Jx0  co;  -  a?       /*, 


dx 


1 


TABLE  6-1 


Mol  fraction 
of  acetic  acid 
in  liquid,  x 

Mol  fraction 
of  acetic  acid 
in  vapor,  y 

c-« 

X 

Relative 
volatility 
of  acetic  acid 
to  water,  a 

0  0677 

0  0510 

0.75 

0  74 

0  1458 

0.1136 

0  2682 

0.2035 

0.76 

0.70 

0  3746 

0.2810 

0.4998 

0.3849 

0.77 

0.625 

0.6156 

0.4907 

0  7227 

0  6045 

0.84 

0.59 

0.8166 

0.7306 

0  9070 

0.8622 

0.95 

0.64 

Clearing  of  logarithms, 

W        (x  V/'-i  ^        /TFV"1 

™r  -  (-)  or       :r*(ir) 

rr  o  \^o/  Xo  \  rr  Q/ 

From  the  table  a  value  of  c  •»  0.75  is  appropriate  and 

254 


(6-4) 


)0  76-1 


0.0757 
x  =  0.107  (10.7  mol  per  cent) 

The  calculated  value  is  in  good  agreement  with  the  experimental  result  and 
indicates  that  the  various  assumptions  made  are  reasonably  satisfied. 

Another  method  is  by  the  use  of  the  relative  volatility  which  is  denned  by 
Eq.  (3-5). 

VA/XA  ^  a 
VB/XB       aAB 

where  <XAB  is  the  relative  volatility  of  component  A  to  component  B.  For  a  large 
number  of  mixtures,  the  variation  of  a.  with  composition  is  small,  and  an  average 
value  may  be  employed.  For  a  binary  mixture,  the  expression  can  be  rewritten 


CfXA 

1  +  (a  -  l)xA 


110  FRACTIONAL  DISTILLATION 

The  use  of  this  relation  with  Eq.  (6-3)  gives 

dx 


.     W         f* da 

In  Tjnr  =*     I     

W0         I  ctx 

JXo    1  +  («  - 


a  -  1       x9(l  —  x)  I  —  x 


Using  a  »  0.74  (see  Table  6-1)  gives  x  -  0.106. 

In  other  cases,  it  will  be  found  that  the  variations  in  c  and  a  are  so  great  that 
these  methods  are  not  satisfactory.  In  such  cases  they  can  be  applied  successively 
over  small  concentration  ranges  but  graphical  integration  is  usually  preferable. 

There  is  an  alternate  form  of  the  differential  distillation  equation 
that  is  frequently  more  convenient  to  use  than  Eq.  (6-3).  Consider  a 
mixture  containing  A  mols  of  one  component  and  B  mols  of  some  other. 
Let  a  differential  quantity  of  vapor  be  produced  containing  —  dA  and 
—  dB  mols  of  the  two  components,  respectively,  then  assuming  vapor- 
liquid  equilibria, 

dA          A  ,c  ,jv 

-=dB  =  aB  (6'6) 

A  similar  equation  can  be  written  between  any  two  components  of  the 
mixture.  If  the  relative  volatility  is  constant,  the  equation  can  be 
integrated  directly, 


M 
)  A. 


B 


where  A0,  B0  =*  mols  of  the  two  components  in  still  at  some  base  time 
A,  J5  =  mols  in  still  at  some  later  time 

i    A  i     B 

In  --  =  a  In  -- 


A  (0 


A 

If  the  vapor  formed  is  not  in  equilibrium  with  the  liquid,  the  value  of 
a  in  Eq.  (6-6)  will  have  to  be  modified  to  express  the  true  relationship. 

Batch  Dehydration  of  Benzene.  As  another  example  of  the  use  of  Rayleigh's 
equation,  consider  the  dehydration  of  benzene.  Benzene  saturated  with  water  at 
20°C.  contains  0.25  mol  per  cent  water,  and  it  is  to  be  given  a  simple  differential 
distillation  at  a  constant  pressure  of  1  atm.  The  operation  is  to  proceed  until  the 
mol  per  cent  water  in  the  liquid  remaining  in  the  still  is  0.00025.  The  following 
data  and  simplifying  assumptions  will  be  used  in  the  calculations: 

1.  Over  the  temperature  range  involved  it  is  assumed  that  the  vapor  pressure 
of  pure  benzene  is  2.1  times  the  vapor  pressure  of  water.  * 


SIMPLE  DISTILLATION  AND  CONDENSATION  111 

2.  At  the  distillation  temperature,  benzene  saturated  with  water  contains 
1.5  mol  per  cent  water,  and  water  saturated  with  benzene  contains  0.039  mol 
per  cent  benzene. 

3.  It  is  assumed  that  the  vapor  leaves  in  equilibrium  with  the  liquid. 

4.  No  condensate  will  be  returned  to  the  still. 

5.  For  the  two  single-phase  regions  of  benzene  containing  water  and  water 
containing  benzene,  it  is  assumed  that  Raoult's  law  applies  to  the  component  in 
large  amount,  and  Henry's  law  to  the  component  in  small  amount. 

Solution.  By  a  procedure  similar  to  that  employed  for  ether  and  water  in 
Chap.  4,  it  is  possible  to  calculate  the  relative  volatility.  The  solubility  of  water 
in  benzene  is  higher  at  the  distillation  temperature  than  at  20°C.,  and  only  a 
benzene  phase  will  be  present  in  the  still.  The  partial  pressure  of  benzene  will 
follow  Raoult's  law. 

PB  -  XaPn 

For  water  the  partial  pressure  will  follow  Henry's  law  with  a  constant  such  that  a 
mol  fraction  of  0.015  will  give  a  partial  pressure  equal  to  that  over  water  saturated 
with  benzene.  Thus, 


« 
Pw  0.015 

and  the  relative  volatility  of  water  to  benzene  is 


0.015P* 
=  31.7 

Even  though  the  vapor  pressure  of  benzene  is  over  twice  that  of  water  at  the 
distillation  temperature,  the  volatility  of  water  in  benzene  is  over  30  times  that  of 
the  benzene  because  of  the  abnormalities  indicated  by  the  low  mutual  solubilities. 

The  value  of  the  relative  volatility  can  be  used  with  Eq.  (6-5)  and  gives 

W  _         1  0.0000025(0.9975)       ,       0.9975 

Wo  ~  31.7  -  1  ln  0.0025(0.9999975)  "*"      0.9999975 

w.  -  °-796 

Therefore,  20.4  per  cent  of  the  charge  should  be  vaporized. 

The  example  given  assumed  that  no  condensate  would  be  returned  to  the  still. 
Actually  the  condensate  will  break  in  two  layers,  a  water  layer  and  a  benzene 
layer,  and  the  benzene  layer  saturated  with  water  could  be  returned  continually 
to  the  still  for  redistillation  or  it  could  be  stored  and  added  to  a  subsequent  cycle. 
The  amount  of  heat  required  for  the  drying  operation  could  be  reduced  by  rectifica- 
tion or  partial  condensation.  These  operations  will  be  considered  in  a  later 
chapter. 

Steam  Distillation.  A  common  example  of  simple  distillation  is  the 
so-called  steam  distillation.  The  term  is  generally  applied  to  distilla- 
tions carried  out  by  the  introduction  of  steam  directly  into  the  liquid 


112  FRACTIONAL  DISTILLATION 

in  the  still  and  is  usually  limited  to  those  cases  in  which  the  solubility 
of  the  steam  in  the  liquid  is  low  at  the  temperature  and  pressure  in 
question.  It  is  usually  applied  to  relatively  high-boiling  organic 
materials  which  would  decompose  if  they  were  distilled  directly  at 
atmospheric  pressure  or  to  liquids  that  have  such  poor  heat-transfer 
characteristics  that  excessive  local  superheating  would  result  with 
indirect  heating.  By  steam  distillation  a  volatile  organic  material 
may  be  separated  from  nonvolatile  impurities,  or  mixtures  may  be 
separated  with  results  about  equivalent  to  those  predicted  by  the  Ray- 
leigh  equation. 

The  molal  ratio  of  the  organic  material  to  the  steam  is  the  ratio  of 
the  mol  fractions  in  the  vapor,  and,  assuming  that  the  gas  laws  apply, 

22  =  Hi  =  P2  =  _££_ 
nw       yw       Pw      TT  —  po 

where  n  =  mols 

y  =  mol  fraction  in  vapor 

p  =  partial  pressure 

TT  =  total  pressure 

Subscripts  0  and  W  refer  to  organic  and  water,  respectively. 
If  the  organic  material  is  immiscible  with  water,  and  equilibrium  is 
attained,  po  would  be  the  vapor  pressure  of  the  organic  material,  and 
the  partial  pressure  of  water  would  be  the  total  pressure  minus  p0.  If  a 
liquid  water  phase  were  present,  the  partial  pressure  of  water  would 
haVfc  to  be  the  vapor  pressure  of  water,  thus  fixing  the  distillation 
temperature  at  a  given  total  pressure.  For  this  case,  the  distillation 
temperature  will  always  be  less  than  that  corresponding  to  the  boiling 
point  of  water  at  the  total  pressure  in  question.  This  illustrates  one 
of  the  real  advantages  of  steam  distillation.  Thus  high-boiling 
organic  material  can  be  steam-distilled  in  an  atmospheric  pressure 
operation  at  a  temperature  below  100°C.  If  a  liquid  water  phase 
is  not  present,  then  both  the  total  pressure  and  the  temperature  can  be 
arbitrarily  chosen,  and  the  partial  pressure  of  water  is  the  difference 
between  the  total  pressure  and  the  vapor  pressure  of  the  organic 
material.  Of  course,  this  partial  pressure  of  water  must  not  be  greater 
than  the  vapor  pressure  of  pure  water,  or  a  liquid  phase  will  form. 

If  water  and  the  material  being  distilled  are  not  immiscible,  the 

vapor-liquid  equilibria  for  the  system  in  question  will  have  to  be  known 

in  order  to  determine  the  relation  between  the  vapor  and  the  liquid 

composition. 

Steam  distillation  can  be  carried  out  in  several  manners.     It  is  possi- 


SIMPLE  DISTILLATION  AND  CONDENSATION  113 

ble  to  pass  steam  directly  through  the  liquid  without  any  other  source 
of  heat.  Because  heat  must  be  supplied  for  the  vaporization  of  the 
organic  material,  stearn  will  condense  and  form  a  liquid  phase  unless  it 
is  very  highly  superheated.  If  water  does  condense,  the  distillation 
temperature  will  be  less  than  the  boiling  point  of  pure  water,  and  at 
such  a  temperature  the  value  of  po  will  frequently  be  very  small  and 
Eq.  (6-8)  would  require  a  large  number  of  mols  of  steam  per  mol  of 
organic  material  distilled.  Actually  the  molal  ratio  given  by  this 
equation  is  simply  the  overhead  vapor  ratio,  and  any  condensation  of 
steam  an  the  still  would  be  in  addition  to  the  values  so  obtained. 
Owing  to  the  low  molecular  weight  of  steam  relative  to  that  of  the 
high-boiling  organic  material  the  consumption  of  steam  may  not  be 
excessive.  However,  in  some  cases  it  may  be  desirable  to  reduce  the 
steam  consumption.  This  can  be  accomplished  by  indirectly  heating 
the  still  and  maintaining  the  distillation  temperature  higher  than  that 
obtained  when  a  liquid  water  phase  is  present.  For  maximum  steam 
economy,  the  temperature  should  be  as  high  as  is  possible  without 
undesirable  thermal  effects.  This  higher  temperature  increases  the 
value  of  po  and  reduces  that  of  IT  —  p0  thus  decreasing  the  ratio  of  the 
steam  required  per  unit  of  organic  material  distilled.  Another 
method  of  reducing  the  steam  consumption  is  to  reduce  the  total 
pressure.  Thus,  if  the  total  pressure  was  reduced  to  the  value  of  p0, 
no  steam  would  be  required  and  the  organic  material  would  boil 
directly.  Frequently  this  is  not  feasible  because  of  the  very  high 
vacuum  required  or  the  undesirable  heat-transfer  characteristics  oHhe 
liquid.  However,  by  the  use  of  reduced  pressure  the  amount  of 
steam  required  can  be  made  relatively  small,  and  if  the  vacuum  is 
adjusted  such  that  the  vapor  mixture  will  condense  with  the  cooling 
water  available,  the  load  on  the  vacuum  pump  will  be  low. 

Steam  Distillation  Example.  As  an  example  of  steam  distillation,  consider  the 
separation  of  a  mixture  of  two  high-boiling  organic  acids  from  a  small  amount  of 
nonvolatile  carbonaceous  material.  The  steam  distillation  is  carried  out  at 
100°C.  under  a  total  pressure  of  150  mm.  Hg.  The  organic  acid  mixture  contains 
70  and  30  mol  per  cent  of  the  low-  and  high-boiling  acids,  respectively,  and  at 
100°C.  the  vapor  pressures  of  the  two  acids  are  20  and  8  mm.  Hg.  It  is  assumed 
that  the  mixture  of  the  two  acids  obeys  Raoult's  law  and  that  they  are  immiscible 
with  water.  The  nonvolatile  carbonaceous  material  is  assumed  to  have  no  effect 
on  the  vapor-liquid  equilibria.  It  will  be  assumed  that  the  vapor  leaves  in  equilib- 
rium with  the  liquid  in  the  still,  and  two  cases  will  be  considered.  In  the  first,  the 
mixture  of  acids  will  be  fed  continuously  to  a  still  of  small  capacity,  and  it  will  be 
assumed  that  steady-state  conditions  have  been  reached  in  which  the  composition 
of  the  organic  acids  in  the  condensate  is  the  same  as  in  the  feed.  In  the  second 


114 


FRACTIONAL  DISTILLATION 


case,  a  batch  distillation  of  the  differential  type  will  be  carried  out.  It  is  assumed 
that  all  the  sensible  and  latent  heat  is  supplied  either  externally  or  by  superheat  in 
the  steam.  The  calculations  are  to  determine  how  many  pounds  of  steam  must 
be  used  per  mol  of  acid  recovered  in  each  case. 

Solution  of  Part  1.  Continuous  Operation.  In  this  case  the  ratio  of  the  two 
organic  acids  in  the  vapor  is  the  same  as  in  the  feed  and,  since  Raoult's  law  has  been 
assumed  for  the  organic  acids  mixture,  it  is  possible  to  calculate  the  ratio  of  the 
two  acids  in  the  still.  Thus, 

Hi  «  P&\  »  a  5! 

2/2          PzXz  Xz 

Al 


where  x  «  mol  fraction  based  on  the  two  acids  only 
At,  Al  —  original  mols  of  the  more  and  less  volatile  acids,  respectively 
With  xt  «  1  -  xij 


Pi   4-  P2   =»  PlXi   4- 


-  l)a?i  4- 


Pi(A?  4-  AQ 


where  PI,  p*  =«  partial  pressures  of  acids 

Pi,  PS  «•  vapor  pressures  of  acids 
and 

pHiO  »  IT  -  (pi  4-  p8) 

The  pounds  of  water  required  per  pound  mol  of  acid  is 


18(ir  -  pi  - 
Pi  *f  Pa 


18 


18    ^ ^^-  -  1 


+r;) 


-  178 

Solution  of  Part  2.  In  this  case,  it  is  assumed  that  the  still  is  charged  with  the 
mixture  of  acids  and  that  the  distillation  is  continued  until  all  the  acids  have  been 
vaporized. 

Considering  the  acids  only,  the  differential  distillation  Eq.  (6-6)  gives 


where  Ai  «•  mols  of  more  volatile  acid  in  still 
At  •»  mols  of  less  volatile  acid  in  still 


SIMPLE  DISTILLATION  AND  CONDENSATION 

and  for  constant  a 

4»  dL4« 


115 


Al 


By  steam  distillation  Eq.  (6-8), 
-dN  -dN 


dA\  -f-  dAz       /    A i    .   . 
where  N  =  mols  of  steam 


^  TT  -  Pl  -  P2  _ 
pi  +p2 


The  pounds  of  steam  per  pound  mol  of  acid  is 


1SN 


18 


-  1 


This  is  identical  to  the  relation  found  in  Part  1,  and  the  steam  requirement  is  the 
same.  Less  steam  per  mol  of  acids  distilled  would  be  required  in  the  first  part  of 
the  batch  distillation,  but  more  would  be  required  in  the  last  portion.  The  above 
analysis  indicates  that  the  two  differences  would  just  balance  out. 

Partial  Condensation.  It  is  frequently  desirable  to  partially  condense 
a  vapor.  Such  an  operation  can  be  used  to  produce  a  separation  of 
the  components  but,  in  general,  it  is  employed  for  obtaining  a  portion 
as  condensate  for  some  specific  purpose. 

The  partial  condensation  of  a  vapor  mixture  can  produce  a  wide 
variation  in  degree  of  separation  obtained.  If  the  condensation  is 
carried  out  rapidly,  the  time  for  interchange  between  the  condensate 
and  the  vapor  may  be  so  short  that  essentially  no  selective  mass  trans- 
fer of  the  components  occurs.  In  this  case,  the  composition  of  the 
condensate  will  be  the  same  as  that  of  the  vapor.  If  the  condensation 


116  FRACTIONAL  DISTILLATION 

is  carried  out  at  a  slow  rate,  mass-transfer  interchange  will  occur  and 
several  different  degrees  of  separation  can  be  obtained.  The  operation 
can  be  carried  out  such  that  the  condensate  is  essentially  in  equilibrium 
with  the  uncondensed  vapor,  and  this  type  of  operation  will  be  termed 
equilibrium  partial  condensation.  Alternatively,  the  condensation  can 
be  carried  out  such  that  the  condensate  as  it  is  formed  is  in  equilibrium 
with  the  vapor,  but  the  condensate  is  removed  continuously  and  thus 
the  total  condensate  would  not  be  in  equilibrium  with  the  uncon- 
densed vapor.  This  type  of  operation  will  be  termed  'differential  par- 
tial condensation.  In  a  third  distinct  type  of  partial  condensation,  the 
vapor  passes  through  the  condenser  unit  countercurrent  to  the  con- 
densate. Assuming  that  efficient  countercurrent  contact  is  obtained, 
it  should  be  possible  to  produce  a  higher  degree  of  separation  than  is 
possible  with  equilibrium  partial  condensation. 

Equilibrium  partial  condensation  is  handled  mathematically  in  a 
manner  completely  analogous  to  that  for  equilibrium  distillation.  It 
is  a  type  of  condition  that  is  frequently  encountered  in  the  partial 
condenser  of  a  rectification  unit  in  which  the  uncondensed  vapor  and 
the  condensate  flow  along  together  and  reach  a  close  approach  to 
equilibrium. 

Differential  partial  condensation  is  not  a  common  type  of  operation, 
but  the  countercurrent  version  of  partial  condensation  is  probably 
approached  in  condensers  in  which  the  vapor  flows  upward  condensing 
on  the  tube  walls  and  the  condensate  flows  down  along  the  walls  in 
contact  with  the  upward  flowing  vapor.  Such  a  wetted-wall  unit  will 
give  mass  transfer  between  the  vapor  and  the  liquid,  but  in  general  the 
sizes  of  the  tubes  desirable  for  heat  transfer  and  fluid  flow  are  such  that 
the  interchange  between  vapor  and  liquid  is  relatively  poor,  and  the 
separation  obtained  in  practice  is  probably  not  much  greater  than  that 
equivalent  to  equilibrium  partial  condensation.  While  a  condenser 
could  be  designed  such  that  it  would  give  more  efficient  countercurrent 
action,  it  has  been  found  to  be  more  economical  to  design  a  condenser 
for  the  liquefaction  function  and  to  obtain  the  desired  separation  of  the 
components  by  more  effective  means. 

Analysis  of  Partial  Condenser  Data.  Gunness  (Ref.  1)  reports  experimental 
data  obtained  on  the  partial  condenser  of  a  rectifying  column  stabilizing  absorption 
naphtha.  The  vapor  from  the  column  was  passed  downward  through  the  partial 
condenser,  and  the  uncondensed  vapor  and  condensate  from  the  bottom  of  the 
condenser  passed  together  to  the  reflux  drum  where  they  were  separated.  Owing 
to  the  concurrent  flow  of  the  vapor  and  the  liquid,  it  would  be  expected  that  this 
system  might  approximate  equilibrium  partial  condensation.  The  data  for  a  test 


SIMPLE  DISTILLATION  AND  CONDENSATION 


117 


taken  when  the  pressure  was  254  p.s.i.a.  and  the  temperature  in  the  reflux  drum 
was  117°F.  are  given  in  the  first  three  columns  of  Table  6-2.  The  fourth  column 
gives  values  of  the  equilibrium  constant  obtained  from  Table  3-2,  page  41,  at  the 
temperature  and  pressure  corresponding  to  the  reflux  drum.  If  the  uncondensed 
vapor  and  the  liquid  were  in  equilibrium  with  each  other,  their  composition  should 
be  related  by  the  equilibrium  constants.  As  a  method  of  making  this  comparison, 
the  last  column  in  the  table  gives  the  values  of  the  vapor  composition  divided  by 
the  equilibrium  constant  for  each  of  the  components.  If  the  vapor  and  liquid  were 
at  equilibrium,  the  compositions  so  calculated  should  be  the  same  as  those  given 
for  the  liquid  reflux.  It  will  be  noted  that  the  experimental  composition  and  the 
calculated  values  are  in  good  agreement.  In  fact  it  is  probable  that  the  agree- 
ment i&  within  the  accuracy  of  the  experimental  data  and  the  equilibrium  con- 
stants. The  close  agreement  indicates  the  reliability  of  the  experimental  data 

and  the  applicability  of  the  vapor-liquid  equilibrium  constant  to  this  system. 
v 

TABLE  6-2 


V  V  i  t 
Coirfp^nt 

Residue  gas,  y 

Liquid  reflux,  x 

K 

**  -  1 

CH4 

0.053 

0.007 

12 

0.0044 

C2H4 

0.011 

0.002 

3.6 

0.003 

C2H6 

0.146 

0.0618 

2.55 

0.057 

C8H6 

0.140 

0.12 

1.05 

0.133 

C8H8 

0.537 

0.580 

0.94 

0.572 

t-C4 

0.081 

0.160 

0.50 

0.162 

n-C4 

0.032 

0.069 

0.40 

0.080 

Nomenclature 


mols  of  A  in  still 
mols  of  B  in  still 
c  =  constant 
F  =  mols  of  feed 

«  mols  of  unvaporized  liquid 
=  mols  of  steam 
••  mols 

•  partial  pressure 

•  vapor  pressure 
«  mols  of  vapor 

W  ~  mols  of  liquid  in  still 
x  =  mol  fraction  in  liquid 
y  »  mol  fraction  in  vapor 
z  «  mol  fraction  in  feed 
a  •»  relative  volatility 
TT  «  total  pressure 


A 
B 


L 

N  . 
n 

P 
P 
V 


References 


1.  GUNNESS,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1936. 

2.  RAYLEIGH,  Phil  Mag.,  6th  series,  4,  521  (1902). 


CHAPTER  7 
RECTIFICATION  OF  BINARY  MIXTURES 

The  separation  of  two  liquids  from  each  other  by  fractional  distilla- 
tion may  be  accomplished  in  two  general  ways:  (1)  the  batch,  or  inter- 
mittent, method  and  (2)  the  continuous  method.  In  the  former,  the 
composition  and  temperature  at  any  point  in  the  system  are  changing 
continually;  in  the  latter,  conditions  at  any  point  are  constant. 

It  will  be  recalled  that  a  fractionating  column  consists  of  a  system 
up  through  which  vapors  are  passing  and  down  through  which  a  liquid 
is  running,  countercurrent  to  the  vapor,  the  liquid  and  vapor  being  in 
more  or  less  intimate  contact  with  each  other.  Furthermore,  the 
vapor  and  liquid  tend  to  be  in  equilibrium  with  each  other  at  any  point 
in  the  column,  the  liquid  and  vapor  at  the  bottom  of  the  column  being 
richer  in  the  less  volatile  component  than  at  the  top.  It  is  evident, 
therefore,  that  the  action  of  such  a  column  is  similar  to  that  of  a 
scrubbing  or  washing  column,  where  a  vapor  is  removed  from  a  gas 
that  is  passing  up  through  the  column,  by  bringing  into  contact  with  it, 
countercurrent,  a  liquid  in  which  the  vapor  is  soluble,  and  that  will 
remove  it  from  the  gas. 

SorePs  Method.  Sorel  (Ref .  17)  developed  and  applied  the  mathe- 
matical theory  of  the  rectifying  column  for  binary  mixtures.  He  cal- 
culated the  enrichment,  the ^change  in  composition .from  plate  to  plate, 
by  making^energy  and  material  balances  around  each  plate  and 
assumecf  that  equilibrium  was  attained  between  the  vapor  and  liquid 
leaving  the  plate.  He  proceeded  stepwise  through  the  column  T>y 
applying  this  method  successively  from  one  plate  to  the  next. 

Owing  to  the  steady-state  condition  involved  in  continuous  dis- 
tillation, its  analysis  is  simpler  than  batch  operation  and  so  will  be 
considered  first. 

The  equations  for  SorePs  method  will  be  derived  for  the  case  illus- 
trated in  Fig.  7-1.  The  column  is  assumed  to  be  operating  continu- 
ously on  a  binary  mixture  with  the  feed  entering  on  a  plate  between 
the  top  and  bottom.  The  column  is  provided  with  heat  for  reboiling 
by  conduction  such  as  steam  coils  in  the  kettle;  the  case  of  the  use  of 
live  or  open  steam  will  be  considered  later.  A  simple  total  condenser 

118 


RECTIFICATION  OF  BINARY  MIXTURES 


119 


is  assumed  where  all  of  the  overhead  vapor  is  liquefied,  this  condensate 
being  divided  into  two  portions,  one  of  which  is  returned  to  the  column 
for  refly  ~c  and  the  other  withdrawn  as  overhead  product.  The  bottoms 
are  cf  atinuously  withdrawn  from  the  still  or  reboiler.  (See  end  of 
chapter  for  nomenclature.) 

Consider  the  region  bounded  by 
the  dotted  line  in  Fig.  7-1. ,  The 
only  material  entering  this  section 
is  the  vapor  from  the  nth  plate 
Fn,  while  leaving  the  section  is 
the  distillate  D  and  the  overflow 
from  the  (n  +  l)th  plate  On+i. 
By  material  balance, 


Condenser 


m+i 


m-f 


Vn   -   On+1  +  D  (7-1) 

Considering  only  the  more  volatile 
component,  the  mols  entering  this 
section  are  the  total  mols  of  vapor 
from  the  nth  plate  multiplied  by 
the  mol  fraction  of  the  more  vol- 
atile component  in  this  vapor 
Vnyn.  Likewise,  the  mols  of  the 
more  volatile  component  in  the 
distillate  are  DX&;  and  in  the 
overflow  from  the  (n  +  l)th  plate 
are  On+i£n-fi.  A  material  balance 
on  the  more  volatile  component  for  this  section  therefore  gives 


FIG.  7*1. .  Diagram  of  continuous  distills 
tion  column. 


or 


Vny«  = 
yn  - 

On+1 


DxD 

~ 

V  n 


(7-2) 

(7-2a) 

(7-26) 


Thus,  starting  at  the  condenser,  the  composition  of  the  reflux  to  the 
tower,  with  the  type  of  condenser  employed,  is  the  same  as  the  com- 
position of  the  distillate,  which  makes  the  composition  of  the  vapor  to 
the  condenser  the  same  as  that  of  the  distillate.  The  mols  of  vapor 
from  the  top  plate  are  equal  to  OR  +  Z>,  and  the  reflux  to  this  plate  is 
OR,  Sorel's  assumption  of  the  vapor  and  liquid  leaving  the  plate  being 
in  equilibrium,  called  a  theoretical  plate,  makes  it  possible  to  calculate 


120 


FRACTIONAL  DISTILLATION 


the  composition  of  the  liquid  leaving  the  top  plate  of  the  tower  from 
the  composition  of  the  overhead  vapor  and  vapor-liquid  equilibrium 
data. 

In  the  design  of  such  a  tower,  it  is  generally  customary  to  set  or  fix 
certain  operating  variables  such  as  the  composition  of  the  distillate 
and  of  the  bottoms,  the  reflux  ratio  OR/D,  and  the  composition  and 
thermalljondition  of  the  feed.  With  these  values  and  a  known  quan- 
tity of  feed  per  unit  time,  by  over-all  material  balances  it  is  possible 
to  calculate  D,  0Rl  and  W.  To  calculate  the  composition  of  the  vapor 


0.2          0.4  0.6          0.8          1.0 

Mol  Fraction  Benzene  in  Liquid 
Fia.  7-2.     Equilibrium  curve  for  benzene-toluene  mixture. 

from  the  plate  below  the  top  plate  by  Eq.  (7-2a),  it  is  necessary  to 
know  the  mols  of  overflow  from  the  top  plate  and  the  mols  of  vapor 
from  the  plate  below  as  well  as  the  known  quantities  D,  XD,  and  the 
mol  fraction  of  the  more  volatile  component  in  the  liquid  overflow  from 
the  top  plate.  Sorel  obtained  the  mols  of  overflow  from,  and  the  mols 
of  vapor  to,  the  plate  by  heat  (enthalpy)  balances.  Thus  the  heat 
brought  into  the  plate  must  equal  that  leaving. 

0RhR  +  F<_i#f-i  -  VtHt  +  Otht  +  losses  (7-3) 

Equation  (7-3)  gives  a  relation  between  7*_i  and  Ot\  if  the  enthalpies 
are  known,  this  equation  can  be  solved  simultaneously  with  Eq.  (7-1) 
to  give  Vt~i  and  Ot.  This,  in  general,  involves  trial-and-error  solu- 
tions, since  Ht-i  is  not  known  and  must  be  assumed  until  the  condi- 
tions of  the  next  plate  are  known.  The  values  of  F<-i  and  Ot  obtained 
in  this  manner  are  used  in  Eq.  (7-2a)  to  give  #t-~i.  From  this  compo- 


RECTIFICATION  OF  BINARY  MIXTURES  121 

sition  the  value  of  xt~\  is  obtained  from  vapor-liquid  equilibrium  data 
as  well  as  the  temperature  on  this  plate.  The  value  of  Ht~i  can  then 
be  accurately  checked,  and  the  calculations  corrected  if  necessary. 
This  operation  is  continued  plate  by  plate  down  the  tower  to  the  feed 
plate. 

A  similar  derivation  for  the  plates  below  the  feed  gives 

Wxw 

T7 

Wxw  (7  .  . 

-  (7'4a) 


These  equations  are  used  in  the  same  way  as  Eq.  (7-2). 

Because  of  the  complexity  of  SqrePs  method,  it  is  usually  modified 
by  certain  simplifying  assumptions.  The  heat  supply  to  any  section 
of  the  column,  above  the  feed  is  solely  that  of  the  vapor  entering  that 
section.  This  supply  of  heat  to  the  next  plate  goes  to  supply  vapor 
from  this  plate,  to  heat  loss  from  the  section  of  the  tower  that  corre- 
sponds to  this  plate,  and  to  heating  up  the  liquid  overflow  across  this 
plate.  In  a  properly  designed  column,  the  heat  loss  from  the  column 
should  be  reduced  as  far  as  is  practicable  and  is  generally  small  enough 
to  be  a  negligible  quantity  relative  to  the  total  quantity  of  heat  flowing 
up  the  column.  Thus  the  enthalpy  of  the  vapor  per  unit  time  tends 
to  be  constant  from  plate  to  plate,  and  in  order  to  simplify  the  calcu- 
lations for  such  systems,  Lewis  (Ref  .  9)  assumed  that  the  molal  vapor 
rate  from  plate  to  plate  was  constant  except  as  changed  by  additions 
or  withdrawals  of  material  from  the  column.  This  assumption  also 
leads  to  a  constant  overflow  rate  for  such  a  section.  In  the  case  illus- 
trated in  Fig.  7-1,  this  simplifying  assumption  would  give  constant 
vapor  and  overflow  rates  above  and  below  the  feed  plate,  but  the  rates 
in  the  two  sections  of  the  tower  would  be  different  due  to,  the  introduc- 
tion of  the  feed.  This  assumption,  together  with  the  theoretical 
plate  concept,  has  been  of  great  assistance  in  the  analysis  and  design 
of  fractionating  column.  The  validity  of  these  two  assumptions  will 
be  considered  in  later  sections. 

On  the  basis  of  Lewis'  assumption  On+i  and  Vn  are  constant  in  the 
section  above  the  feed  plate,  and  the  relation  between  yn  and  x*+\ 
becomes  a  straight  line  with  the  slope  equal  to  0/7.  Similarly,  below 
the  feed,  ym  is  linear  in  xm+i.  On  the  basis  of  the  operating  variables 
previously  fixed,  On+i,  F»,  D,  and  XD  are  known,  and  the  equation 
between  yn  and  #n+i  is  completely  defined;  likewise  for  ym  and  xm+i. 

A  plate  on  which  SorePs  conditions  of  equilibrium  are  attained  is 


122  FRACTIONAL  DISTILLATION 

defined  as  a  "theoretical  plate/'  i.e.)  a  plate  on  which  the  contact 
between  vapor  and  liquid  is  sufficiently  good  so  that  ffie  vapor  leaving 
the  plate  has  the  same  composition  as  the  vapor  in  equilibrium  with 
the  overflow  from  the  plate.  For  such~a  pfate  the  vapor  and  liquid 
leaving  are  related  by  the  equilibrium  y,x  curve  (see  page  18).  Rec- 
tifying columns  designed  on  this  basis  serve  as  a  standard  for  compar- 
ing actual  columns.  By  such  comparisons  it  is  possible  to  determine 
the  number  of  actual  plates  equivalent  to  a  theoretical  plate  and  then 
to  reapply  this  factor  when  designing  other  columns  for  similar  service. 

Sorel-Lewis  Method.  As  an  illustration  of  the  Sorel-Lewis  method,  consider  the 
rectification  of  a  50  mol  per  cent  benzene  and  50  mol  per  cent  toluene  mixture  into  a 
product  containing  5  mol  per  cent  toluene  and  a  bottoms  containing  5  mol  per  cent 
benzene.  The  feed  will  enter  as  a  liquid  sufficiently  preheated  so  that  its  intro- 
duction into  the  column  does  not  affect  the  total  mols  of  vapor  passing  the  feed 
plate;  i.e.,  such  that  F»  -»  Vm.  A  reflux  ratio  On/A  equal  to  3,  will  be  employed, 
and  the  column  wfll  o'pef  ate  with  a  total  condenser  and  indirect  heat  in  the  still. 
The  y,x  equilibrium  curve  is  given  in  Fig.  7-2. 

Taking  as  a  basis  100  Ib.  mols  of  feed  mixture,  an  over-all  benzene  material 
balance  on  the  column  gives 

0.5(100)  -  0.95D  -f  0.05TF 

*  0.95D  -f  0.05(100  -  D) 

gives 

D  »  50  Ib.  mols 

W  «  50  Ib.  mols 
Since 

On         « 

D       3 

On   -  150 

Fft   -  On  +  D  =  200 

by  Eq.  (7-2a), 

V*  -  (15%oo)*«+i  +  (5%oo)(0.95)  -  0.75sn+i  +  0.2375  (7-5) 


Since  a  total  condenser  is  used, 

yt  «  XD  =  XR  =»  0.95 

from  the  equilibrium  curve  at"jT*=  0.95,  x  *»  0.88;  i.e.,  xt  in  equilibrium  with  yt  is 
0.88, 

Equation  (7-5)  then  gives 

yM  -  0.76x«  -f  0.2375  -  0.75(0.88)  -f  0.2375  -  0.8975 
by  equilibrium  curve,  Xt-i  at  y^i  **  0.8975  is  0.77  and 

yi_2  *  0.75(0.77)  4*  0.2375  -  0.8145 

ak-i  -  0.64 

yt-3  -  0.75a;«-2  +  0.2375  -  0.75(0.64)  -f  0.2375  -  0.7165 

St.,  «  0.505 


RECTIFICATION  OF  BINARY  MIXTURES  123 

Since  thejraJue  of  Xt-s  is  close  to  the  composition  of  the  feed,  this  plate  will  be 
taken  asThe  feed^E*!^  portion  of 

the  tower  must  be  used.    Since  the  feed  was  preheated  such  that 

Vn    **    V 

Y!  -  200 
W-SQ 

.a.-*o 

and 

i  -  (5%oo)(0.05)  -  1.25**+i  -  0.0125 


since  xt-s  «  a?/  =  0.505 


y/_i  -  1.25(0.505)  -  0.0125  -  0.615 

aj/-1  »  0.392  from  equilibrium  curve 

y/_2  -  1.25(0.392)  -  0.0125  «  0.478 

*/-«  -  0.275 

^/_8  -  1.25(0.275)  -  0.0125  *  0.323 

as/-,  -  0.172 

y/-4  -  1.25(0.172)  -  0.0125  -  0.21 

z/_4  -  0.100 

y/_6  -  1.25(0.100)  -  0.0125  -  0.122 

z/_6  -  0.058 

2//_8  -  1.25(0.058)  -  0.0125  - 


The  desired  strength  of  the  bottoms  was  xw  ~  0.05;  #/_6  is  too  high,  and  #/_e 
is  too  low.  Thus  it  is  impossible  to  satisfy  the  conditions  chosen  and  introduce 
the  feed  on  the  fourth  plate  from  the  top  with  an  even  number  of  theoretical  plates. 
Hdwever,  by  slightly  reducing  the  reflux  ratio  it  would  be  possible  to  make  s/-8 
equal  to  xw  ,  or  by  increasing  the  reflux  ratio  to  make  #/_6  equal  to  xw*  In  general, 
such  refinements  are  not  necessary,  and  it  is  sufficient  to  say  that  between  eight 
and  nine  theoretical  plates  are  required  in  addition  to  the  still,  three  plates  above 
the  feed  plate,  the  feed  plate,  and  four  or  five  plates  below  the  feed,  and  the  still, 
approximately  8M-  The  percentage  difference  between  eight  and  nine  is  much 
less  than  the  accuracy  with  which  the  ratio  of  actual  to  theoretical  plates  is  known; 
whichever  is  used,  a  sufficient  factor  of  safety  must  be  utilized  to  cover  the  varia- 
tion of  this  latter  factor. 

McCabe  and  Thiele  Method  (Ref  .  11).  By  the  Sorel-Lewis  method, 
the  relation  between  yn  and  xn+i  is  a  straight  line,  and  the  equation  of 
this  line  may  be  plotted  on  the  y,x  diagram.  Thus,  for  the  example 
worked  in  the  preceding  section, 

0.2375 


This  is  a  straight  line  of  slope  0.75  *»  On/Vn  which  crosses  the  y  *»  x 
diagonal  at  yn  »  x«+i  -  0.95  -  XD.  On  the  y,x  diagram  for  benzene- 
toluene,  a  line  of  slope  0.75  is  drawn  through  y  *  x  «  XD  (see  line 


124  FRACTIONAL  DISTILLATION 

AB,  Fig.  7-3).    Likewise,  below  the  feed, 
0 


-  0-0125 


This  represents  a  straight  line  of  slope  Om/Vm  =  1.25  and  passes 
through  the  y  =  x  diagonal  &t  x  —  Xw  ~  0.05  (see  line  CD,  Fig.  7-3) 
These  two  lines  are  termed  the  operating  lines,  since  they  are  deter- 


0.2  0,4  0.6 

x=Mol  Fraction  Benzene  in  Liquid 

FIG.  7-3.     McCabe  and  Thiele  diagram. 


0.8 


1.0 


mined  by  the  tower  operating  conditions,  AB  being  the  operating  line 
for  the  enriching  section  and  CD  the  operating  line  for  the  stripping, 
or  exhausting,  section.  To  determine  the  number  of  theoretical 
plates  ty  £ig.  7-3,  start  at  XD]  as  before,  yt  =  XD  ==  0.95,  and  the  value 
of  Xt  is  determined  by  the  intersection  of  a  horizontal  line  through 
yt  **  0.95  with  the  equilibrium  curve  at  1,  giving  xt  =  0.88.  Now, 
irustead  of  using  Eq.  (7-2a)  algebraically  as  in  the  Sorel-Lewis  method, 
it  is  used  graphically  as  the  line  AB.  A  vertical  line  at  xt  =  0.88 
intersects  this  operating  line  at  2,  giving  yt~i  -  0.89.  By  proceeding 


RECTIFICATION  OF  BINARY  MIXTURES  125 

horizontally  from  intersection  2,  an  intersection  is  obtained  with  the 
equilibrium  curve  at  3.  Since  the  ordinate  of  intersection  3  is  yt-i,  the 
abscissa  must  be  the  composition  of  the  liquid  in  equilibrium  with  this 
vapor;  i.e.,  Xt-\  =  0.77.  As  before,  the  intersection  of  the  vertical 
line  through  the  point  3  with  the  operating  line  at  4  gives  the  y  on  the 
plate  below,  or  yt~2  =  0.815.  This  stepwise  procedure  is  carried  down 
the  tower.  At  intersection  8,  xt-s  is  approximately  equal  to  xy;  and 
at  this  plate,  the  feed  will  be  introduced.  The  stepwise  method  is  now 
continued,  using  the  equilibrium  curve  and  the  operating  line  CD. 

Such  a  stepwise  procedure  must  yield  the  same  answer  as  the  previ- 
ous calculations,  since  it  is  the  exact  graphical  solution  of  the  algebraic 
equation  previously  used.  It  has  a  number  of  advantages  over  the 
latter  method:  (1)  It  allows  the  effect  of  changes  in  equilibrium  and 
operating  conditions  to  be  visualized.  (2)  Limiting  operating  coudi- 
tion§~afe  easily  determined,  and  if  a  column  contains  more  than  two  or 
thTee~ptates,  it  is  generally  more  rapid  than  the  corresponding  algebraic 
procedure.  Because  of  the  importance  of  this  diagram  it  will  now  be 
considered  in  further  detail. 

Intersection  of  Operating  Lines.  In  Fig.  7-3,  the  operating  lines 
intersected  at  x  =  XF-  This  intersection  is  not  fortuitous,  since  the 
positions  of  the  two  operating  lines  are  not  independent  but  are  related 
to  each  otherHby  the  composition  and  thermal  condition  of  the  feed. 
This~relation  is  mtf^Teasity  "showtf  by :  writing  a  Heat  balance  around 
the  feed  plate.  Let  p  be  the  difference  between  the  mols  of  overflow 
to  and  from  the  feed  plate  divided  by  the  mols  of  feed. 

P  =  °f+1~  °f  (7-6) 

A  material  balance  gives 

p  +  1  =  Vf  ~FVf~l  (7-6a) 

Let  Xi  and  yi  be  the  coordinates  of  the  intersection  of  the  operating 
lines.  At  this  intersection,  yn  must  equal  ym,  and  xn  must  equal  xm. 
An  over-all  material  balance  on  the  more  volatile  component  gives 
DXD  +  Wxw  =  FzFj  where  ZF  is  the  average  mol  fraction  of  this  com- 
ponent in  the  feed.  Writmg  Eqs.  (7-2a)  and  (7-4)  for  the  intersection 
and  using  the  values  y%  and  x%J 

-  DxD  (7-2a') 

-  Wxw  (7-4') 


126  FRACTIONAL  DISTILLATION 

and  subtracting, 

(7,  -  7/-OW  -  (0/+i  -  0,)a*  +  Dar^ 
«  (0/+i  -  0/)a*  +  FzF 
(Vf  -  7/-Jy,  _  /Q/+!  -  O 

-  - 


Substituting  values  of  p  and  p  +  1  gives  the  point  on  the  diagram  at 
which  the  intersection  must  occur. 

(p  +  l)yi  -  pxi  +  zf 


Equation  (7-7)  together  with  Eq.  (7-26)  gives 

— (7-8) 

and 


= 
'         (0/D)  -  p 

This  line  of  intersections  crosses  the  y  =  x  diagonal  at 

y.  =  Xi  =  2j. 

and  has  a  slope  of  p/(p  +  1).  The  effects  of  various  values  of  p  are 
shown  in  Fig.  7-4  for  a  given  slope  of  the  operating  line  above  the  feed. 
Thus,  if  p  «  0,  the  mols  of  overflow  above  and  below  the  feed  are 
equal,  and  the  operating  lines  must  intersect  in  a  horizontal  line 
through  the  diagonal  at  Zr.  A  value  of  p  =  —  1,  i.e.,  F/  =  7/-i, 
would  put  the  intersection  on  a  vertical  line  at  ZF. 

The  value  of  p  is  best  obtained  by  an  enthalpy  balance  around  the 
feed  plate.  However,  when  the  molal  enthalpy  of  the  overflow  from 
the  feed  plate  and  the  plate  above  is  essentially  the  same  and  the 
enthalpy  of  the  vapor  from  the  feed  plate  and  the  plate  below  is  also 
the  same,  then,  by  Eqs.  (7-6)  and  (7-6a),  —  p  becomes  approximately 
the  heat  necessary  to  vaporize  1  Ib.  mol  of  the  feed  divided  by  the 
latent  heat  of  vaporization  of  the  feed.  Thus,  an  all-vapor  feed  at  its 
boiling  point  would  have  a  value  of  p  =  0,  for  an  all-liquid  feed  at  its 
boiling  point,  p  would  equal  —  1;  p  would  be  less  than  —  1  for  a  cold 
feed,  between  —  1  and  zero  for  a  partially  vaporized  f eed7  and  greater 
than  zero  for  a  superheated  vapor  feed. 


RECTIFICATION  OF  BINARY  MIXTURES 


127 


A  little  study  of  Fig.  7-4  indicates  that  for  a  given  0/D  fewer  plates 
are  required  for  a  given  separation  the  colder  the  feed.  This  results 
from  the  fact  that  the  cold  feed  condenses  vapor  at  the  feed  plate  and 
increases  the  reflux  ratio  in  the  lower  portion  of  the  column.  This 
higher  reflux  ratio  is  obtained  at  the  expense  of  a  higher  heat  con- 
sumption in  the  still. 


xw       Mol  Fraction  in  Liquid 

FIG.  7-4.     The  effect  of  the  thermal  condition  of  the  feed  on  the  intersection  of  the 
operating  lines. 

1,  pis  greater  than  0  (superheated  vapor  feed) 

2,  p  -  0  (O/+i  -  Of) 

3,  0  >  p  >  —  1  (partly  vapor  feed) 
4fp  -  -KF/-1  -  F/) 

5,  p  <  —  1  (cold  feed) 

Logarithmic  Plotting.  When  the  design  involves  low  concentrations 
at  the  terminals  of  the  tower,  it  is  necessary  to  expand  this  part  of  the 
diagram  in  order  to  plot  the  steps  satisfactorily.  This  may  be  done  by 
redrawing  these  regions  of  the  y,x  diagram  to  a  larger  scale.  In  some 
cases,  it  may  be  necessary  to  make  more  than  one  expansion  of  suc- 
cessive portions  of  the  diagram.  Alternately,  the  y,x  diagram  may  be 
plotted  on  logarithmic  paper,  and  the  steps  constructed  in  the  usual 
manner.  On  this  type  of  plot  in  the  low-concentration  region,  the 
equilibrium  curve  is  generally  a  straight  line,  since,  for  small  values  of 

becomes  y  —  ax]  however,  the  operating  line 


x:y 


ax 


1  +  (a  -  l)x 


128 


FRACTIONAL  DISTILLATION 


which  is  of  the  form  ym  =  axm+i  +  6  is  a  curved  line  unless  6  =  0. 
The  operating  line  is  constructed  from  points  calculated  from  the 
operating-line  equation. 

Minimum  Number  of  Plates.  The  slope  of  the  operating  line  above 
the  feed  is  On/Vn,  and  as  this  slope  approaches  unity  the  number  of 
theoretical  plates  becomes  smaller.  When  On/Vn  is  equal  to  1,  0R/D 
is  equal  to  infinity,  and  only  an  infinitesimal  amount  of  product  can 
be  withdrawn  from  a  finite  column.  Frequently  it  is  assumed  that 
total  reflux  corresponds  to  the  addition  of  no  feed  or  to  the  removal  of 
no  products.  If  such  is  the  case,  the  tower  is  not  meeting  the  design 
conditions.  It  is  better  to  visualize  a  tower  with  an  infinite  cross 
section,  which  is  separating  the  feed  at  a  finite  rate  into  the  desired 

products.     Under  such  conditions 

the  column  is  said  to  operate  at 
total  reflux  or  with  an  infinite  re- 
flux ratio,  and  both  operating  lines 
have  a  slope  of  unity  causing  them 
to  coincide  with  the  y  =  x  diago- 
nal. Since  a  higher  reflux  ratio 
than  this  is  not  possible,  the  size  of 
the  steps  on  the  y,x  diagram  is  a 
maximum,  and  a  minimum  num- 
ber of  theoretical  plates  to  give  a 
given  separation  is  obtained. 
This  number  is  determined  by 
simply  using  the  y  =  x  diagonal 
as  the  operating  line  and  con- 
A  column  with  the  minimum  num- 
ber of  plates  serves  as  a  reference  below  which  no  column  with  fewer 
plates  can  give  the  desired  separation,  but  such  a  column  would  have 
a  zero  capacity  per  unit  volume  and  would  require  infinite  heat  con- 
sumption per  unit  of  product. 

Minimum  Reflux  Ratio.  In  general,  it  is  desired  to  keep  the  reflux 
ratio  small  in  order  to  conserve  heat  and  cooling  requirements.  As 
the  reflux  ratio  0R/D  is  reduced  from  infinity,  the  slope  of  the  operating 


x«Mol  Fraction  in  Liquid 
FIG.  7-5.     Plot  for  minimum  reflux  ratio. 

structing  the  steps  from  XD  to  xw . 


linc  0*  „       0/D 
Vn      (0/D)  +  I 


decreases  from  unity.     Thus,  in  Fig.  7-5  a  reflux 


ratio  of  infinity  would  correspond  to  operating  lines  coinciding  with 
the  diagonal  as  acb,  and  a  lower  reflux  ratio  would  correspond  to  adb. 
It  is  obvious  that  the  average  size  of  the  steps  between  the  equilibrium 
curve  and  the  line  adb  will  be  much  smaller  than  the  size  of  the  steps 


RECTIFICATION  OF  BINARY  MIXTURES  129 

between  the  equilibrium  curve  and  the  line  acb.  Thus,  a  reduction  of 
the  reflux  ratio  requires  an  increase  in  the  number  ^TflEeoretical  plates 
to  effect  a  given_sejgara|raa;  S1the"re9ux  ratio  iifuriber  decreased, 
thcTsfze  oTthe  steps  between  the  operating  lines  and  the  equilibrium 
curve  becomes  still  smaller,  and  still  more  theoretical  plates  are 
required,  until  the  conditions  represented  by  aeb  are  encountered,  when 
the  operating  line  just  touches  the  equilibrium  curve.  In  this  final 
case,  the  size  of  the  step  at  the  point  of  contact  would  be  zero,  and  an 
infinite  number  of  plates  would  be  required  to  travel  a  finite  distance 
down  the  operating  line.  The  reflux  ratio  corresponding  to  this  case 
is  called  the  minimum  reflux  ratio  and  represents  the  theoretical  limit 
below  wMch^tMtf*^^  Be  ireHuced  and  produce  the  desired 

separation  even  if  an  infinite  column  is  employed.  This  reflux  ratio 
is  easily  determined  by  laying  out  the  operating  line  of  the  flattest 
slope  through  XD  that  just  touches  but  does  not  cut  the  equilibrium 

curve  at  any  point;  the  slope  of  this  line  ^  =  (n/  ^  ._  .  gives  the 
value  of  0/D.  Alternately,  it  may  be  calculated  from  the  equation 

(7-10) 
' 


D       yc  -  xc 

where  xc  and  yc  are  the  coordinates  of  the  point  of  contact.  For  mix- 
tures having  normal-shaped  equilibrium  curves,  such  as  benzene- 
toluene,  the  point  of  contact  of  the  operating  line  with  the  equilibrium 
curve  will  occur  at  the  intersection  of  the  operating  lines.  For  cases 
that  deviate  widely  from  Ilaoult's  law,  the  operating  line  may  become 
tangent  to  the  equilibrium  curve  before  the  intersection  of  the  operat- 
ing lines  touches  the  equilibrium  curve,  and  in  such  cases  it  is  usually 
best  to  plot  the  diagram  and  determine  the  slope  On/Vn> 

Optimum  Reflux  Ratio.  The  choice  of  the  proper  reflux  ratio  is  a 
matter  of  economic  balance.  At  the  minimum  reflux  ratio,  fixed 
charges  are  infinite,  because  an  infinite  number  of  plates  is  required. 
At  total  reflux,  both  the  operating  and  the  fixed  charges  are  infinite. 
This  is  due  to  the  fact  that  an  infinite  amount  of  reflux  and  a  column 
of  infinite  cross  section  would  be  required  for  the  production  of  a  finite 
amount  of  product.  The  tower  cost  therefore  passes  through  a  mini- 
mum as  the  reflux  ratio  is  decreased  above  the  minimum.  The  costs 
of  the  still  and  condenser  both  increase  as  the  reflux  ratio  is  increased. 
The  heat  and  cooling  requirements  constitute  the  main  operating 
costs,  and  the  sum  of  these  increases  almost  proportionally  as  the 


130 


FRACTIONAL  DISTILLATION 


reflux  ratio  is  increased.  The  total  cost,  the  sum  of  operating  and 
fixed  costs,  therefore  passes  through  a  minimum. 

Optimum  Reflux  Ratio  Example.  The  following  estimates  illustrate  these 
factors  for  the  fractional  distillation  of  a  methanol-water  mixture  to  produce  250 
gal.  of  methanol  per  hour.  In  making  the  calculations,  it  was  assumed  that  the 
heat-transfer  surface  required  was  proportional  to  the  vapor  rate  which  is  equal  to 

D  (  £  +  1 )  and  that  the  tower  costs  were  proportional  to  the  total  square  feet  of 

plate  area.  The  charges  on  the  equipment,  including  maintenance,  repairs,  depre- 
ciation, interest,  etc,,  were  taken  at  25  per  cent  per  year,  and  the  heating  costs  were 
based  on  the  heat  load.  The  costs  per  hour  as  a  function  of  the  reflux  ratio  are 
summarized  in  Table  7-1 .  Labor  charges  have  been  excluded  since  these  should  be 

TABLE  7-1.    ESTIMATED  COST  FOB  THE  FRACTIONATION  OF  A  METHANOL- WATER 

MIXTURE 


Costs,  dollars  per  hour 

f) 

D 

V 

Charges 
on  tower 

Charges  on  con- 
denser and  reboiler 

Steam  and 
cooling  water 

Total 

0  65 

0.39 

00 

0.023 

0.33 

00 

0.68 

0.40 

0.11 

0.024 

0.335 

0.47 

0  71 

0.41 

0  09 

0.025 

0.34 

0  455 

0.84 

0  46 

0.064 

0.026 

0.37 

0.46 

1.1 

0  52 

0  057 

0.030 

0.42 

0  51 

1.6 

0.61 

0.06 

0.037 

0.52 

0.62 

2.6 

0.72 

0.07 

0.051 

0.71 

0.82 

6.5 

0.87 

0.12 

0.017 

1.48 

1.71 

QO 

1.0 

00 

00 

oo 

OO 

relatively  independent  of  the  reflux  ratio.  These  results  are  plotted  in  Fig.  7-6. 
It  will  be  noted  that  the  total  of  these  costs  passes  through  a  minimum  at  a  reflux 
ratio,  0/V,  equal  to  about  0.43,  (0/D  -  0.75).  This  is  very  close  to  the  mini- 
mum reflux  ratio,  0.65  and  is  a  result  of  the  fact  that  the  heating  and  cooling  costs 
are  large  and  increase  rapidly  with  the  reflux  ratio. 

The  calculated  economic  reflux  ratio  for  most  cases  is  so  close  to  the  minimum 
reflux  ratio  that  the  accuracy  of  the  latter  becomes  a  critical  matter.  In  the  pres- 
ent case,  the  most  economical  reflux  ratio,  0/D,  is  only  15  per  cent  above  the  mini- 
mum, and  it  is  doubtful  whether  the  equilibrium  data  available  are  sufficiently 
accurate  to  make  the  calculation  of  the  minimum  reflux  ratio  better  than  ±10 
per  cent.  For  this  reason,  it  is  industrial  practice  to  employ  a  reflux  ratio  some- 
what higher  than  the  most  economical,  and  values  of  J.3  to  2  times  the  minimum 
reflux  ratio  are  common.  For  the  case  here  considered,  reflux  ratios  in  this  range 
would  give  operating  costs  only  slightly  greater  than  the  minimum.  This  small 
increase  in  cost  gives  a  design  that  will  be  less  sensitive  to  slight  inaccuracies  in  the 
data  employed. 


RECTIFICATION  OF  BINARY  MIXTURES 


131 


One  of  the  important  pieces  of  data  needed  for  such  an  economic 
study  is  the  number  of  theoretical  plates  as  a  function  of  the  reflux 
ratio.  Approximate  methods  for  the  rapid  estimation  of  such  data  are 
given  in  Chap.  12,  page  348. 


to 

0.9 
08 
,07 

•?0.6 
S. 

I05 
•"0.4 

02 
O.I 

°c 

^ 

\ 

/ 

T 

To  o» 

*/7 

/  /i 

f    ' 

/  u  — 

D/ 

Toe 

to 

/ 

> 

'* 

a/V 

r  -O.v 

5? 

[s 

/ 

* 

***«..•» 

^ 

^ 

/A 

-Cost 
-Chat 
-Chat 
-  To/a 

sofs 

*ges  i 

"ges  c 
r/of. 

team 
in  fra 
m  sti 
A*B* 

Me/cooling  watei 
icfionafing  co/urr 

fl  and  conofenset 
/% 

, 

t 

^^ 

^ 

C 
n 

n 

'.§ 

U 

C 

*1 

^^j 

\  —  i 

$  -0.39 

—  B~ 

.To  oo  at 

to?  • 

7-> 

h 

- 

C 

'\ 

,3             0.4              0.5              06              07              0.8              09             1 

0. 
V 
FIG.  7-6.     Optimum  reflux  ratio. 

Feed-plate  Location.  One  step  between  the  equilibrium  curve  and 
the  operating  line  for  the  enriching  section  corresponds  to  one  the- 
oretical plate  in  the  enriching  section  above  the  feed,  and  one  step 
between  the  equilibrium  curve  and  the  other  operating  line  corresponds 
to  one  theoretical  plate  below  the  feed.  Therefore  the  step  that  passes 
from  one  operating  line  to  the  other  corresponds  to  the  feed  plate. 
Thus,  in  Fig.TjT^jr^^  the  operating  line 

a6cj  it  is  not  possible  toji£iLQp.-tihe-  ^p^M^g^3^-^^  "  ntil  the  yalue  jpf 
x  IslesS  tliiaKTHe^Sue  correspondinjgj2,J^in^  e-    However,  as  soon  as 

possible  to  shift  to  the  other  operating 


line,  but  it  is  not  necessary  to  do  so  at  this  value,  since  steps  can  be 
continued  down  abc  until  they  are  pinched  in  at  c,  but  a  value  less  than 
c  cannot  be  obtained  unless  the  shift  is  made.  The  step  from  one 


132 


FRACTIONAL  DISTILLATION 


operating  line  to  the  other  must  therefore  occur  at  some  value  of  x 
between  the  values  corresponding  to  c  and  e,  and  a  change  at  any 
value  within  this  range  will  give  an  operable  design.  In  general,  for  a 
given  reflux  ratio,  it  is  desired  to  carry  out  the  rectification  with  as  few 
plates  as  possible  in  order  to  reduce  the  plant  costs;  i.e.,  the  minimum 
number  of  steps  from  a  to  d  between  the  equilibrium  curve  and  the 
operating  line  is  desired.  This  minimum  number  of  steps  for  the 
design  conditions  selected  is  obtained  by  taking  the  largest  possible 
steps  at  all  points  between  a  and  d.  It  is  obvious  that  for  values 
between  e  and  b  larger  steps  will  be  obtained  between  the  equilibrium 

curve  and  operating  line  abc  than 
would  be  obtained  with  operating 
line  dbe.  Likewise,  for  values  be- 
tween c  and  b  larger  steps  will  be 
obtained  by  using  line  dbe  than  by 
using  abc.  Therefore  it  is  desir- 
able to  use  operating  line  abc  for 
values  from  a  to  6,  and  line  dbe  for 
values  between  d  and  6;  and  by 
making  the  feed  plate,  i.e.j  the 
shift  from  one  line  to  the  other, 
straddle  the  value  f>,  the  minimum 
number  of  theoretical  plates  will 
be  obtained  for  the  operating  con- 
ditions chosen.  If  a  step  happens 
to  fall  directly  on  6,  then  the  feed  may  be  introduced  either  at  b  or  on 
the  plate  below  without  changing  conditions. 

X  Partial  vs.  Total  Condenser,  In  the  foregoing  discussion  the 
column  was  assumed  to  be  operating  with  a  total  condenser,  i.e.,  a 
condenser  that  completely  liquefies  the  overhead  vapor  and  returns  a 
portion  of  the  condensate  as  reflux,  removing  the  remainder  as  product. 
However,  partial  condensers  are  quite  frequently  used  in  commercial 
operations,  especially  where  complete  liquefaction  of  the  overhead 
would  be  difficult.  In  this  case,  only  enough  condensate  for  the  reflux 
to  the  column  may  be  produced,  and  the  product  is  withdrawn  as  a 
vapor.  In  other  cases,  mixtures  of  vapor  and  liquid  are  withdrawn. 
For  example,  in  gasoline  stabilizers  employed  by  the  petroleum  indus- 
try, where  the  overhead  contains  appreciable  percentages  of  methane, 
ethane,  and  ethylene,  together  with  Cs  and  €4  hydrocarbons,  in  order 
to  condense  the  methane  and  C2  hydrocarbons,  very  low  temperatures 


FIG.  7-7.     Diagram  for  limits  of  feed-plate 
composition. 


RECTIFICATION  OF  BINARY  MIXTURES  133 

would  be  required  with  resulting  high  refrigeration  costs.  However, 
sufficient  of  the  C8  and  C4  hydrocarbons  can  be  liquefied  at  moderate 
temperatures  and  pressures  to  serve  as  reflux,  and  the  remainder  of  the 
overhead  containing  a  large  portion  of  the  Ci  and  C%  hydrocarbons  can 
be  removed  as  vapor  and  sent  to  the  gas  lines. 
A  partial  condenser  may  operate  in  any  of  several  ways : 

1.  The  cooling  may  be  so  rapid  and  the  contact  between  condensate 
and  uncondensed  vapor  so  poor  that  essentially  no  transfer  of  com- 
ponents back  and  forth  is  obtained,  with  the  result  that  the  condensate 
and  uncondensed  vapor  are  of  the  same  composition.     (This  is  possible 
if  part  of  the  vapor  condenses  completely  and  the  balance  does  not 
condense  at  all.)     In  this  case,  the  partial  condenser  is  equivalent  to 
the  total  condenser  with  the  exception  that  the  product  is  removed  as 
vapor  instead  of  as  liquid. 

2.  The  vapor  product  may  be  in  sufficiently  good  contact  with  the 
returning  reflux  for  the  two  to  be  in  equilibrium  with  each  other,  in 
which  case  the  partial  condenser  acts  as  a  theoretical  plate,  and  one 
less  theoretical  plate  may  be  used  above  the  feed  plate  in  the  column 
when  this  condition  exists  than  when  a  total  condenser  is  employed. 
Such  a  condition  can  be  approximated  by  requiring  an  overhead  vapor 
to  bubble  through  a  pool  of  reflux  to  the  column. 

3.  The  vapor  is  differentially  condensed,  and  the  equilibrium  con- 
densate continually  removed,  giving  a  differential  partial  condensation. 
Alternately,  the  vapor  may  be  condensed  on  vertical  tubes  such  that 
the  condensate  flows  countercurrent  to  the  rising  vapor,  and  fractiona- 
tion  occurs  between  the  vapor  and  condensate.    Theoretically,  such  a 
condenser  can  give  a  separation  equal  to  a  number  of  theoretical 
plates;  actually,  such  conditions  are  seldom  employed,  since  to  obtain 
efficient  transfer  of  components  from  vapor  to  liquid,  low  rates  of  con- 
densation per  unit  area  are  required,  thus  necessitating  large  and  costly 
condensers,  and,  in  general,  it  is  found  more  satisfactory  and  cheaper 
to  obtain  additional  rectification  by  adding  more  plates  to  the  column 
and  using  a  condenser  to  produce  condensate  rather  than  make  it  per- 
form composite  duties. 

Actual  partial  condensers  usually  operate  somewhere  between  Cases 
1  and  2.  For  an  absorption  naphtha  stabilizer,  Gunness  (Ref .  6)  (see 
page  116)  found  good  agreement  with  Case  2.  In  actual  design  cal- 
culation, the  conservative  assumption  is  to  assume  operation  as  in  Case 
1,  and  any  fractionation  that  does  occur  will  act  as  a  factor  of  safety; 
with  ordinary  condenser  design,  with  the  most  optimistic  assumption, 


134 


FRACTIONAL  DISTILLATION 


m-H 


not  more  than  one  theoretical  plate  should  be  taken  for  the  partial 
condenser. 

Open  vs.  Closed  Steam.  When  rectifying 
mixtures  in  which  the  residue  is  water  and  in 
some  cases  where  the  mixture  undergoing  f  rac- 
tionation  is  immiscible  with  water,  the  steam 
for  heating  may  be  introduced  directly  into  the 
still.  Such  a  procedure  may  materially  reduce 
the  temperature  and  pressure  of  the  steam  nec- 
essary for  the  distillation  by  giving  in  effect  a 
steam  distillation. 

Considering  the  distillation  of  an  ethyl  alco- 
hol-water mixture,  the  lower  operating  line  when 
a  closed-steam  heating  is  used  was  shown  to 
have  a  slope  of  (0/V)m  and  to  pass  through  the 

y  =  X  line  at  X  =  Xw.     In  Fig.  7-8,  a  column  Op- 
erating  ^^  g  molg  rf  jjve  gteam 


Fia,  7-8.     Diagram  of 
column  using  live  steam. 

material  balance  between  the  m  and  m  +  1  plate  gives 


o, 


.M.I 


3  -  Vm  +  W 


(7-11) 


and  with  the  usual  simplifying  assumptions,  S  would  equal  Fm,  making 
0  =  W.     An  alcohol  balance  gives 


W 


~ 

y  m 


This  is  an  operating  line  of  slope  Om/Vm;  but  at  x  =  y,  x  is  equal  to 

(w       ciixw  instead  of  xw]  and  at  x  equal  to  3wr,  y  becomes  zero 
W  -  S/ 

corresponding  to  the  composition  of  the  vapor  (steam)  to  the  bottom 
plate.  For  a  given  0/D  and  feed  condition,  Om/Vm  must  be  the  same 
whether  closed  or  open  steam  is  used,  so  that  the  lower  operating  line 
must  cross  the  y  =  x  diagonal  at  the  same  x  value  in  both  cases,  Xw  for 
the  live  steam  being  lower  than  xw  for  closed  steam  because  of  the 
dilution  effect  of  the  steam.  In  stepping  off  theoretical  plates,  the 
step  must  start  at  y  =  0  and  Xw]  i.e.,  in  Fig.  7-9,  the  bottom  plate 
corresponds  to  the  step  dbc.  In  such  a  case,  the  introduction  of  live 
steam  can  eliminate  the  still,  but  it  dilutes  the  bottoms  and  requires 
more  plates  in  the  lower  section  of  the  column.  Since  the  steps  in  the 


RECTIFICATION  OF  BINARY  MIXTURES 


135 


case  of  live  steam  start  lower,  it  always  requires  at  least  a  portion  of  a 
step  to  come  up  to  the  intersection  of  the  operating  line  and  the  y  =  # 


100  mots  Oil  \       f 


2,SSmols  C3Hfl 
4/no/s  Steam 


FIG,  7-9.     y,x  diagram  for  case  of  live  steam. 

diagonal,  and  more  plates  are  required  with  live  steam  than  with 
closed  steam.  In  Fig.  7-9,  about  1J^  more  plates  would  be  necessary. 
One  plate  is  needed  to  replace  the  still, 
and  the  additional  "fraction  of  a 
plate"  is  required  to  offset  the  dilu- 
tion. 

As  an  example  of  using  open  steam 
to  obtain  a  steam  distillation,  consider 
the  steam  stripping  of  an  oil  containing 
2.54  mol  per  cent  propane  at  20  p.s.i. 
The  temperature  will  be  maintained 
constant  at  280°F.  by  internal  heat- 
ing. The  molecular  weight  of  oil  may 
be  taken  as  300,  and  4  mols  of  steam 
will  be  used  per  100  mols  of  oil  stripped. 
It  is  desired  to  estimate  the  number  of 


4  mols  Steam 


FIG.  7-10. 


fOOmo/s  Oil 
O.OSmols  CjHg 

Figure  for  illustration. 


theoretical  plates  necessary  to  reduce  the  propane  content  of  the  oil  to 
0.05  mol  per  cent.    The  oil  may  be  assumed  nonvolatile,  and  the  vapor- 


136 


FRACTIONAL  DISTILLATION 


liquid  relation  of  the  propane  in  the  oil  may  be  expressed  as  y  =  33.4$. 
It  is  obvious  that  the  mols  of  vapor  will  increase  up  the  tower,  since 
the  steam  does  not  condense  under  the  conditions  given,  and  the 
propane  vaporizes  into  it  as  it  passes  up  the  tower.  This  will  cause 
0/V  to  vary  through  the  tower,  and  points  on  the  operating  line  must 


0.50 


0.01 
Mol  Fraction 


FIG.  7-11. 


0.02 
in  Liquid 

Steam  stripping  diagram. 


be  calculated,  since  it  will  not  be  a  straight  line  on  a  y,x  diagram. 
This  is  easily  done  by  taking  a  basis  of  100  mols  of  entering  oil,  for 
which  the  terminal  conditions  are  given  in  Fig.  7-10.  Now,  assume 
that  the  liquid  flowing  down  the  tower  at  some  position  contains  1.3 
mols  of  CaHg.  The  vapor  at  this  point  must  then  contain  1.25  mols  of 
C3H8,  giving  xn+i  =  1.3/101.3  =  0.0128  and  yn  -  1.25/5.25  «  0.238. 


RECTIFICATION  OF  BINARY  MIXTURES 


137 


tVn 


fVs 


In  a  similar  manner,  other  values  on  the  operating  line  are  calculated 
and  plotted  in  Fig,  7-11  together  with  the  equilibrium  curve,  and  a 
little  more  than  six  steps  are  required  to  give  the  desired  stripping. 

Side  Streams.  Side  streams  are  removed  from  a  column  most 
often  in  multicomponent  mixtures; 
however,  they  are  occasionally  used  in 
the  distillation  of  binary  mixtures. 
Thus,  a  plant  separating  alcohol  and 
water  might  have  uses  for  both  80  and 
95  per  cent  alcohol  mixtures,  which  of 
course  could  be  produced  by  making 
only  95  per  cent  alcohol  and  diluting 
with  water  to  produce  the  required  80 
per  cent;  or  alternately  liquid  could  be 
tapped  off  of  a  plate  in  the  column  on 
which  the  concentration  was  approxi- 
mately 80  per  cent.  The  proper  plate 
in  the  tower  can  be  determined  by  con- 
structing the  usual  operating  lines. 
Figure  7-12  illustrates  the  removal  of 
a  liquid  side  stream  L.  Considering 
this  figure  and  making  the  usual  sim- 
plifying assumptions,  the  operating 
line  above  the  side  stream  is,  as  before, 


FIG.  7-12.     Diagram  of  continuous 
column  with  side  stream. 


DxD 


A  material  balance  around  the  top  of  the  column  and  some  plate 
between  the  feed  plate  and  the  side-stream  plate  gives 


W.  *• 


Lxi 


V. 


(7-12) 


The  operating  line  above  the  side  stream  passes  through  y  =  x  =  XD 
and  has  a  slope  of  (0/F)«,  while  the  operating  line  below  the  side 

stream  passes  through  y  =  x  =    x*       J*^  (i.e.,  the  molal  average 

composition  of  the  product  and  side  stream)  and  has  a  slope  (0/V)8. 
Since  XL  is  less  than  XD  and  0,  is  less  than  On,  (0«  =  On  -  L);  this 
latter  operating  line  will  cross  the  y  =  x  diagonal  at  a  lower  value  than 


138 


FRACTIONAL  DISTILLATION 


the  upper  operating  line  and  will  have  a  flatter  slope.  The  two 
operating  lines  will  intersect  at  x  =  XL.  Figure  7*13  illustrates  these 
lines. 

Theoretical  plates  are  stepped  off  in  the  usual  manner,  using  the 
operating  line  ac  from  a  to  c,  the  operating  line  bcf  from  c  to 
some  value  between  e  and  /,  and  then  the  line  edg  from  there  to  g. 
Although  the  feed  plate  may  have  any  composition  between  e  and  / 
and  the  feed-plate  step  may  be  made  at  any  value  between  these  two, 
the  side-stream  step  must  fall  exactly  on  c.  This  is  because  the  feed 


XW 


FIG.  7-13.     y,x  diagram  for  column  with  side  stream. 

can  be  actually  introduced  into  plates  of  different  composition,  but  a 
side  stream  has  to  be  of  the  same  composition  as  the  plate  from  which 
it  was  withdrawn  unless  a  partial  separation  of  the  side  stream  is  made 
and  a  portion  returned  to  the  column.  By  altering  the  reflux  ratio 
slightly,  the  step  can  be  made  to  fall  very  close  to  c. 

Unequal  Molal  Overflow.  1.  Above  Feed  Plate.  The  analysis 
given  in  the  preceding  sections  was  based  on  constant  molal  overflow 
rate,  and  it  is  necessary  to  consider  the  validity  of  this  assumption. 

Equations  (7-1)  and  (7-2)  can  be  combined  with  the  following 
enthalpy  balance: 


V«Hn 


+  On+ihn+i  +  DhD  +  losses 


RECTIFICATION  OF  BINARY  MIXTURES  139 

where  Hn  »  molal  enthalpy  of  vapor  entering  plate  n  +  1 

Q0  =  heat  removed  in  condenser 
hn+i  =  molal  enthalpy  of  liquid  leaving  plate  n  +  1 
hD  «  molal  enthalpy  of  product 

to  give 

Xp    ~-  Xn+l     _     _Xj>   ""  J/n  /7   jq% 

~  MD  -  Hn  ('~L6) 


where  MD  equals  -j?  +  ftx>  +  .  °®fes  or,  in  general,  DMD  equals  the 

total  enthalpy  removed  from  the  section  in  question  other  than  by 
Vn  and  On+i.     Equation  (7-13)  can  be  rearranged  as 

M  D    —  H  n  i      *fn          kn+1    M  /"7   i  A\ 


n 
/I  ^n    —  hn+i\ 

-.  i  i  _     -  —  i  ^ 

\  MD    "^  fan+l/ 


--  -_  - 

MD   ~ 

By  comparison  with  Eq.  (7-2)  it  is  obvious  that 


fr~    —    THF  T 

Kn          MD   —  Aln-fl 

and 

0»-.l  Af/)    —   ffn  Hn    — 


Vn          MD  -  An+i  MD  - 

The  condition  for  constant  molal  overflow  rate  is  that  the  group, 

tin  —  hn+i  should  be  a  constant.  In  general  the  losses  are  or  should 
M  D  —  An+i 

be  small  making  MD  a  constant  for  a  section  having  no  additions  or 
withdrawals  except  at  its  ends,  and  the  value  of  MD  is  usually  large  in 
comparison  to  hn+i.  Thus,  a  constant  value  of  Hn  —  hn+i  will  lead  to 
essentially  constant  molal  rates  of  vapor  and  overflow.  The  difference 
Hn  —  An+i  is  not  a  conventional  latent  heat  but  is  the  difference  in 
molal  enthalpy  between  the  vapor  entering  and  the  liquid  leaving  a 
plate.  In  order  to  analyze  this  difference,  it  is  desirable  to  evaluate 
Hn  and  hn+i* 

In  the  case  of  .the  liquid  phase,  the  enthalpy  will  be  calculated  on 
the  basis  of  heating  the  pure  components  to  the  mixture  temperature 
and  then  mixing  at  this  condition. 


140  FRACTIONAL  DISTILLATION 

where  (xi)n+i,  (a^n+i  =  niol  fractions  of  components  1  and  2  in  liquid 

leaving  (n  +  l)th  plate 
hn+i  =  molal  enthalpy  of  liquid 
ci,  c2  =  molal  specific  heat 

tn+i  =  temperature  of  liquid  leaving  (n  +  l)th  plate 

fa)  fa  =  base  temperatures  for  calculating  enthalpy, 

enthalpy  of  pure  component  1  taken  as  zero 

at  fa,  enthalpy  of  pure  component  2  taken  as 

zero  at  £2& 

Ahm  =  enthalpy  change  on  mixing  pure  liquids  to 

give  1  mol  of  desired  composition 

In  most  cases  it  is  desirable  to  choose  the  base  temperatures  reasona- 
bly close  to  the  distillation  temperatures.  For  such  cases  Ci  and  c2 
can  be  taken  at  constant  average  values  and 

fen+l    =    (3l)n+l(Cl)(*n+l    ~   fa)    +    (Za)»+l(Cs)(*n+l    —   fa)    +  Akm        (7-19) 

In  the  case  of  the  vapor  enthalpy,  several  paths  for  calculating  the 
mixture  value  relative  to  the  pure  components  are  possible:  (1)  The 
pure  components  can  be  heated  as  liquids  from  the  base  temperature 
to  their  boiling  point  at  the  pressure  under  consideration,  (2)  the 
liquids  can  be  vaporized,  and  (3)  the  vapors  can  be  mixed  and  then 
heated  or  cooled  to  the  desired  mixing  temperature.  Alternately,  the 
pure  components  can  be  (1)  heated  from  the  base  temperature  to  the 
desired  mixture  temperature,  (2)  vaporized  at  this  temperature,  and 
(3)  mixed.  Thus,  for  the  two  cases, 

Hn    =    (yi)n     ftlB  Cl  dt   +    (y2)n    ('**  C2  dt   +    foi) 
Jtlb  Jtti, 

+  AHm  +  (yi)n  f^  Ci  dt  +  (y,)n        C2  dt     (7-20) 
or 

ci  dt  +  (y2)n  ^  c2  dt  +  (yi)n  AH[ 

(7-21) 
where  Hn  =  molal  enthalpy  of  vapor 

2/i,  ^2  ^  niol  fractions 
i,  c2,  fi6,  fa  =  same  as  for  Eq.  (7-18) 

tin,  UB  =  boiling  temperatures  for  pure  components  at  pressure 

in  question 

tn  =  temperature  of  vapor  entering  plate  n  *+  1 
i,  AJEf2  =  latent  heats  of  vaporization  of  the  pure  components  at 

IB  and  t%B 

i,  &H'Z  =«  latent  heats  of  vaporization  of  the  pure  components  at  tn 
AHm,  AH'm  =  enthalpy  changes  on  mixing  pure  vapors 
Ci,  C2  «  molal  heat  capacities  of  the  pure  vapors 


RECTIFICATION  OF  BINARY  MIXTURES  14 

For  most  of  the  following  discussion  the  second  equation  will  b< 
employed,  and  the  heat  capacities  will  be  taken  constant  at  averag< 
values,  giving 

Hn    =    (yi)nCl(tn    —   tib)    +   (y^nC^(tn    ~   *2&) 

+  (Vl)n  AH(  +  (y2)n  AH't  +  AH'm     (7-22 

At  low  and  moderate  pressures,  the  enthalpy  effects  on  mixini 
vapors  are  small  and  AJET^  and  AHm  will  be  neglected,  although  at  higi 
pressures  this  procedure  could  lead  to  large  errors. 

Combining  Eqs.  (7-19)  and  (7-21), 

Hn  -  hn+i  =  (yi)n  &H{  +  (y,) 


(7-23 

The  values  of  tw  and  £2&  can  be  arbitrarily  chosen.  For  convenient 
they  will  be  taken  as  the  boiling  points  of  the  pure  components  at  th< 
pressure  in  question.  On  this  basis,  (tn+i  —  tib)  and  (tn+i  —  £2& 
seldom  exceed  50  to  100°F.  and  (yn  ~  xn+i)  and  (tn  —  tn+i)  are  small 
As  a  result,  the  last  two  brackets  of  Eq.  (7-23)  are  usually  only  a  fev 
per  cent  of  the  value  of  (Hn  —  hn+i)  and  are  not  of  real  significance  ii 
determining  the  constancy  of  this  difference.  The  most  importan 
factors  are  (yi)n  &H(  and  (y2)n  AH'z,  although  Ahm  may  be  large  ii 
some  cases.  As  the  calculations  are  made  from  plate  to  plate,  (yi) 
will  vary,  and  (Hn  —  7&n+i)  will  vary  or  remain  constant  depending  01 
whether  or  not  AH(  equals  A-H^.  Thus  for  most  cases,  the  criterion  o 
the  constancy  of  (Hn  —  An+i)  will  be  the  difference  in  the  latent  hea 
of  vaporization  of  the  pure  components  at  the  operating  pressure. 

Figure  7-14  gives  the  values  of  the  molal  latent  heat  divided  by  tin 
absolute  temperature  plotted  as  a  function  of  the  ratio  of  the  vapo 
pressure  divided  by  the  absolute  temperature.  (See  also  page  479. 
As  a  rule,  the  maximum  variation  in  (yi  AH(  +  y2  AH'2)  is  obtained  ty 
going  from  yi  =  1.0  to  yi  =  0,  although  abnormal  mixtures  may  shov 
greater  values  at  some  intermediate  point.  As  an  example,  con 
SJder  a  mixture  of  ethanol  and  water  at  atmospheric  pressure.  Fo 
ethanol  P/T  =  760/351.5  =  2.16,  AH/T  =  27,  AH'  =  9,470;  fo 
water  P/T  -  760/373  -  2.04,  AH/T  -  26.2,A  H'  =  9,750.  Unles 
this  mixture  has  a  very  large  heat  of  mixing  for  the  liquid  phase,  i 
would  be  expected  that  the  value  of  (Hn  —  hn+i)  would  be  essentiall; 
constant  and  the  assumption  of  constant  rates  of  vapor  and  over 
flow  would  be  well  justified.  At  higher  pressures  the  variation  i 


142 


FRACTIONAL  DISTILLATION 


of  vaporization  of  ethanol.     As  another  case,  consider  the  distillation 
of  an  ammonia-water  mixture  at  20  atm.  abs.     For  ammonia 


P 
T 


(20  X  760) 


321.5 
AH/T  =  13.5,  and  AH'  =  4,340;  for  water 

P 
T 


47.3 


(20  X  760)  _ 
484.5       "" 


AH/T  =  17.0,  and  AJET  =  8,230,  and  the  variation  in  vapor  and  over- 
flow rates  from  plate  to  plate  would  be  large  and  calculations  based  on 
the  constancy  of  these  rates  would  be  appreciably  in  error. 


44 
40 
36 
32 
28 

20 
16 
12 
8 

4 

0 
01 

I 

III 

1 

IIMI       1    1 

Nomenclature 
1  -Molecular  wejaht 
-  Latent  heat,(5m,  Cal/6m 
"  -  Absolute  temperatur*,°K. 
"Vapor  pressure  Mm.  Hg. 

^*x 

4 

* 

^s, 

r 

i!ll^> 

^X 

*v 

•4 

P 

M 

S|6 

„  . 

1 

^s 

1  *  •  '^x 

V 

Sv.     ' 

^< 

hS 

^s 

A             ^5 

V 

»C 

N 

'^v 

s 

s 

J     ' 

s 

N 

s 

sr* 

*•  lfev«- 

N 

>, 

vxS> 

V 

S  J  J 

^  ^O 

vs. 

s|'v 

"^N^ 

Leger 
-Wbter 
!-EmylA 
i-EthylEI 
-Ammo 
i-  Ethane 
i-Benzer 
7-n-Octa 

e! 

cohol 
tier 
nia 

le 
nil 

'v^s 

w 

s 

\^ 

s 

'    V 

i 

^ 

»    ^L 

S    ^W 

A 

\  ^ 

i     ^ 

! 

1    > 

\  V\ 

\ 

_7  1  ^ 

\    } 

\ 

1 

\ 

\ 

)l                     0.1 

1.0              to              no             ijooo 

P/T 
FIG.  7-14.     Hildebrand  chart  for  estimating  latent  heats  of  vaporization.     (Ref.  19.) 

It  is  possible  to  have  mixtures  of  similar  components  that  would 
not  satisfy  the  constant  0/V  assumption  even  approximately  and  to 
have  mixtures  of  unlike  components  that  give  good  agreement.  Thus, 
in  a  mixture  of  hydrocarbons  the  operating  pressure  could  be  such  that 
one  of  the  components  is  near  its  critical  pressure  as  a  pure  com- 
ponent while  the  other  could  be  at  more  normal  conditions.  Under 
such  conditions,  the  difference  in  latent  heats  of  vaporization  of  the 
two  pure  components  could  be  very  large.  At  high  pressures,  most 
mixtures  give  large  variation  in  (H n  —  /&w+i)  because  one  of  the  com- 
ponents will  be  nearer  its  critical  conditions  than  the  other. 

In  order  to  carry  out  more  exact  calculations,  it  is  necessary  to 


RECTIFICATION  OF  BINARY  MIXTURES  143 

evaluate  the  other  terms  of  Eq.  (7-23).  The  specific  heat"  terms  require 
the  heat  capacity  of  the  pure  liquid  components.  Values  for  a  number 
of  materials  can  be  obtained  from  Fig.  2  in  the  Appendix.  (Ref.  10.) 
The  value  of  Ahm  is  difficult  to  obtain,  owing  to  the  lack  of  published 
data.  Such  data  lire  available  in  a  few  cases  and  can  be  calculated 
if  vapor-liquid  equilibrium  data  are  available  at  several  temperatures 
and  pressures.  Thus,  by  Eqs.  (3-37)  and  (3-38), 

RTd  In  71  = 
and 


»  RT  In  7l 
=  ft?7  In  T2 

It  can  be  shown  that,  at  constant  composition  and  total  pressure, 

d(*P./T)  _     -   _  _ 
d(l/T)    ~~    ^e  " 
and 


(7-24) 

Thus  if  vapor-liquid  equilibrium  data  are  available  to  evaluate  the 
activity  coefficients  at  constant  composition  as  a  function  of  the  tem- 
perature, the  heat  of  mixing  can  be  calculated.  This  heat  of  mixing, 
so  calculated,  will  be  slightly  in  error  due  to  neglecting  the  effect  of 
pressure  on  the  enthalpy  of  liquids. 

It  is  to  be  noted  that  Ahm  is  usually  of  such  a  sign  as  to  make  the 
enthalpy  of  the  liquid  more  nearly  constant.  In  using  Eq.  (7-19) 
with  the  base  temperatures  taken  at  the  boiling  point  for  each  pure 
component,  the  enthalpy  of  the  liquid  is  equal  to  zero  at  both  x\  =  1 
and  Xi  =  0,  but  at  intermediate  values  the  enthalpy  may  be  greater  or 
less  than  zero.  For  mixtures  with  positive  deviations  from  Raoult's 
law,  i.e.,  activity  coefficients  greater  than  unity,  the  temperatures  of 
the  mixtures  at  constant  pressure  are  lower  than  for  ideal  solutions, 
and  the  sum  of  the  sensible  heat  terms  of  Eq.  (7-23)  is  negative. 
This  is  partly  offset  by  the  fact  that  (d  In  y/dT)  is  usually  negative 
for  such  mixtures,  making  Ahm  positive.  Mixtures  with  negative 
deviations  from  Raoult's  law  give  similar  compensations.  Thus,  in 
most  cases  the  variation  in  the  enthalpy  of  the  liquid  (with  the  base 
temperature  assumed)  with  composition  is  small. 

If  the  Van  Laar  equation  for  the  activity  coefficient  is  used,  Eq. 


144  FRACTIONAL  DISTILLATION 

(7-24)  becomes 

Ahm  »  RT(xl  In  71  +  z2  In  72)  (7-25) 

and,  using  the  Van  Laar  relations, 

Ahm  =    **f*^  (7-26) 

Comparison  with  experimental  data  for  a  number  of  mixtures  of 
organic  liquids  indicates  that  the  Van  Laar  relations  give  high  values 
and  that  better  results  are  obtained  by 

Ahm  =  Q.5RT(x!  In  7!  +  x2  In  72) 

The  use  of  the  modified  Margules  relations,  Eq.  (3-34a),  would  give  a 
similar  relation  with  the  constant  equal  to  0.25. 
2.  Below  Feed  Plate.     A  similar  analysis  using 

Om+l    -    Vm   +   W 

Om+1xm+1  =  Vmym  +  Wxw 
Om+ihm+i  +  Q*  =  VmHm  +  Whw 

where  Q8  is  the  heat  added  in  the  still,  gives 

ym  -  xw         _         xm+1  -  xw  (?-27) 


Hm  -  \hw  -  p~ 


or 

9I    -   Mw  —  Hm  Hm  — 

ym  — 


~  -  -  -    m+i        XT  -  z^  -  •wr 

Mw    —  /Im-fl  MTT   —  km+1 

-    A     -    gm    "~   ^+A  r  J_  -^m    -   Am+1 

—    II—    J-f  -  jr  -  I  Xm+l   "f  TTf  -  1  -  Xw 

—  — 


where  Mw  *  hw  —  TF?> 


^r 


-  hm+l 


These  equations  are  similar  to  Eqs.  (7-16)  and  (7-17),  and  the  same 
considerations  relative  to  the  constancy  of  molal  rates  of  vapor  and 
overflow  apply  in  this  case, 

3.  General  Case.  The  derivations  given  in  the  two  previous  sections 
were  for  the  case  with  no  side  streams,  but  the  general  case  is  similar. 


RECTIFICATION  OF  BINARY  MIXTURES 


145 


In  the  system  shown  in  Fig.  7-15,  the  over-all  and  component  material 
balances  above  plate  a  are 

Fa  =  Oa+i  +  N  (7-32) 

F«2/a  =  Oa+iXa+i  +  NxN  (7-33) 

where  Fa  =  vapor  rate  to  plate 


^u 


~D 
~Li 


Oa+i  —  overflow    rate    from 

plate  a  +  1 

N  =  net  mols  leaving  sec- 
tion =   Fa  —  Oa+l 

2/a>  #a-f i  —  mol  fractions 

NXN  =  total  mols  of  compo- 
nent leaving  section 
other  than  with  Fa 
and  00+i 

The  enthalpy  balances  are 

VaHa  =  Oa+iha+1  +  NM     (7-34) 

where  Ha  =  molal  enthalpy  of  va- 
por   entering    plate 


ha+i  =  molal      enthalpy      of       FIG.   7-15.     Schematic  diagram  of  frac- 
liduid     leaving!    plate       tionating  system  with  side  streams. 

a  +  1 

NM  =  net  enthalpy  removed  from  the  column  above  plate  a  +  1 
For  Fig.  7-15, 

NhN  =  Li(hL)i  +  £2(^)2  +  DhD  —  FhF  +  Qc  +  losses 
These  equations  can  be  combined  to  give 


or 


If  -  Ha 

M   —  Ha 


M  - 


M  - 
=  I  1  - 


, 
~\ 


where 


M  -  h< 

Hg     —    hg+1 

M  -  ha+1 

1 ~% r^ — 


M  -  h 

Xa-4-1     i 


-  ha. 


'•XN 


.+1 


V 


(7-35) 

(7-36) 

(7-37) 
(7-38) 
(7-39) 


146 


FRACTIONAL  DISTILLATION 


These  equations  reduce  to  Eqs.  (7-14),  (7-17),  (7-27),  and  (7-31)  for 
the  special  cases  considered. 

Method  of  Ponchon  and  Savarit.  Ponchon  (Ref.  14)  and  Savarit 
(Ref.  15)  showed  that  Eqs.  (7-13)  and  (7-27),  or  in  general  any  equation 


0      0.1     0.2     03     0.4    05    0.6     07     0.8    0.9     1.0 

y  or  x ,  Mol  fraction 
FiQ,  7-16,     Enthalpy-composition  diagram. 

of  this  type,  could  be  easily  solved  by  plotting  the  enthalpy  (or  other 
property)  of  the  saturated  vapor  and  liquid  vs.  the  mol  fraction.  For 
example,  in  Fig.  7-16,  if  the  value  M D  for  the  upper  section  is  plotted 
at  XD,  it  is  easily  shown  that  any  straight  line  drawn  through  the 
point  (M  D,  XD)  will  intersect  the  enthalpy  lines  to  give  values  of  x  and 


RECTIFICATION  OF  BINARY  MIXTURES  147 

y  that  will  satisfy  Eq.  (7-13).  Thus  if  yn  is  known,  a  line  through  the 
vapor  enthalpy  curve  at  this  composition  and  the  point  (M»,  XD)  will 
cut  the  liquid  enthalpy  at  hn+i  and  xn+i.  When  xn+i  is  known,  yn+i  is 
obtained  from  the  equilibrium  curve  and  xn+$  is  determined  by  drawing 
a  new  line  through  (yn+i,  Jffn+i),  (MD,  xx>),  etc. 

Similarly  if  Mw  for  the  lower  section  is  plotted  at  Xw,  then  a  straight 
line  through  this  point  intersects  the  two  enthalpy  curves  at  values 
that  satisfy  Eq.  (7-27)  and  these  straight  lines  give  the  relation 
between  ym  and  xm+i. 

A  heat  and  material  balance  around  the  whole  column  (no  side 
streams)  gives 

Fhr  =  Whw  +  DhD  +  QC-Q,  +  losses  =  WMW  +  DMD 
and 

Fzp  =  Wxw  +  DxD 


By  rearranging, 

ZF  —  Xw 


hF  ~  Mw 


(7-40) 


Equation  (7-40)  is  of  the  same  type  as  Eqs.  (7-13)  and  (7-27),  and 
similar  reasoning  leads  to  the  conclusion  that  a  straight  line  through 
(ZP,  hp)  and  (xw,  Mw)  will  also  pass  through  (XD,  M&).  In  other 
words,  the  point  (ZF,  hp)  lies  on  a  straight  line  between  (XD,  MD)  and 
(xw,  Mw).  This  line  will  be  termed  the  terminal  tie  line. 

In  general  the  same  type  of  information  given  by  the  constant  0/V 
method  can  be  obtained  by  the  use  of  the  Ponchon  and  Savarit 
method.  For  example,  the  cases  of  total  reflux,  minimum  reflux 
ratio,  and  optimum  feed-plate  location  can  be  easily  solved. 

Total  Reflux.  In  the  case  of  total  reflux  the  values  of  M D  and  M w 
are  infinite,  and  lines  drawn  through  them  and  the  enthalpy  curves  will 
therefore  be  vertical.  Thus,  for  this  case  the  diagram  shows  that  the 
composition  of  the  liquid  leaving  the  plate  is  equal  to  the  composition 
of  the  vapor  entering  the  plate,  and  the  same  number  of  theoretical 
plates  will  be  obtained  by  both  the  constant  0/V  method  and  Ponchon- 
Savarit  methods,  regardless  of  the  value  of  the  enthalpies. 

Minimum  Reflux  Ratio.  The  case  of  the  minimum  reflux  ratio 
corresponds  to  conditions  that  require  an  infinite  number  of  plates  to 
obtain  the  desired  separation.  As  in  the  case  of  the  y,x  diagram,  this 
necessitates  a  region  in  which  succeeding  plates  differ  only  differen- 
tially in  composition,  i.e.,  a  pinched-in  region. 

The  step  equivalent  to  a  theoretical  plate  on  the  enthalpy  diagram 


148  FRACTIONAL  DISTILLATION 

involves  going  from  the  vapor  below  a  plate  to  the  liquid  on  a  plate  by 
means  of  the  enthalpy  operating  line  through  one  of  the  terminal 
enthalpy  points  and  then  proceeding  from  the  composition  of  the 
liquid  on  the  plate  to  the  vapor  above  the  plate  by  the  equilibrium 
relationship,  i.e.,  by  an  equilibrium  tie  line.  If  the  composition  of 
this  vapor  above  the  plate  is  to  be  equal  to  the  composition  of  the 
vapor  entering  the  plate,  it  is  necessary  for  the  enthalpy  operating 
line  to  coincide  with  the  equilibrium  tie  line.  In  the  general  case  it 
is  a  trial-and-error  procedure  to  determine  the  least  value  of  MD  that 
can  be  employed.  However,  if  the  pinched-in  region  occurs  at  the 
feed  plate,  the  minimum  reflux  ratio  can  be  easily  determined  by  find- 
ing the  equilibrium  tie  line  that  passes  through  the  point  (hp,  ZF)  and 
extrapolating  this  line  until  it  intersects  the  vertical  line  at  x&.  The 
enthalpy  value  at  this  intersection  will  correspond  to  the  minimum 
M &.  In  other  cases,  the  equilibrium  tie  lines  for  a  number  of  compo- 
sitions above  the  feed  plate  can  be  extrapolated  to  the  vertical  line 
through  XD,  and  the  maximum  value  of  MD  so  obtained  corresponds  to 
the  minimum  reflux  ratio  for  this  section.  A  similar  procedure  can  be 
used  below  the  feed  plate  and  the  minimum  value  of  Mw  determined. 
Since  M />  and  M w  must  fall  on  the  line  through  the  feed  point,  it  can 
be  determined  which  of  the  two  values  is  the  limiting  one,  and  thus 
whether  the  pinched-in  region  is  above  or  below  the  feed  plate. 

Optimum  Feed-plate  Location.  The  optimum  feed-plate  location 
again  corresponds  to  making  the  total  number  of  theoretical  plates 
required  for  the  operating  conditions  chosen  a  minimum,  which  is 
equivalent  to  making  the  change  in  composition  per  plate  a  maximum 
at  all  points.  In  general,  it  will  be  found  that,  when  the  enthalpy 
operating  line  is  on  the  XD  side  of  the  terminal  tie  line  through  (MD,  XD) 
and  (Mw,  xw),  larger  steps  will  be  obtained  by  using  enthalpy 
operating  lines  drawn  through  (MD,  XD)  than  those  drawn  through 
(Mw,  xw)9,  on  the  xw  side  of  the  terminal  tie  line  the  reverse  will  be 
true.  The  enthalpy  operating  lines  should  be  drawn  through  Mw  for 
values  of  x  less  than  the  composition  given  by  the  intersection  of  the 
terminal  tie  line  with  the  liquid  enthalpy  curve,*  and  the  (M&,  XD) 
point  should  be  used  for  all  operating  lines  corresponding  to  liquid 
mol  fractions  greater  than  this  value.  It  should  be  noted  that  the 
value  of  the  mol  fraction  at  which  the  change-over  is  made  is  not  equal 
to  the  composition  of  the  feed.  It  will  be  equal  to  the  composition  of 
the  feed  when  the  feed  enters  with  an  enthalpy  equal  to  that  of  a 
saturated  liquid.  If  the  enthalpy  of  the  feed  is  greater  than  that  of  a 
saturated  liquid,  the  change-over  value  will  be  at  composition  lower 


RECTIFICATION  OF  BINARY  MIXTURES  149 

than  that  of  the  feed.  If  the  enthalpy  of  the  feed  is  less  than  that  of 
the  saturated  liquid,  the  reverse  will  be  true.  This  is  similar  to  the 
conditions  found  for  the  y,x  diagram. 

General.  It  is  frequently  advantageous  to  use  enthalpy  diagrams 
to  determine  a  series  of  values  of  yn,  xn+i,  and  ym,  xm+i,  which  can  be 
plotted  on  the  y,x  diagram  to  give  the  actual  operating  lines  which  are 
then  employed  in  the  usuaLstepwise  manner.  The  y,x  curves  and  the 
enthalpy  composition  values  can  be  plotted  on  the  same  diagram, 
and  the  combined  graphical  procedure  shown  in  Fig.  7-16  completes 
both  the  y,x  and  the  enthalpy  diagrams.  To  illustrate  the  procedure, 
(1)  starting  at  (hw  ,  xw),  a  vertical  line  is  drawn  to  the  equilibrium  curve 
giving  yw,  (2)  this  value  is  transposed  to  the  H,y  line  by  going  horizon- 
tally to  y  =  x  and  then  vertically  to  the  //  curve  giving  the  point  (Hw, 
yw),  (3)  the  line  through  this  point  and  (Mw,  xw)  gives  the  value  of  xi  on 
the  h  curve,  and  (4)  the  process  is  repeated.  The  intersection  of  the 
vertical  lines  with  the  horizontal  lines  on  the  y,x  diagram  corresponds 
to  points  of  the  operating  lines,  and  the  triangles  above  the  lines  drawn 
through  these  points  are  the  usual  steps  on  the  y,x  diagram. 

Heat  losses  from  the  column  can  be  taken  into  account  by  shifting 
the  value  of  MD  from  plate  to  plate  by  an  amount  equal  to  the  heat 
loss  per  plate  divided  by  N.  A  side  stream  in  the  upper  section  of  the 
column  is  handled  by  locating  a  point  (Ms,  XN$)  where 

,  ,        LhL  +  DhD  +  Qc  ,  LxL  +  DxD 

Ms  -  -  —-  -  -        and        XM  -  --- 


and  drawing  lines  through  this  point. 

In  this  case,  (ZF,  hr)  lies  on  a  straight  line  through  (xw,  Mw)  and 
(XNS,  Ms),  and  (XNS,  Ms)  lies  on  a  straight  line  through  (XD,  Mn)  and 
(XL,  hL).  In  a  similar  way  any  number  of  side  streams  or  feeds  can  be 
handled. 

In  general  the  Ponchon-Savarit  diagram  is  somewhat  more  difficult 
to  use  than  the  constant  0/V  diagram,  but  it  is  the  exact  solution  for 
theoretical  plates  assuming  that  the  enthalpy  data  employed  are 
correct.  This  graphical  procedure  suffers  because  the  absolute  values 
of  M  D  and  M  w  are  frequently  large,  and  to  plot  them  on  the  diagram 
.requires  the  use  of  an  ordinate  scale  such  that  the  enthalpy  curves  for 
the  liquid  and  vapor  are  crowded  together,  making  it  difficult  to 
obtain  accurate  results.  Likewise  when  low  concentrations  are 
encountered,  it  is  impossible  to  obtain  accurate  results  from  the  dia- 
gram unless  the  plot  is  greatly  expanded.  For  such  regions  Eqs. 
(7-14)  and  (7-28)  can  be  used  algebraically,  or  in  most  cases  the  con- 


150 


FRACTIONAL  DISTILLATION 


t«150C,H«3°l30 

.P- 

* 

<l4J5mols99.9%Oz,| 
t'15°C  H33IIO 

3= 

{ 

"100  mols  air 

100  mols 

t_  OAO/"> 

p*l 

*  i\J  L 

p=ZOdtmos 

c 

o 

ol 

o 

.—> 

h=330 

H*3I20 

85.85  mols 

t--!85°C. 
94  3%  Liquid 

Refrigeration 

o  < 

c 

i 

92%  N2 
p=1 

5.7%  Vapor 

< 

4 

< 

a 
1 

t=H94°C. 
H  =  I630 

Heat  leakage  in 

< 
4 

o 

c 

| 

t> 

H> 

< 
4 

1 

100  mols 

1 

a 

Air 

+•-  —  IIA0^ 

< 

? 

r 

14  15  mols  99  9%  02 

^ 

h*330 

p*l  t=-!8295°C. 

N< 

X 

/ 

\ 

u  «.  mA 

'  /« 

^ 

L£~t 

^ 

!*SA 

100  mols  air 
p=20atmos 
H  =  I645 
FIG.  7-17.    Liquid-air  distillation. 

stant  0/V  diagram  is  useful  since  in  such  regions  the  value  of  (Ha  —  ha+i) 
is  essentially  constant. 

EXAMPLES  OF  ENTHALPY-COMPOSITION  METHOD 

Liquid-Air  Fractionation  Example.  The  enthalpy-composition  method  will  be 
illustrated  by  several  examples.  Figure  7-17  gives  a  schematic  diagram  of  a  liquid- 
air  fractionating  system.  In  this  particular  case  the  system  employed  involves 

TABLE  7-2 


Xn+i 

2/ncalo 

Constant  0/V 

Heat  balance 

0.001 

0.001 

0.002 

0.00218 

0.004 

0.0045 

0.01 

0.0116 

0.02 

0.0233 

0  024 

0.04 

0.0468 

0  048 

0.1 

0.1175 

0.12 

0.2 

0.235 

0.24 

0.4 

0.470 

0.475 

0.6 

0.705 

0.71 

0.8 

0.94 

0.94 

RECTIFICATION  OF  BINARY  MIXTURES 


151 


2000 


only  a  stripping  column,  and  the  enthalpy  diagram  (Fig,  7-18)  is  only  for  the  lower 

portion.    The  heat  supply  to  the  still 

is  from  the  ingoing  air  and,  on  the 

basis  given  for  Fig.  7-17,  Q,  is  equal 

to  100(1,645  -  330)  -  131,500.     The 

molal  enthalpy  of  the  bottoms  from 

the  still,  Hw,  is  1,730,  and  for  Eq. 

(7-27)  the  terminal  enthalpy  point 

becomes 


Bw--  -7,550 

This  construction  is  given  in  Fig.  7-18. 

The   enthalpy  values  employed  are 

those  given  by  Keesom  (Ref  .  7)  and, 

since  they  were  already  available,  they 

were  not  recalculated  to  the  basis  sug- 

gested in  the  foregoing  section.     The 

diagram  of  Fig.  7-18  is  not  very  suit- 

able for  a  major  portion  of  the  con- 

centration region  under  consideration. 

This  diagram  and  Eq.  (7-28)  were  used 

to  calculate  y  and  x  values  for  the  op- 

erating line.    The  results  of  these  cal- 

culations are  compared  with  those  for 

constant  0/V  in  Table  7-2.     It  will  be 

noted  that  within  the  accuracy  of  the 

calculation  the  operating  line  coordi- 

nates calculated  by  the  two  methods  are  in  agreement,  indicating  that  the  constant 

0/V  method  is  satisfactory  for  this  case. 

The  operating  line  values  and  the  equilibrium  data  of  Dodge  (Ref.  4)  are  shown 
in  Fig.  7-19.  Logarithmic  plotting  is  employed  so  that  steps  can  be  satisfactorily 
made  in  the  low  concentration  region.  Nine  theoretical  plates  in  addition  to  the 
still  give  the  desired  separation. 

TABLE  7-3.     02-NZ  AT  10  ATM. 


04         06 
x  ory 


0.8 


1.0 


FIG.    7-18. 
separation. 


Heat    diagram   for  liquid-air 


2/noalo 

X»+l 

Heat  balance, 

Constant  0/V 

data  of  Bosn- 

jakovic  (Ref.  2) 

1.0 

1.0 

1.0 

0.9 

0.955 

0.953 

0.8 

0.91 

0.91 

0.7 

0.865 

0.867 

0.6 

0.82 

0.823 

152 


FRACTIONAL  DISTILLATION 


1.0 
0.8 

0.6 

0.4 

0.2 

01 
t   0.08 

R  0.06 

0 

.£   0.04 

M 

•j:   0.02 
o 

£ 

o  0.01 
2  0.008 

aooe 

0002 

• 

Ws 

^l"^H 

'/* 

^^ 

// 

^^x  \/ 

X 

x   ^ 

/ 

X 

X 

Vs 

^ 

^ 

S 

^ 

'* 

^ 

.^y 

/ 

y 

.vVy^ 

/ 

Y 

/ 

Y 

Sf 

Vj 

\4^ 

'/ 

/ 

' 

$ 

^ 

A 

/ 

V 

/: 

f» 

/ 

/, 

s 

/ 

/ 

/ 

x 

'/ 

/ 

^ 

/ 

0,001 

/ 

0.001       0.002        00040.006    001          0.02         0.04    0.06      01           02           04     0.6       1.0 
Mol  Fraction  N2  in  Liquid 

FIG.  7-19.     x,y  diagram  for  liquid-air  separation. 

TABLE  7-4.     NH3-H2O  AT  10  ATM. 


2/n  oalc 

Xn+l 

Heat  balance, 

Constant  0/V 

data  of  Bosn- 

jakovic  (Ref.  2) 

0.1 

0.701 

0  705 

0.09 

0.624 

0.625 

0.08 

0  547 

0.547 

0.07 

0.47 

0.46 

0.06 

0.393 

0.385 

0.05 

0.316 

0.305 

RECTIFICATION  OF  BINARY  MIXTURES 


153 


Oxygen-Nitrogen  and  Ammonia-Water  Example.  Tables  7-3  and  7-4  give  similar 
comparisons  for  other  systems.  In  Table  7-3  the  comparison  is  for  the  oxygen- 
nitrogen  system  at  10  atm.,  and  again  the  agreement  between  the  two  methods  is 
good.  The  other  data  are  for  ammonia-water  at  10  atm.,  and  even  in  this  case  the 
two  methods  are  in  fair  agreement. 

Ammonia-Water  Example.  In  order  to  illustrate  the  fact  that  the  assumption  of 
constant  molal  overflow  rate  is  not  always  justified,  consider  the  following  example 
in  which  the  molal  latent  heat  of  vaporization  varies  approximately  twofold  over 
the  tower.  An  aqueous  solution  containing  20  weight  per  cent  ammonia  is  to  be 
separated  into  a  distillate  containing  98  mol  per  cent  ammonia  and  a  bottoms  con- 
taining 0.1  mol  per  cent  ammonia.  The  tower  and  total  condenser  will  operate 
at  an  absolute  pressure  of  20  atm.  The  feed  will  enter  the  system  at  20°C.  and 
be  heated  to  40°C  in  the  condenser  as  shown  in  Fig.  7-20. 
Using  the  data  and  notes  given  below,  calculate: 

1.  The  minimum  reflux  ratio  0/D,  using  the  enthalpy-composition  method. 

2.  The  number  of  theoretical  plates  required  at  0/D  equal  to  1.5  times  the 
minimum  value  found  in  Part  1,  using  the  enthalpy-composition  method. 

3.  The  number  of  theoretical  plates  required  at  total  reflux. 

4.  The  minimum  reflux  ratio  0/D,  assuming  constant  molal  overflow  rates. 

5.  The  number  of  theoretical  plates  required  for  0/D  equal  to  1.5  times  the 
value  obtained  in  Part  4,  using  constant  molal  overflow  rates. 

Data  and  Notes.  As  a  basis,  the  enthalpies  of  pure  liquid  water  and  pure  liquid 
ammonia  are  taken  as  zero  at  their  boiling  points.  On  this  basis  a  20  weight  per 
cent  ammonia  solution  at  40°C.  has  an  enthalpy  equal  to  —3,180  cal  per  g.  mol. 
Other  enthalpy  and  equilibrium  data  (Ref.  2)  for  20  atm.  follow. 


Mol  fraction  NH3, 
in  liquid 
or  vapor 

Enthalpy  liquid 
cal.  per  g.  mol  vs. 
mol  fraction 
in  liquid 

Enthalpy  vapor 
cal.  per  g  mol  vs. 
mol  fraction 
in  vapor 

T,  °C.  vs. 
mol  fraction 
in  liquid 

0.0 

0 

8,430 

211.5 

0.1053 

-500 

7,955 

182  5 

0  2094 

-970 

7,660 

165  5 

0.312 

-1,310 

7,495 

131  5 

0  414 

-1,540 

7,210 

113  0 

0  514 

-1,650 

6,920 

91  0 

0.614 

-1,580 

6,620 

76  0 

0.712 

-1,400 

6,300 

65  5 

0.809 

-960 

5,980 

58.0 

0.905 

-520 

5,470 

52.5 

0.923 

-430 

5,350 

51.6 

0.942 

-310 

5,240 

50.8 

0.962 

-210 

5,050 

50.0 

0.981 

-110 

4,870 

49.3 

1.000 

0 

4,250 

48.5 

154 


FRACTIONAL  DISTILLATION 


VAPOR-LIQUID  EQUILIBBIA  AT  20  ATM. 

x  Mol  Fraction  NH3  y  Mol  Fraction  NH» 

0.0529  0.262 

0.1053  0.474 

0.2094  0.742 

0.312  0.891 

0.414  0.943 


0.514 
0.614 
0.712 
0.809 
1.00 

Solution  of  Part  1.     Basis;  100  mols  of  feed. 
Feed  composition 


NHg  balance, 


0.977 
0.987 
0.990 
0.995 
1.000 


0.209 


0.209(100)  -  0.98D  -f  0.001(100  -  D) 
D  -  21.25        and         W  -  78.75 


An  enthalpy  balance  on  the  exchanger  between  the  bottoms  and  the  feed  is 
WCP(tB  -  45)  -  F(hr  -  h,0) 
Feed 


FIG.  7-20.     Flowsheet  for  ammonia-water  example. 


RECTIFICATION  OF  BINARY  MIXTURES 


155 


0.3     03     0.4     05     06     0.7     0.8     0.9     1.0 
x  or  y ,  Mol  fraction  of  ammonia 
FIG.    7-20a.     Enthalpy-composition    diagram    for    ammonia-water    example. 

where  IB  **  temperature  in  reboiler  «•  211.2°C. 
hp  —  enthalpy  of  feed  entering  tower 
CP  «  heat  capacity  of  bottoms  «  19.0  cal.  per  g.  mol/°C,  (used  same  as 

Cp  of  H2O  at  20  atm.) 
hto  -  enthalpy  of  feed  at  40°C.  -  -3,180 
78.75(19.0)  (211.2  -  45)  -  100 (hr  +  3,180) 
hp  a*  —690  cal.  per  g.  mol 

This  point  is  located  on  the  enthalpy-composition  diagram  (Fig.  7-20a).     Utiliz- 
ing the  y,x  diagram  it  is  found  that  the  equilibrium  tie  line  that  passes  through  this 
point  corresponds  to  x  »  0.193  and  y  -  0.70.    The  extrapolation  of  this  tie  line 
to  XD  -  0.98  gives  M D  -  10,230  and  at  xw  -  0.001,  M w  ~  -3,630. 
Also 

VTHT  -  0RkR  4-  Qc  +  DhD  -  Osh*  + 


156 


FRACTIONAL  DISTILLATION 


Assuming  that  HR  =  ho  ™  enthalpy  of  saturated  liquid,  and  using  VT  «*  OB  -f  D, 
by  a  derivation  similar  to  that  for  Eq.  (7-16), 

OR  _  MD  -  HT 
D   **   H T  -  hn 


0un  4,850  +  100 
The  above  calculations  assume  that  the  minimum  reflux  ratio  corresponded  to  a 
pinched-in  condition  at  the  feed  point.  A  check  of  the  other  tie  line  above  and 
below  the  feed  plate  indicated  that  this  corresponded  to  the  maximum  value  for 
MD.  It  is  apparent  from  Fig.  7-21  that  the  operating  line  on  the  ytx  diagram  would 
have  to  be  very  curved  in  order  to  have  the  pinch  occur  at  any  place  but  the  feed 
plate. 


0.1      08     0.3     04      0.5     06      07      0.8     09      10 
Mol  fraction  NH3  in  liquid 

FIG.  7-21.     y,x  diagram  for  NHa-water  example., 
Solution  of  Part  2 


MD  -  12,940         and         Mw  -   -4,360 

Using  this  value,  the  steps  corresponding  to  theoretical  plates  are  stepped  off. 
In  this  manner  the  lines  on  Fig.  7-20a  were  constructed  for  the  top  seven  plates 
giving  the  vapor  entering  the  T  —  6  plate  equal  to  0.073.  Below  this  value  the 
diagram  becomes  difficult  to  use,  and  the  calculations  were  completed  by  the  use  of 
Eq.  (7-31),  using  Hm  -  8,400,  and  hm+l  «  -20.  ' 

8,420 


-. 

'      ~  MW  - 
2.92xw+1  -  1.92(0.001) 


-4,360  +  20 


RECTIFICATION  OF  BINARY  MIXTURES  157 

Over  the  low  concentration  region  involved,  Fig.  7-21  indicates  that  the  vapor- 
liquid  equilibrium  curve  can  be  expressed  as 


The  liquid  on  plate  T  —  7  will  be  in  equilibrium  with  a  vapor  composition  of  0.073 
and,  from  the  equilibrium  relationship  just  assumed,  XT-I  would  equal  0.0145. 
This  lower  section  is  to  reduce  this  mol  fraction  to  0.001  and,  by  Eq.  (7-77), 


_5_  J  A/00146  _    \ 
2.92         A  0.001  / 


4/2.92 

3.9  plates 


In  (5/2.92) 

Thus  a  still  and  about  eleven  (7  +  3.9)  theoretical  plates  are  required. 

Solution  of  Part  3.  At  total  reflux  the  plates  are  stepped  off  on  Fig.  7-21,  and  a 
still  plus  four  theoretical  plates  are  required. 

Solution  of  Part  4.  The  determination  of  the  minimum  reflux  ratio  on  the  basis 
of  the  usual  simplifying  assumptions  requires  determining  the  difference  Vm  —  Vn. 
A  given  enthalpy  for  the  feed  does  not  determine  this  difference  for  the  general 
case  because  the  feed-plate  location  is  still  a  variable;  however,  in  this  case,  a 
definite  value  is  obtained  at  the  minimum  reflux  ratio,  because  the  pinched-in 
condition  occurs  at  the  feed  plate. 

By  an  enthalpy  balance  around  the  feed  plate, 

Fhp  +  F/_iF/-.i  +  Of+ihf+i  =  VfHf  +  Ofhf 

where  subscripts/,  /  -f  1,  /  —  1,  refer  to  feed  plate,  plate  above,  and  plate  below, 
respectively. 

For  the  pinched-in  condition, 

H/-.I  =  Hf  —  H        and        hf  =  TI/+I  «  h 
Also, 

F  +  F/-I  +  Of+i  **V/+Of 

Solution  of  these  equations  gives 

V    -  F  -i  -  F  (hp  ~  h) 

For  a  first  approximation  take  the  enthalpy  values  H  and  h,  corresponding  to  a 
mol  fraction  of  ammonia  in  the  liquid  of  0.19;  h  =  —900  and  H  =  6,300.  From 
Part  1,  hp  =  -690. 

F(-—  690  -f-  900) 
Vf  -  F/-I  =     6,300  +  900      "  °'029F 

The  intersection  of  the  operating  lines  for  this  condition  falls  on  a  line  of  slope 
-(0.971/0.029)  =  -33.5,  which  passes  through  the  y  =  x  line  at  ZF  =  0.209. 
Such  a  line  on  Fig.  7-21  cuts  the  equilibrium  curve  at  x  —  0.193,  y  —  0.71  which  is 
close  enough  to  the  assumed  value,  and 

0.98  -  0.71   _  0_343 


F/min         0.98  -  0.193 

°'343     ,  0.522 


D  /„,„      0.657 


158  FRACTIONAL  DISTILLATION 

It  will  be  noted  that  this  value  is  much  lower  than  that  obtained  in  Part  2,  and 
designing  for  an  actual  reflux  ratio  even  twice  this  value  would  still  give  heat 
requirements  less  than  the  true  minimum. 

Solution  of  Part  5.  For  this  case,  the  actual  value  of  (0/D)  is  to  be  1 .5  times  that 
obtained  in  Part  4,  and  the  plates  are  to  be  calculated  with  the  usual  simplifying 
assumptions. 

(<jL\      **  1.5(0.522)  -  0.783 

On  -  0.783(21.25)  -  16.65 
Vn  -  37.9 

In  this  case  the  value  of  Vn  •—  Vm  will  not  be  quite  equal  to  that  calculated  in 
Part  4  because  H/  will  not  equal  H/+I,  and  h/  will  not  equal  h/+i,  but  the  change 
will  not  be  large  and  the  same  difference  will  be  used  as  an  approximation. 

Vm  -  Vn  -  -2.9 
Vm  -  35        Om  -  113.75 


(?).--*«  - 


These  lines  are  drawn  on  Fig.  7-21,  and  the  corresponding  steps  are  shown.  A  still 
plus  approximately  12  theoretical  plates  are  required.  This  is  in  close  agreement 
with  the  result  of  Part  2,  and  this  is  generally  the  case.  Thus,  if  a  tower  is  calcu- 
lated on  the  basis  of  a  certain  factor  times  the  minimum  reflux  ratio,  the  theoretical 
plates  required  are  usually  approximately  the  same  on  either  the  (y,x)  or  enthalpy- 
composition  basis,  provided  the  method  chosen  is  used  consistently  throughout  the 
calculation.  However,  the  heat  and  cooling  requirements  may  be  seriously  in 
error  when  calculated  on  the  constant  overflow  basis. 

Modified  Latent  Heat  of  Vaporization  Method.  The  use  of  the 
Ponchon-Savarit  method  has  two  major  disadvantages:  (1)  The 
enthalpy-composition  relations  for  the  vapor  and  liquid  are  required 
over  the  whole  operating  range  and  (2)  it  is  limited  to  binary  mix- 
tures. On  page  141  it  was  shown  that  the  variation  in  0/V  was  due 
mainly  to  a  variation  in  the  difference  between  the  enthalpy  of  the 
vapor  entering  and  the  liquid  leaving  a  plate,  and  since  the  enthalpy 
of  either  the  vapor  or  liquid  does  not  vary  greatly  over  the  concentra- 
tion range  corresponding  to  one  plate,  this  difference  in  enthalpy  is 
essentially  equal  to  the  latent  heat  of  vaporization  for  the  concentra- 
tion region  involved.  Neglecting  heat  of  mixing  effects,  this  heat  of 
vaporization  can  be  approximated  as 


(H  -  h)n  -  [yi(H  -  k)i  +  y*(H  -  A), 

where  (H  —  A)i,  (H  —  A)  2  are  the  latent  heats  of  vaporization  of  the 
pure  components. 

The  values  of  this  molal  enthalpy  difference  vary  chiefly  because  the 
heats  of  vaporization  of  the  individual  components  are  different  rather 


RECTIFICATION  OF  BINARY  MIXTURES  159 

than  because  their  values  vary  over  the  temperature  range  of  the 
column.  Thus,  if  the  latent  heats  of  vaporization  of  all  the  com- 
ponents were  the  same,  the  enthalpy  difference  would  be  approximately 
constant,  independent  of  the  composition.  It  is  pointed  out  on  page 
141  that  this  condition  leads  to  essentially  straight  operating  lines  for 
most  cases. 

If  pseudo  mol  fractions,  y',  x',  z',  and  flow  quantities,  F',  0',  W,  F', 
could  be  defined  such  tha£  the  following  five  relationships  apply: 

Vryf(H  —  K)B  =  Vy(H  —  h)  for  each  component1 

y\  +  y'*  +  y*  +  •  •  •  =  1 
x(  +  xi  +  4  +  '  *  •  =  1 

n  =  O'n+1  +  D' 

V'ny'n  =  #n+i<+i  +  D'XB  for  each  component 

then  the  design  could  be  carried  out  with  the  pseudo  quantities,  and 
the  molal  enthalpy  difference  (H  —  h)  would  be  equal  to  (H  —  h)R  for 
all  compositions  and  would  make  the  usual  simplifying  assumptions  a 
good  approximation.  Such  relations  for  the  pseudo  values  in  terms 
of  the  normal  variables  are 


where  2fty  =  ftiyi  +  pzy*  +  foys  + 


where  %$x  =  0iXi  +  j32x2  + 


Dr  ==  D20XD 
Ff  « 

W  - 


In  these  equations  the  values  of  ft  are  equal  to  the  (H  —  h)  values  for 
the  component  in  question  divided  by  (H  —  h)R.  If  the  (H  —  h) 
values  for  individual  components  are  constant  over  the  operating 
range  involved,  ft  will  be  constant  and  a  design  problem  in  the  new 

1  Where  (H  —  K)R  «  same  constant  for  all  components  and  (H  -  h)  -  enthalpy 
difference  value  for  the  individual  component  and  is  assumed  to  be  constant  over 
the  temperature  range  involved, 


160  FRACTIONAL  DISTILLATION 

coordinates  will  give  straight  operating  lines.  This  method  necessi- 
tates calculating  the  equilibrium  curves  and  all  the  flow  quantities 
over  to  the  new  coordinates.  At  the  same  temperature,  the  relative 
volatility  is  the  same  for  both  coordinates.  However,  if  the  relative 
volatility  varies  with  the  temperature  the  y',xf  curve  will  be  different 
from  the  y,x  curve  because  the  temperature  will  be  different  for  equal 
values  of  x  and  xf.  The  design  problems  can  then  be  carried  out  with 
these  new  variables,  making  the  usual  simplifying  assumptions.  After 
the  number  of  theoretical  plates  and  design  conditions  is  obtained  on 
this  pseudo  basis,  the  results  can  be  converted  to  the  true  flow  quanti- 
ties and  mol  fractions.  The  (H  —  h)R  value  can  be  arbitrarily  chosen 
although  it  is  usually  convenient  to  make  it  equal  to  the  (//  —  h) 
value  for  one  of  the  components  because  this  makes  0  for  that  com- 
ponent equal  to  1.  The  use  of  these  pseudo  variables  is  equivalent  to 
assigning  a  fictitious  molecular  weight  to  a  component  such  that  its 
molal  heat  of  vaporization  on  the  new  basis  will  be  a  specified  value. 
The  use  of  this  method  will  be  illustrated  by  the  ammonia-water 
example  already  considered. 

Ammonia -Water  Example.  From  page  153  the  latent  heats  of  vaporization  for 
ammonia  and  water  are  4,250  and  8,430  cal.  per  g.  mol,  respectively.  Using 
(H  -  h)R  =  8,430  gives  0H2o  =  1.0  and  j3NH3  =  4,250/8,430  =  0.504.  Recal- 
culating the  feed  and  terminal  concentrations  to  the  pseudo  units, 


"      0.504(0.98)  +  0.02 

0.504(0.209)          _01175 
*      0.504(0.209)  +  0.791 


Per  100  mols  of  original  feed, 

D'  -  21.25(0.504(0.98)  +  0.02]  =  10.9 
F'  -  100(0.504(0.209)  +  0.791]  -  89.6 
W  =  [0.504(0.001)  +  0.999]  -  78.7 

The  vapor-liquid  equilibrium  data  were  recalculated  to  the  new  basis  and  the 
results  are  plotted  in  Fig.  7-2 la. 

To  determine  the  minimum  reflux  ratio,  it  is  necessary  to  locate  the  intersection 
of  the  operating  line  and  the  equilibrium  curve.  From  page  155  the  enthalpy  of 
the  feed  is  —690  cal.  per  g.  mol  and  the  enthalpy  of  saturated  liquid  of  the  feed 
composition  is  —970  cal.  per  g.  mol  (both  values  on  original  basis).  Thus  the 
excess  enthalpy  of  the  feed  above  saturated  liquid  is  100(280)  -  28,000  cal.  The 
latent  heat  on  the  new  basis  is  8,430  cal.  per  g.  mol  and  the  fraction  vapor  in  the 


RECTIFICATION  OF  BINARY  MIXTURES 


161 


feed  is 


Fraction  vapor  - 


0.037 


The  line  of  intersection  of  the  operating  line  will  cross  the  yf  «•  x*  line  at 
=  0.1175  and  will  have  a  slope  equal  to   -0.963/0.037  =  —26.     This  line  is 


0      0.1      0.2     0.3     0.4     0.5     06     07     0.8    09     1.0 


FIG.  7-21a. 

shown  in  the  figure.     This  line  intersects  the  equilibrium  curve  at  x'  =  0.102. 
y'  ss  0.517  and  the  minimum  reflux  ratio  corresponds  to 

QL\       =  0.961  -  0.517 

F'/mln          0.961    -  0.104          U'&1/ 

0.517 


(°*\       =  °'517 

VP'/nnx          0.483 


1.07 


Because  the  compositions  of  OR  and  D'  are  the  same,  the  conversion  factors  to 
OR  and  D  will  be  the  same  and 

(°*\       «  1.07 


This  value  is  close  to  one  obtained  on  page  156,  and  the  two  values  probably  agree 
within  the  accuracy  of  the  calculation.     For 


(°*\      =1.5^)       =1.605 
\/>7.ct  \D'Jma 


1.605 


=  0.616 


^V'Jn      2.605 
The  corresponding  operating  lines  are  drawn  on  the"  diagram,  and  the  theoretical 


162  FRACTIONAL  DISTILLATION 

plates  stepped  off  in  the  usual  manner.  The  diagram  indicates  that  a  still  and 
11  theoretical  plates  are  required.  This  checks  the  result  obtained  by  the 
enthalpy-composition  method. 

This  modified  latent  heat  of  vaporization  method  generally  gives 
good  results  and  would  be  only  appreciably  in  error  if  (1)  the  heats  of 
mixing  were  large  or  (2)  the  latent  heats  of  vaporization  of  the  com- 
ponent varied  appreciably  over  the  temperature  range  involved.  In 
the  latter  case  the  results  can  be  improved  by  using  separate  average 
values  for  stripping  and  enriching  section. 

Theoretically  the  enthalpy-composition  method  is  more  exact  than 
the  use  of  the  modified  latent  heats  of  vaporization,  but  in  most  cases 
the  two  agree  within  the  accuracy  of  the  calculation.  For  binary 
mixtures,  if  the  necessary  enthalpy  data  are  already  available,  the 
enthalpy-composition  method  is  the  easier  to  apply;  if  the  data  are  not 
available  and  must  be  calculated,  then  the  other  method  is  the  more 
convenient.  For  multicomponent  mixtures  the  modified  latent  heat 
method  is  more  convenient  even  if  the  complete  enthalpy  data  are 
available. 

HEAT  ECONOMY 

In  the  section  on  the  Optimum  Reflux  Ratio  it  was  pointed  out  that 
the  operating  costs,  i.e.,  the  heating  and  cooling  charges,  were  fre- 
quently the  major  cost  of  a  distillation  process  and  that  fixed  charges 
on  the  equipment  were  small.  In  such  cases,  it  would  appear  desirable 
to  decrease  the  heating  and  cooling  requirement  at  the  expense  of 
additional  equipment,  and  this  section  will  consider  a  number  of 
methods  of  accomplishing  that  result. 

From  a  thermodynamic  viewpoint  the  inefficiencies  of  a  distillation 
operation  can  be  grouped  into  two  main  categories:  (1)  those  that  are 
a  function  of  the  distillation  process  itself  and  (2)  those  that  are  related 
to  supplying  the  necessary  energy  to  the  materials  being  separated. 

Separation  Efficiency  of  a  Distillation  Column.  The  minimum 
isothermal  thermodynamic  work  required  for  separating  1  mol  of  a 
liquid  binary  mixture  into  its  pure  liquid  constituents  is  given  by  the 
following  equation: 


Minimum  work  -  AF  =  -RT  (xi  In  |~  +  x2  In  ^  )     (7-41) 

V       ^i  *V 


(7-42) 
-RT[xi  In  (y&i)  +  X*  In  (y&$]       (7-43) 


RECTIFICATION  OF  BINARY  MIXTURES  163 

where  AF  «  minimum  work  of  separation  (free-energy  change)  per 

mol  of  mixture 
R  =  gas  law  constant 
T  =  absolute  temperature 
Xi,  £2  =  mol  fractions  of  components  in  mixture 
71,  72  =  activity  coefficients  of  two  components 
IT  *=  total  pressure  - 
p  =  partial  pressure, 

or  fugacity  of  component  in  mixture 
P  =  vapor  pressure, 

or  fugacity  of  pure  saturated  liquid 
2/i,  2/2  =  mol  fractions  in  equilibrium  vapor 

For  an  ideal  solution,  this  expression  has  a  maximum  value  at  a  mol 
fraction  of  0.5,  and  for  this  condition  the  minimum  work  required  is 
equal  to  0.693flT'.  Thus  to  separate  1  Ib.  mol  of  an  ideal  mixture  of 
this  composition  at  a  temperature  of  212°F.  would  require  about 
920  B.t.u. 

In  case  the  mixture  is  not  completely  separated,  the  minimum  work 
required  per  mol  is  obviously  less  and  can  be  calculated  from  the 
following  equation: 

AF  =  —  RT[(xi  In  yiXi  +  x2  In  72z2)  —  D(XID  In  yiDXiD 

+  XZD  In  72i>z2z>)  —  W(xiw  In  71^1^  +  x^w  In  72^2^)]     (7-44) 

where  D  refers  to  distillate  and  W  refers  to  bottoms. 

Mixtures  with  positive  deviations  from  Raoult's  law,  i.e.,  solutions 
with  activity  coefficients  greater  than  1,  require  lower  minimum  work 
for  separation  than  do  ideal  solutions;  the  reverse  is  true  for  those  solu- 
tions with  negative  deviations.  For  example,  consider  the  minimum 
work  for  separating  a  3  mol  per  cent  solution  of  ethanol  and  water  at  its 
normal  boiling  point  into  pure  water  and  87  mol  per  cent  ethanol.  The 
boiling  point  of  such  a  solution  is  173°F.,  and  the  activity  coefficients 
are  4.4  and  1.01  for  the  3  mol  per  cent  solution  for  ethanol  and  water, 
respectively;  the  corresponding  values  for  the  87  per  cent  solution  are 
1.01  and  2.2.  On  the  basis  of  Eq.  (7-44),  the  minimum  work  of  sepa- 
rating 1  mol  of  this  mixture  into  the  desired  product  wo.uld  be  about 
106  B.t.u.  Separating  an  ideal  solution  of  the  same  composition  into 
the  same  products  would  require  153  B.t.u.  The  actual  energy 
requirement  is  lower  than  the  theoretical  because  ethyl  alcohol  and 
water  have  positive  deviations  from  Raoult's  law,  indicating  a  tend- 
ency to  immiscibility.  An  immiscible  system  requires  essentially  no 
work  for  separation.  On  the  other  hand,  systems  of  negative  devia- 


164  FRACTIONAL  DISTILLATION 

tions  from  Raoult's  law,  i.e.,  those  that  tend  to  maximum  boiling 
azeotropes,  would  need  more  work  for  separation  than  an  ideal 
solution. 

The  energy  required  for  separating  a  mixture  is  supplied  by  adding 
heat  to  the  fluid  in  the  still  and  removing  heat  at  a  lower  temperature 
level  in  the  condenser.  The  available  work  energy,  based  on  an 
isentropic  process,  in  the  heat  supplied  to  the  liquid  in  the  still  can  be 
calculated  on  the  basis  of  the  following  equation: 

Ws  =  available  work  =  Q  — ^-^  (7-45) 

1  w 

where  Q  =  heat  added 

Tw  —  absolute  temperature  of  liquid  in  still 
To  =  absolute  temperature  at  which  heat  can  be  discharged,  i.e., 

temperature  of  cooling  water 

The  available  work  equivalent  to  the  heat  removed  from  the  con- 
denser can  be  calculated  by  the  following  equation: 

Wc  -  available  work  =  Qc  — ^-°  (7-46) 

JL  c 

where  Qc  =  heat  removed  from  condenser 

Tc  =  temperature  of  condensation  of  distillate 
In  order  to  simplify  the  following  discussion,  it  will  be  assumed  that 
the  heat  added  to  the  still  is  equal  to  the  heat  removed  in  the  con- 
denser. This  is  essentially  true  when  the  feed  enters  as  a  liquid  at  its 
normal  boiling  point  and  when  the  distillate  and  bottoms  leave  as 
liquid  at  their  boiling  points.  Other  cases  can  be  handled  in  a  similar 
manner  but  are  more  involved.  For  a  system  in  which  Q3  =  Qc,  the 
net  available  work  supplied  in  the  heat  to  the  distillation  process  itself 
is  equal  to 


Te 


The  minimum  heat  that  can  be  utilized  in  a  distillation  corresponds 
to  the  minimum  reflux  condition  and,  for  a  binary  mixture,  can  easily 
be  determined  from  the  enthalpy-composition  diagram. 

It  is  interesting  to  compare  the  thermodynamic  minimum  work  with 
that  required  by  the  actual  distillation  process  at  the  minimum  reflux 
condition.  For  ideal  mixtures  of  close  boiling  compounds  Eq.  (7-47) 
can  be  combined  with  Eqs.  (7-44)  and  (7-56),  and  approximations  for 


RECTIFICATION  OF  BINARY  MIXTURES 


165 


the  relative  volatility  and  latent  heats  as  a  function  of  temperature 
can  be  made  to  give  simplified  expressions,  but  they  are  misleading  for 
most  actual  distillations.  It  is  more  instructive  to  study  the  ratio  of 
Eq.  (7-44)  to  Eq.  (7-47)  for  actual  cases.  Such  calculations  for  the 
systems  benzene-toluene  and  ethanol-water  at  atmospheric  pressure 
are  given  in  Tables  7-5  and  7-6.  In  the  case  of  the  first  system,  it  was 
assumed  that  complete  separation  was  being  obtained;  in  the  second 
case,  that  the  distillate  was  87  per  cent  mol  alcohol  and  the  bottoms 
was  pure  water.  In  both  cases,  actual  activity  coefficients  were 
employed,  minimum  heat  requirements  were  estimated  from  enthalpy- 
composition  diagrams,  and  pressure  drop  and  heat  losses  were  neglected. 

TABLE  7-5.     FRACTIONATION  EFFICIENCY  FOR  ETHANOL-WATER  MIXTURE 
(Tw  =  672°R.,  To  =  G33°H.,  T0  =»  546°R.) 


Mol  fraction 
ethanol  m 
feed 

Minimum  heat 
requirement, 
B.t.u.  per  Ib. 
mol  of  feed 

Thermo- 
dynamie  work 
equation  (7-44), 
B.t.u.  per  Ib. 

Available  work 
equation  (7-47), 
B.t.u.  per  Ib. 
mol  of  feed 

Thermo- 
dynamic 
efficiency, 
per  cent 

mol 

0 



0 



0 

0.03 

2,620 

106 

J31 

81 

0.1 

8,740 

195 

437 

45 

0.2 

17,500 

261 

875 

30 

0.3 

26,200 

310 

1,310 

24 

0.4 

34,900 

288 

1,745 

17 

0.5 

43,600 

283 

2,180 

13 

0.6 

52,400 

234 

2,620 

9 

0.7 

61,200 

195 

3,060 

6 

0.8 

69,600 

90 

3,480 

3 

The  benzene-toluene  system  shows  a  maximum  thermodynamio 
efficiency  of  about  80  per  cent  for  a  feed  composition  of  40  per  cent 
benzene.  The  efficiency  decreases  for  feeds  both  weaker  and  stronger, 
but  in  the  range  of  feed  compositions  from  0.1  to  0.8  it  is  good.  Equa- 
tion (7-44)  shows  that  the  efficiency  is  zero  both  for  XF  approaching  0 
and  1.0.  In  the  case  of  the  ethanol-water  system,  the  heat  require- 
ment per  mol  of  distillate  is  essentially  independent  of  the  feed  compo- 
sition over  the  range  shown  because  the  minimum  reflux  ratio  condi- 
tion corresponds  to  a  tangency  with  the  equilibrium  curve  at  about 
84  mol  per  cent  alcohol  Thus  the  reflux  ratio  is  the  same  over  the 
whole  concentration  region  considered.  This  reflux  ratio  corre- 


166 


FRACTIONAL  DISTILLATION 


spends  closely  to  a  pinched-in  region  for  the  0.03  feed  as  well  as  to  the 
pinched-in  condition  at  the  mol  fraction  of  0.84.  Thus  the  operating 
line  follows  the  equilibrium  curve  approximately,  and  the  efficiency  for 
the  0.03  feed  concentration  is  high.  For  the  higher  strength  feeds, 
the  operating  lines  over  most  of  the  lower  concentration  regions  are  a 
considerable  distance  from  the  equilibrium  curve.  This  leads  to  low 
efficiency. 

TABLE  7-6.    FRACTIONATION  EFFICIENCY  FOR  BENZENE-TOLUENE  MIXTURE 
(TV  -  692°R,  Tc  -  636°R.,  T0  -  546°R.) 


Mol  fraction 
benzene  in 
feed 

Minimum  heat 
requirement, 
B.t.u.  per  Ib. 
mol  of  feed 

Eq.  (7-43) 
thermo  dynamic 
work,  B.t.u. 
per  Ib.  mol 

Eq.  (7-47) 
net  available 
work,  B.t.u.  per 
Ib.  mol 

Thermo- 
dynamic 
efficiency, 
per  cent 

0 



0 

— 

0 

0.1 

13,100 

439 

895 

49 

0.2 

14,100 

662 

964 

69 

0.3 

15,200 

810 

1,040 

78 

0.4 

16,200 

882 

1,110 

80 

0.5 

17,200 

900 

1,180 

76 

0.6 

18,300 

880 

1,250 

70 

0.7 

19,300 

785 

1,320 

60 

0.8 

20,300 

630 

1,390 

45 

0.9 

21,400 

410 

1,460 

28 

1.0 

— 

0 

— 

0 

The  values  given  in  the  tables  were  the  maximum  efficiencies  and 
correspond  to  the  minimum  reflux  ratio.  In  an  actual  tower,  the  heat 
supply  would  be  greater  and  the  efficiency  values  should  be  reduced. 
The  reduction  factor  is  approximately 


Efficiency  of  Supplying  and  Removing  Energy  from  a  Distillation 
Tower,  In  a  simple  distillation  tower,  the  greatest  inefficiency 
usually  results  from  the  methods  employed  in  supplying  energy  to  the 
system  rather  than  in  the  separation  process.  For  example,  consider 
the  benzene-toluene  mixture  of  Table  7-6.  Even  allowing  for  the 
actual  reflux  ratio  being  greater  than  the  minimum  reflux  ratio,  a 
thermodynamic  efficiency  of  50  per  cent  or  greater  would  be  obtained 
over  a  wide  range  of  feed  compositions.  However,  in  such  a  case  it 


RECTIFICATION  OF  BINARY  MIXTURES 


167 


would  be  common  practice  to  use  steam  as  a  heating  fluid  with  a  con- 
densation temperature  of  at  least  250°F.  This  would  give  a  tempera- 
ture difference  for  heat  transfer  in  the  still  of  18°F.  There  would  then 
be  a  temperature  drop  through  the  distillation  system  corresponding 
to  56°F.,  and  with  cooling  water  available  at  80°F.  there  would  be  tem- 
perature loss  equivalent  to  96°F.  in  going  from  the  overhead  vapor  to 
the  available  heat  sink.  'These  temperatures  and  the  corresponding 
losses  in  availability  are  given  in  Table  7-7.  It  is  evident  that  the 

TABLE  7-7 


Temp., 
°R. 

Net  available  work,  B.t.u.  per 
B.t.u.  of  heat 

Per  cent 
of  total 

Steam  ... 

710 

Steam  to  still,                       0  02 

s 

Still       

692 

Still  to  distillate,                  0  069 

29 

Distillate      

636 

Distillate  to  cooling  water,  0.151 

63 

Cooling  water  

540 

Total 

0  24 

100 

major  inefficiency  is  in  adding  and  removing  the  heat.  In  this  case 
over  70  per  cent  of  the  total  available  work  in  the  heat  was  dissipated 
in  these  ways.  In  the  case  of  a  3  per  cent  ethanol  mixture  distilled  at 
atmospheric  pressure,  the  loss  in  available  energy  of  the  steam  due  to 
heating  and  cooling  could  be  over  80  per  cent.  These  heat-transfer 
processes  are  the  source  of  the  major  thermal  inefficiencies  in  most 
industrial  distillations.  For  these  cases,  the  greatest  improvement  in 
efficiency  will  be  obtained  by  reducing  these  losses.  The  most  obvious 
way  of  accomplishing  this  result  is  by  utilizing  a  higher  fraction  of  the 
temperature  difference  between  the  source  and  sink  temperatures  for 
the  actual  distillation  process  and  less  for  the  heat  transfer. 

METHODS  OF  INCREASING  EFFICIENCY 

1.  Separation  Process.  One  method  of  reducing  the  irreversibility 
in  the  fractionating  tower  is  to  make  the  vapor  stream  entering  a  plate 
more  nearly  in  equilibrium  with  the  liquid  on  the  plate.  The  extreme 
in  this  case  is  for  this  vapor  and  the  liquid  to  be  in  equilibrium,  but  this 
obviously  means  no  enrichment  per  plate.  To  accomplish  this  condi- 
tion throughout  the  tower  requires  that  the  operating  line  coincide 
with  the  equilibrium  curve,  and  this  in  turn  involves  a  different  reflux 
ratio  for  each  plate,  i.e.,  heat  must  be  added  to  or  removed  from  each 
plate  in  the  tower.  It  is  obvious  that  such  a  system  is  impractical. 


168  FRACTIONAL  DISTILLATION 

However,  certain  approaches  have  been  made  in  this  direction.  For 
example,  in  most  systems  the  vapor  load  is  limited  by  the  region  around 
the  feed  plate.  Actually  a  lower  quantity  of  vapor  would  be  suitable 
for  the  regions  near  the  top  or  bottom  of  the  tower,  and  it  would  be 
possible  to  operate  a  fractionating  tower  by  supplying  less  heat  at  the 
bottom  than  corresponds  to  the  required  reflux  ratio  and  then  to 
supply  additional  heat  at  some  point,  or  points,  intermediate  between 
the  still  and  the  feed  region.  This  would  be  advantageous  where  an 
additional  heat  source  was  available  which  was  not  at  a  temperature 
level  sufficient  to  operate  the  still  but  was  suitable  for  the  intermediate 
region.  Likewise,  above  the  feed  plate,  heat  could  be  removed  at 
some  intermediate  position  and  used  for  other  heating  purposes. 

For  example,  in  the  ethanol  separation  given  in  Table  7-5  high  feed 
concentrations  gave  low  fractionating  efficiency.  This  was  because 
lower  quantities  of  heat  could  have  been  used  for  the  concentration 
region  below  a  mol  fraction  of  ethanol  less  than  0.8,  and  the  high  heat 
requirement  resulted  from  the  pinched-in  region  at  the  top  of  the 
tower.  In  this  case,  a  high  percentage  of  the  heat  could  have  been 
added  at  a  temperature  level  only  a  few  degrees  above  that  of  the  con- 
denser rather  than  at  the  higher  temperature  of  the  still.  This  would 
have  given  a  much  higher  thermodynamic  efficiency. 

Such  modifications  result  in  a  lower  enrichment  per  plate  and  thus 
require  more  plates,  but  they  might  be  attractive  in  certain  cases. 
Actually,  they  are  seldom  advantageous. 

2.  Heat -transfer  Process.  In  most  cases,  using  a  higher  percentage 
of  the  available  work  in  the  heat  supplied  to  the  still  is  more  practical 
than  attempting  to  reduce  the  irreversibility  in  the  distillation  column. 
The  available  work  utilized  is  relatively  small  because  the  difference 
in  the  absolute  temperature  in  the  still  and  the  condenser  is  not  large. 
If  the  heat  could  be  employed  over  a  wider  temperature  differential, 
the  efficiency  would  be  increased.  The  most  common  method  of 
accomplishing  this  result  is  by  the  multieffect  principle  or  a  modifica- 
tion of  it.  For  example,  the  mixture  can  be  fractionated  in  two  or 
more  towers  arranged  with  operating  pressures  such  that  the  distillate 
condensing  temperature  of  one  tower  is  high  enough  to  serve  as  the 
heating  medium  for  the  succeeding  towers.  This  is  entirely  analogous 
to  the  use  of  multieffect  evaporators.  Figure  7-22  illustrates  the 
multieffect  system,  and  Figs.  7-23  to  7-26  show  modifications  of  it. 

Multieffect.  This  system  reduces  the  heat  required  roughly  in  pro- 
portion to  the  number  of  columns  employed.  This  is  not  quite  true 
because  there  will  be  changes  in  the  relative  volatility  with  pressure 


RECTIFICATION  OF  BINARY  MIXTURES 


169 


that  may  make  some  difference.  It  is  possible  to  make  further  heat 
savings  by  taking  the  bottoms  from  the  higher  pressure  columns  and 
expanding  them  directly  into  the  bottom  of  the  succeeding  columns, 
thereby  obtaining  a  certain  amount  of  sensible  heat.  The  sensible 
heat  of  the  condensate  from  a  high-pressure  stage  can  also  be  used  in  a 
low-pressure  stage. 


FIG.  7-22.     Multitower  system. 

If  the  relative  volatility  does  not  change  with  pressure,  the  total 
volume  of  the  distillation  system  will  be  essentially  the  same  as  for  a 
single  column  operating  at  about  the  average  pressure  of  the  multi- 
effect  system.  Thus  the  column  costs  will  be  of  the  same  order  for  the 
single-  and  multieffect  system.  However,  the  heat-transfer  surfaces 
required  will  be  larger  for  the  latter  due  to  the  lower  temperature 
differences  involved,  and  usually  more  cooling  water  will  be  required. 
The  control  problem  will  be  greater  than  for  a  single  column,  but  it 
would  not  be  particularly  difficult.  Thus,  the  multieffect  system 
should  be  attractive  where  the  heating  and  cooling  costs  are  large 
relative  to  the  equipment  costs. 

The  number  of  stages  that  can  be  employed  is  relatively  limited 
because  of  the  high  temperature  drop  per  stage.  For  example,  in 
distilling  the  mixture  ethanol  and  water,  it  would  be  difficult  to 
operate  without  having  each  stage  cover  a  temperature  differential  of 
the  order  of  60  to  80°F.  Thus,  a  system  having  more  than  two  stages 
would  require  a  high-temperature  heat  source  and  would  involve  rela- 
tively high  pressures. 

In  order  to  evaluate  the  heat  requirement  for  the  multitower  system, 


170 


FRACTIONAL  DISTILLATION 


consider  the  case  of  the  separation  of  a  binary  mixture  in  a  single 
tower,  for  which  the  relative  volatility  is  independent  of  the  pressure 
and  composition.  Making  the  usual  simplifying  assumptions,  the 
minimum  vapor  for  a  system  with  Vn  =  Vm  and  xj>  =  1,  xw  =  0  is 


v 

Kmln 


a  -I 


(7-48) 


where  Fmi» 
F 


compressor 


minimum  vapor  rate 
feed  rate 

a  =  relative  volatility 
XF  —  feed  composition 
For  a  multieffect  system  based  on  the  same  assumptions 

Vi  _  1  +  (a .—  l)xy 

JJT        "~" "  /  ^^     •*  \ 

where  Vi  =  minimum  vapor  generation  required  for  first  effect 

n  =  number  of  effects 

Vapor  Recompression.  One  method  that  has  been  proposed  and 
utilized  on  a  limited  scale  is  vapor  recompression,  a  schematic  diagram 

of  which  is  shown  in  Fig.  7-23.  In  this 
case,  the  overhead  vapor  is  compressed  to 
a  pressure  such  that  its  condensation 
temperature  will  be  suitable  for  heating 
the  still.  Theoretically,  such  a  method 
appears  attractive  because  the  heat  needs 
to  be  pumped  through  a  relatively  small 
temperature  rise.  However,  in  practice 
the  system  has  not  been  particularly  at- 
tractive due  to  the  high  cost  of  efficient 
vapor  compressors  of  the  size  required 
for  distillation  units,  and  wheii.  the  ef- 
ficiency of  the  compressor  and  the  effi- 
ciency for  producing  the  power  to  drive 
the  compressor  are  included,  a  great  deal  of  the  theoretical  advantage 
is  lost. 

Vapor  Rewe.  A  modification  of  the  multieffect  system,  shown  in 
Fig.  7-24,  has  been  termed  the  vapor  reuse  system  (Ref.  12).  In  it, 
the  feed  is  introduced  into  a  stripping  tower,  and  the  vapor  from  it  is 
used  as  the  source  of  heat  and  the  feed  of  the  lower  pressure  column. 
The  two  columns  may  not  be  in  balance;  i.e.,  the  minimum  heat 
required  for  the  stripping  section  may  be  more  or  less  than  that 


w      0 

7-23.     Vapor   recompres- 


FIG. 

aion.  system. 


RECTIFICATION  OF  BINARY  MIXTURES 


171 


required  for  the  fractionating  column.     This  heat  differential  can 
either  be  added  or  be  removed  from  the  still  of  the  lower  pressure 
fractionating   column.    In   case  the 
heat    requirement    of   the   low-pres- 
sure tower  is  greater  than'  for  the 
stripper,  it  might  be  thought  that 
the  additional  heat  could  be  employed 
advantageously  through  the  stripping 
column  as  well  as  through  the  frac- 
tionating column.     However,  this  is 
not  the  case  because  this  additional 
heat  would  require  that  more  vapor 
leave  the  top  of  the  stripping  column 
and  thus  it  must  be  a  diluter  vapor, 
which  makes  the  work  of  the  frac- 
tionating column  more  difficult.     For        FlG-  7"24'    Vapor  reuse  sysfcem- 
the  same  idealized  case  considered  in  the  multieffect  section,  the 
minimum  vapor  is 

For  the  stripping  tower, 

F 

(a  -  l)xr]  (7-49) 


Flmin  =  - 


and,  for  the  fractionating  tower, 

V  2  mm    == 


a(a  — -  1) 

If  XF   <   ox 1-\ '    Vl  *"n    >    F2  mm 


(7-60) 


If 


2(«  -  I)' 
a  -  2 


'i...  -  F2min 


If  Xp    >   H7 :r~TY'    ^l'»'n    <    ^2min 

Split  Tower.  Figure  7-25  shows  a  split-tower  system  which  partly 
separates  the  feed  in  one  fractionating  column  and  completes  the  frac- 
tionation  in  the  second  column.  It  is  similar  to  the  previous  case 
except  that  a  fractionating  column  with  reflux  is  employed  instead  of 
the  stripping  column.  This  system  can  be  adjusted  so  that  the  two 
towers  are  in  heat  balance.  The  feed  is  shown  entering  the  high- 
pressure  tower.  This  makes  relatively  pure  bottoms  but  impure  dis- 
tillate, which  is  then  fractionated  in  the  second  tower.  This  arrange- 
ment is  desirable  for  feeds  with  low  concentrations  of  the  more  volatile 
component.  If  the  feed  concentration  is  high,  it  is  more  desirable  to 


172 


FRACTIONAL  DISTILLATION 


fractionate  the  feed  first  into  pure  distillate  and  impure  bottoms  and 
refractionate  the  bottoms  in  the  lower  pressure  column. 

This  system  involves  a  lower  over-all  temperature  drop  than  a  two- 
stage  multieffect  system  and  thereby  makes  the  design  of  the  heat- 
transfer  surfaces  easier.  However,  the  multieffect  system  has  the 


»l  *t 

FIG.  7-25.     Split-tower  system. 

advantage  that  it  uses  all  the  heat  over  the  whole  concentration 
region  each  time.  This  system  utilizes  the  heat  over  a  limited  con- 
centration region  twice  but  over  other  regions  only  once.  Thus,  it 
tends  to  give  a  higher  reflux  ratio  in  the  central  portion  of  the  column 
and  lower  reflux  ratios  at  the  ends.  The  minimum  vapor  requirement 
for  this  system  for  the  assumptions  made  in  the  other  cases  is  given 
byEq,  (7-51): 

?[1  +  (a  - 


r  mm    — 


(a  -  1)[2  +  (a  - 


(7-51) 


This  arrangement  is  very  similar  to  that  employed  in  the  so-called 
double  towers  for  the  separation  of  oxygen  and  nitrogen.  In  this  case 
the  feed  is  relatively  rich  in  the  lower  boiling  component,  nitrogen, 
and  the  first  tower  is  used  to  produce  an  overhead  that  contains  a  high 
concentration  of  nitrogen  and  an  impure  bottoms  containing  40  to  50 
per  cent  oxygen.  The  liquid  nitrogen  overhead  product  from  the  high- 
pressure  tower  is  added  to  the  top  of  the  low-pressure  tower  and  serves 
as  the  only  reflux  for  it.  There  is  no  other  source  of  refrigeration  to 
produce  additional  reflux.  The  impure  bottoms  from  the  high-pres- 
sure tower  is  introduced  into  the  middle  portion  of  the  low-pressure 
tower.  Such  a  two-tower  system  will  give  high-purity  nitrogen  and 
oxygen  low  in  nitrogen  although  it  will  contain  appreciable  quantities 


RECTIFICATION  OF  BINARY  MIXTURES 


173 


of  noble  gases.  It  is  common  practice  in  this  case  to  build  the  low- 
pressure  tower  on  top  of  the  high-pressure  tower,  and  the  condenser- 
reboiler  is  a  common  unit  of  both  columns. 

A  further  modification  of  the  split-tower  system  is  given  in  Fig.  7-26 
in  which  the  feed  is  rectified  into  impure  bottoms  and  an  impure  dis- 
tillate, and  these  are  then  retreated  in  the  lower  pressure  tower.  In 
general,  this  system  gives  slightly  lower  heat  requirements  than  the 
previous  system,  and  it  utilizes  the  same  principle,  namely,  reusing  the 
heat  several  times  in  the  middle  region  concentrations  where  the  heat 
requirements  are  the  highest.  The  systems  of  Figs.  7-22  to  7-26  can 
be  utilized  with  three  or  more  towers  to  obtain  still  further  heat 
reductions. 


FIG.  7-26.     Modified  split-tower  system. 

In  order  to  show  the  comparison  between  the  different  systems,  the 
various  equations  have  been  plotted  in  Fig.  7-27  for  the  cases  of  feed 
concentrations  equal  to  0  and  1.0.  The  equation  for  the  multieffect 
system  has  been  plotted  for  n  =  1,  i.e.,  a  single  tower.  If  a  two-stage 
system  is  used,  the  values  should  be  divided  by  2,  etc.  For  the  vapor- 
reuse  system  the  maximum  vapor  requirement  for  the  two  columns  is 
given,  and  no  credit  has  been  applied  for  the  low-temperature  heat 
that  could  be  withdrawn  for  the  cases  with  the  relative  volatility 
greater  than  2.0.  It  will  be  noted  that  none  of  the  modifications  is  so 
effective  in  reducing  the  heat  requirements  as  the  multistage  system, 
although  some  of  them  are  equal  to  it  for  specialized  conditions.  Of 
the  specialized  arrangements,  vapor  reuse  would  not  appear  to  be  so 
attractive  as  the  split-tower  system  shown  although,  in  cases  where  the 
waste  heat  was  of  real  utility,  it  could  be  attractive.  The  relative 


174 


FRACTIONAL  DISTILLATION 


attraction  of  the  systems  will  be  shifted  somewhat  as  they  are  com- 
pared at  reflux  ratios  lower  than  the  minimum  (or  greater  than  total 
reflux).  It  still  appears  that,  for  a  given  number  of  towers,  the  multi- 
effect  system  is  the  most  attractive  from  the  heat  viewpoint.  Like- 
wise, for  a  given  total  quantity  of  heat  supplied  to  the  system  and  for 
the  same  number  of  towers,  the  total  plate  area  required  is  less  for  the 


V/F 


1.0 
0.6 
06 

0.4 


Key 


S  -  Single  tower 
VR.  -  Vapor  reuse 
JJ  -  Split  tower 
(For  multiple  towers  ctmcte  " 
values  for  $  by  number 
of  towers) 


I 3 V 

oc, Relative  volatility 
FIG.  7-27.     Comparison  of  vapor  requirements. 

multieffect  system  than  for  the  modifications.  This  greater  efficiency 
is  due  to  the  more  efficient  utilization  of  the  heat  over  a  wide  tempera- 
ture range. 

ANALYTICAL  EQUATIONS 

For  a  few  special  cases  of  limited  applicability,  it  has  been  possible 
to  develop  mathematical  solutions,  some  of  the  most  useful  of  which  are 
considered  in  this  section. 

Total  Reflux.  Fenske  (Ref .  5)  developed  an  algebraic  method  of 
calculating  the  minimum  number  of  theoretical  plates  by  utilizing  the 
relative  volatility  together  with  the  fact  that  at  total  reflux  the  operat- 
ing line  becomes  the  y  »  x  diagonal.  Thus,  considering  the  two 


RECTIFICATION  OF  BINARY  MIXTURES  175 

components  x'  and  x"  and  starting  with  the  still, 


and  at  total  reflux  the  operating-line  equation  gives  yw  —  x\,  giving 


Continuing,  in  the  same  manner, 

(x'\        i       v       ^  (x'\ 

1-77  I    =  (an-i)(<Xn-2)  '  '  '  aiawi—} 

\X    /n  \X    /w 


In  some  cases  the  relative  volatility  does  not  vary  widely,  and  an 
average  value  for  the  entire  column  can  be  used,  giving 


where  N  is  the  number  of  plates  in  the  column.     This  can  be  changed  to 

N  =  10g  (*/'")*('"/«%  (7.52) 

log  aav  v         ' 

If  a  total  condenser  is  employed,  this  becomes 

N  log  (MWM* 

log  «av  V  ' 

or,  if  a  partial  condenser  equivalent  to  one  theoretical  plate  is  employed, 

N  +  2  .  ^  W/fWM*  (7.54) 

log  aav 

These  equations  offer  a  simple  and  rapid  means  of  determining  the 
number  of  theoretical  plates  at  total  reflux  and  avoid  the  necessity 
of  constructing  the  y,x  diagram.  Principally  an  arithmetic  average  of 
the  relative  volatility  at  the  temperature  of  the  still  and  the  top  of  the 
tower  is  employed.  This  is  a  satisfactory  average  only  if  the  relative 
volatility  is  reasonably  constant  over  the  concentration  region  involved. 
For  larger  variations  the  geometric  average  would  be  more  satisfactory, 
i.e.,  aav  =  -VatOiW)  or  still  better  aav  =  1  +  V(««  —  l)(«ir  —  1),  where 
at  is  the  relative  volatility  at  the  top  of  the  column.  In  the  cases  of 
abnormal  volatility  such  as  are  exhibited  ty  ethyl  alcohol  and  water, 


176  FRACTIONAL  DISTILLATION 

the  use  of  an  average  relative  volatility  is  not  satisfactory  over  an 
appreciable  concentration  range;  however,  the  equation  may  be 
applied  successively  to  small  concentration  ranges,  but  the  operation 
becomes  more  time-consuming  than  constructing  the  y}x  diagram  and 
stepping  off  the  plates.  The  use  of  the  y,x  diagram  has  the  advantage 
that  it  gives  a  picture  of  the  concentration  gradient,  and  after  the 
diagram  has  been  constructed,  the  number  of  theoretical  plates  for 
other  reflux  ratios  can  be  easily  determined.  This  method  is  not 
applicable  to  conditions  other  than  total  reflux. 

Minimum  Reflux  Ratio.  The  analytical  equation  for  this  case  can 
be  easily  derived  in  terms  of  the  coordinates,  yc  and  xe,  of  the  point  of 
contact  of  the  equilibrium  curve  and  the  operating  line. 


(7  w\ 

^        J 


(a-  l)(xc)(l  -XC)(XD-XW) 
Xp[l  +  (a  —  l)xe]  -  axc 


_  F(xc  -  XW)(XD  ~  ZF*)[l  +  («  -  l)xc]  /7  5?) 

(a  -  l)(xc)(l  -  £)(XD  -  Xw)  V  " 

when  Vn  ==  Vm   and   XD   and   xw   are    approximately  1.0  and  0.0, 
respectively, 

T7  £>[!  +  (a  -  l)xr]  _  F(l  +  (a  -  l)xF] 

Vmin  =  -  («  ---  " 

and 


These  equations  are  exact  for  the  case  of  constant  molal  flow  rates 
where  the  value  of  the  relative  volatility  is  taken  at  the  composition 
xc.  Equations  (7-55)  to  (7-57)  involve  so  many  factors  that  it  is 
usually  easier  to  obtain  the  minimum  reflux  ratio  values  graphically 
or  by  using  yc  and  xc  with  Eq.  (7-10). 

Theoretical  Plates  Required  for  Finite  Reflux  Ratios.  The  equation 
for  the  number  of  theoretical  plates  at  total  reflux  is  convenient  for 
mixtures  in  which  a  is  relatively  constant,  and  it  would  be  useful  to 


RECTIFICATION  OF  BINARY  MIXTURES 


177 


have  a  comparable  method  applicable  at  other  reflux  ratios.    Two 
basic  equations  were  used  in  the  total  reflux  derivation: 

fa)    .  an  fa)         and        fa)    =  fa) 

\JlB/n  \X*/n  \3/B/n  VWn+l 


The  first  of  these  two  equations  is  true  at  any  reflux  ratio,  but  the 
latter  is  true  only  at  total  reflux.  However,  new  variables  may  be 
defined  such  that 


and 


-A 

~     W 


\ 


(7-60) 


(7-61) 


It  is  obvious  that  the  equation  developed  from  these  will  be  identical 
in  form  with  Eq.  (7-52).     Thus,  above  the  feed 

\r    _i_  i   _  *°&  (XA/XB)I>(XB/XA)P 


log 


(7-62) 


It  remains  to  relate  the  new  variables  to  the  actual  compositions  and 
the  usual  operating  variables.  In  addition  it  would  be  convenient, 
but  not  necessary,  to  have  the  following  relation, 

xfA  +  x'B  =  1.0 

A  number  of  solutions  for  the  new  variables  have  been  obtained 
(Refs.  16,  18),  and  one  set  is  given  here. 
Above  the  feed, 


o 


V 


8*0 
aV 


(*--7hnh)        (7'63) 


Below  the  feed, 


aV 


^  /     (       MW     xw   \ 
"__  1        \XA»  +  0  1$^  \) 


(7-64) 


Above  the  feed  plate, 


s  -  i  - 


=  1  —  x' 

i        xAm 

2(D/O)xp(cl  -  1) 


'  (<*(V/0)  -  1  -  (D/Q)xD(a  -  I)]2  | 
(7-65) 


178  FRACTIONAL  DISTILLATION 

Below  the  feed  the  same  equation  is  employed  to  calculate  the  value 
of  8  except  that  —  (Wxw/0m)  is  used  instead  of  DxD/On.  The  value 
of  An  is  calculated  as  follows: 

^«  =  ™  (7-66) 

These  equations  are  rather  involved  to  use,  and  they  become  par- 
ticularly difficult  where  the  number  of  plates  is  large.  The  most  effort 
is  involved  in  evaluating  the  term  /S,  and  Fig.  7-28  is  arranged  to 
facilitate  this  calculation. 

The  equation  above  the  feed  plate  should  be  employed  down  to  the 
feed-plate  conditions;  i.e.,  for  minimum  plates  at  a  given  0/D,  it 
should  be  applied  down  to  the  conditions  given  by  the  intersection  of 
the  operating  lines. 

For  special  conditions  found  to  apply  in  a  great  mg/ny  design  prob- 
lems, it  is  possible  to  simplify  the  calculation  of  S.  It  is  often  found 
that  the  value  of  S  below  the  feed  plate  is  equal  to  a(F/0),  which 
makes  the  calculation  of  S  simple.  It  can  be  shown  that  this  is  the 

W 
exact  value  of  S  for  the  case  of  -^-  xw(a  —  1)  equal  to  zero.     For 

W 
values  of  -^  xw(a  —  1)  that  are  less  than  1  per  cent  of  a(V/O)  —  1, 

good  results  are  obtained  by  using 

(V\       W 
~J+-~^(«-  1) 

and 

w 


-»  (7-68) 

Above  the  feed  plate  similar  approximations  can  be  made  giving, 
S  =  «[l  +  ^  (1  -  xAO)(a  -  1)]  (7-69) 


«  -1)] 


=  l+2(l  -**„)(«  -1)  (7-70) 

Lewis  Method.  Lewis  (Ref.  8)  expresses  the  rate  of  increase  of 
concentration  of  the  liquid  in  the  column  from  one  plate  to  the  next  by 
the  differential  dx/dn;  therefore, 

Xn  +  (7-71) 


RECTIFICATION  OF  BINARY  MIXTURES 


179 


FIG.  7-28.     Plot  for  analytical  equation. 


180  FRACTIONAL  DISTILLATION 

The  material-balance  equation 


yn  =  xn+i  +  ^  (XD  -  yn) 


can  be  written,  using  V  =  0  +  D, 

0 
From  Eq.  (7-71), 

dn  1 


dx  Xn+l    -  Xn  _  _  D 


(7-72) 


n   -  Xn-          (XD    -  yn) 


The  integration  of  this  equation  gives  the  number  of  theoretical  plates 
between  the  x  limits  chosen;  thus,  the  number  of  theoretical  plates 
above  the  feed  Nn  is 

CXD  dx 

Nn  =  /  zn        ~ 

/  yn   —  Xn   —  ~fi  (XD   "~  y*) 

J  xf  V 

Similarly,  below  the  feed  the  theoretical  plates  Nm  become 


The  equilibrium  y}x  curve  gives  the  relation  between  yn  and  xn  and 
ym  and  xm-  thus  by  assuming  values  for  xn,  values  of  yn  can  be  obtained 
from  the  equilibrium  data,  and 

1 

yn   —  Xn   ~  0   (XD   —  2/n) 

can  be  calculated  and  plotted  vs.  the  assumed  values  of  x.  The  area 
under  such  a  curve  from  x/  to  XD  is  the  number  of  theoretical  plates  by 
the  Lewis  method. 

Although  this  method  is  based  on  a  continuous  change  of  concentra- 
tion instead  of  the  actual  step  wise  concentration  increase,  the  error 
involved  is  not  serious  when  the  change  in  concentration  per  plate  is 
small.  This  latter  condition  is  generally  true  when  the  number  of 
plates  involved  is  large. 

When  the  relative  volatility  is  reasonably  constant  over  the  range  of 
concentrations  involved,  the  equilibrium  curve  may  be  approximated 
by  the  relative-volatility  relation  (see  page  30) 


RECTIFICATION  OF  BINARY  MIXTURES  181 


and  the  foregoing  equation  may  be  integrated  directly,  avoiding  the 
necessity  for  the  graphical  integration.  Thus,  for  total  reflux  at  which 
D/0  =  0  and  (F  -  D)/(0  -  pF)  =  0,  the  foregoing  equations  give 

Nn  +  Nm  +  1  *  (_!_)  in  *5  (1—  *lY  (7-75) 

\Q!    -    I/          0V  \1    -   #/>/  V  J 

where  ATn  +  Nm  +  1  is  the  total  number  of  steps  including  the  still 
step  and  any  condenser  enrichment.  Noting  that  In  a  for  values  of  a 
near  to  1  is  approximately  equal  to  (a  —  1)  makes  the  foregoing  equa- 
tion approach  the  equation  for  total  reflux.  The  values  at  total 
reflux  were  calculated  for  an  example  with  a  equal  to  1.07,  XD  =  0.96, 
and  xw  =  0.033.  Equation  (7-75)  gave  97  theoretical  plates,  while 
Eq.  (7-53)  gave  94.  Equation  (7-53)  is  applicable  only  at  total  reflux, 
while  the  Lewis  equation  can  be  integrated  for  the  case  of  finite  reflux 
ratios  and  constant  relative  volatilities.  Thus  Eq.  (7-73)  can  be 
integrated  to  give 

Nn  .  i+jvao.  ln  r  A^JLA  (*L±*\\ 

Vb*  -  4ac      l\xf  +  A)  \x»  +  B/j 


where  a  =  1  —  a 

b  =  (a  -  1)[1  -  (D/0)xD]  +  <xD/0 
c  =  —(D/0)xD 


A  __  b  -  Vb*  -  4ac  6  +  V62  -  4ac 

A  _  ~—~2a  and         B  --  2^  - 

A  similar  equation  is  obtained  below  the  feed  with  —  W/Om  replacing 
D/0,  —(W/Om)xw  replacing  (D/0)xD,  and  using  x/  instead  of  XD  and 
xw  instead  of  x/. 

The  integrated  equations  are  most  useful  in  cases  where  the  number 
of  plates  is  large  and  the  graphical  method  would  be  long  and  tedious. 

Example  Illustrating  the  Use  of  Analytical  Equations.    As  an  example  of  the 
use  of  the  analytical  equations,  consider  the  benzene-toluene  rectification  problem 
of  page  122. 
Total  reflux  [Eq.  (7-53)]: 

Using  a  relative  volatility  of  2.5, 


loe  0.95 

iog 


N  «  5.4  olates 


182  FRACTIONAL  DISTILLATION 

Minimum  reflux  ratio  [Eq.  (7-56)): 

(9\  0.95[1  4  1.5(0.5)]  -  2.5(0.5) 

\DJmm  "  1.5(6:5)(0.5)  " 

Theoretical  plates  at  (0/D)  -  3.0: 
Above  the  feed  plate,  using  Fig.  7-28, 


(0.95)  (1.5)  -0.475 


Q2  »  2.5  (%)  -  3.333 
4-  0.475  ~  3.333 

- 


n  _ 
°'65 


0475 

S  -  3.333  -  1.65(0.475)  -  2.55 


1  -  0.9  -  0.1 


3  f  (1.5/2.5)  (2.1 

4  L      1.95  -  1 


.55)  0.95 


3(2.55  - 
1  -  0.357  =  0.643 


Below  the  feed  plate,  TF/0  -  0.2,  7/0  «•  0.8,  and  using  Fig.  7-28  and  the 
equations  gives  S  «  2.03,  A  =  2.06,  x^F  »  0.731,  x'BF  «  0.269,  a?^  —  0.086, 
and  a;^  -  0.914. 


W 

o.269 


Thus  the  total  number  of  steps  including  the  still  is  8.9,  corresponding  to  7.9 
theoretical  plates  plus  the  still.  This  compares  with  8H  plates  estimated  by  the 
stepwise  method  (page  123). 

The  values  of  S  and  A  can  be  evaluated  by  the  approximations,  thus, 
Above  the  feed  plate, 

S  -  2.5(1  4-  H(0.05)  (1.5)1  *  2.55 
A  -  2.5(M)[1  4-  ?i(0.05)  (1.5)]  -  1.95 

Below  the  feed  plate, 

S  -  2.5(«)  4  H  (0.05)  (1.5)  -  2.02 
A  -  2.5(#)  4-  %  (0.05)  (1.5)  -  2.03 

These  values  are  so  close  to  those  obtained  before  that  the  result  is  essentially 
the  same. 


RECTIFICATION  OF  BINARY  MIXTURES 


183 


In  certain  cases  the  terminal  compositions  are  small,  and  to  obtain 
any  accuracy  by  the  usual  y,x  method  requires  expanding  the  diagram. 
However,  in  most  of  these  cases,  the  number  of  steps  between  the 
operating  line  and  the  equilibrium  curve  can  be  easily  calculated, 
utilizing  the  fact  that  the  equilibrium  curve  and  the  operating  line  are 
straight  in  this  region.  The  equilibrium  curve  can  be  expressed  as 
y  «  Kx  and,  for  values  of  a; -near  zero,  K  is  equal  to  the  relative  vola- 
tility. The  operating  line  in  such  a  region  will  be  straight  for  most 
cases,  even  on  an  enthalpy  basis,  because  the  variation  of  composition 
is  so  small  that  it  makes  essentially  no  difference  in  the  enthalpy 
values. 

For  the  bottom  of  the  tower  when  the  two  lines  are  straight,  the 
number  of  steps  or  theoretical  plates  required  is  given  by 


In 


Nw  = 


(V/0)(K  - 


-'-  +  1 


(7-77) 


In  (VK/O) 


where  Nw  =  number  of  plates  including  plate  m,  but  excluding  the 

still 

x  =  mol  fraction  of  most  volatile  component 
This  relationship  is  similar  to  Eq.  (7-62)  for  small  values  of  Xw 
Above  the  feed  plate,  the  corresponding  relationship  for  a  column  and 
total  condenser  is  obtained  by  using  a  similar  analysis  for  the  least 
volatile  component, 


In 


Nn+    I    = 


0 - 


1  -  K 


(7-78) 


In  (0/K'V) 


where  Nn  =  plates  above  plate  n 

y,  x  =  mol  fractions  of  least  volatile  component 

Kr  =  equilibrium  constant  for  least  volatile  component 
Packed  Towers.  Packed  towers  can  be  used  in  fractional  distilla- 
tion as  well  as  bubble-plate  columns.  Instead  of  bubbling  through  a 
pool  of  liquid  as  in  a  bubble-plate  tower,  the  interaction  between 
vapor  and  liquid  can  be  obtained  by  causing  the  reflux  to  flow  over  the 
surface  of  the  packing  material  while  the  vapor  flows  up  through  the 
voids.  Thg  iffie  of  packed  towers isgenerally  limited  to  towers  of  small 
sizes  or  to  special  distillations  Well  an'MlB  tiUUUtiZLU'UllOft  olTaitrlFacid7 
TiTsmall  laboratory  and  pilot-plant  size,  the  packed  tower  necessary 


184 


FRACTIONAL  DISTILLATION 


for  a  given  separation  is,  in  general,  less  expensive  than  a  corresponding 
bubble-plate  tower;  in  large  diameter,  the  reverse  may  be  true.  Aside 
from  the  economic  aspects,  packed  towers  are  ^asily  constructed  and 
can  be  made  of  noncorrosive  refractory  earffienware.  glass,  and  carBon 

—     ,_    _      ,    .uu, ~.— '~«—~>~~.|  ,-  -.-— V.-..- »^,ww       j^^r*,^'*****™"-       "*••"*«•»»•»-.,<        ^^ 

as  well  as  the  usualmetals  used  in  bubble-plate  tower  construction. 
They  have  the  disadvantage  that  it  may  be  difficult  to  clean  the  tower 
without  completely  dismantling  the  unit,  and  often  they  channel 
badly;  i.e.,  the  liquid  and  vapor  segregate  from  each  other,  and  the 
efficiency  of  contact  between  them  is  poor.  The  packed  towers,  in 

general,  have  very  low  pressure  drops 
from  top  to  bottom  relative  to  an 
equivalent  bubble-plate  tower. 

For  tower  packing,  a  wide  variety 
of  materials  have  been  used,  e.g., 
coke,  stone,  glass,  earthenware,  car- 
bon and  metal  rings,  wood  grids,  jack 
chain,  carborundum,  and  metal  and 
glass  helices,  as  well  as  a  large  num- 
ber of  other  packings  including  many 
manufactured  packings  of  special 
shapes. 

Design  of  Packed  Tower.  Ide- 
ally, the  interaction  between  va- 
por and  liquid  in  packed  towers  is 
true  counter  current  rather  than  the 
stepwise-countercurrent  propess  of 
a  bubble-plate  tower  with  theoreti- 
cal plates.  Instead  of  finite  steps,  the  true  countercurrent  action 
should  be  treated  differentially.  Consider  the  schematic  drawing  of 
the  packed  tower  in  Fig.  7-29.  Let  0  be  the  mols  of  overflow,  V  the 
mols  of  vapor,  x  and  y  the  average  mol  fraction  in  the  liquid  and  vapor, 
respectively,  n  distance  above  and  m  distance  below  the  feed.  Focus- 
ing on  the  differential  section  dn,  V(dy/dri)  must  equal  0(dx/dri)  for 
each  component,  and  this  transfer  must  be  due  to  an  exchange  of 
components  back  and  forth  between  the  liquid  and  vapor.  This 
transfer  is  due  to  the  fact  that  the  vapor  and  liquid  at  a  given  cross 
section  are  not  in  equilibrium  with  each  other,  and  the  rate  of  transfer 
will  be  a  function  of  the  distance  from  equilibrium;  thus, 


•w 


FIG.     7-29.     Schematic    diagram    for 
packed  tower. 


•  dy 


=  kA(y*  -  t/)  -  k'A(x  - 


(7-79) 


RECTIFICATION  OF  BINARY  MIXTURES  185 

where  &,  kf  =  proportionality  constants 

A  =  area  per  unit  height 

y*  =  vapor  in  equilibrium  with  x 

x*  =  liquid  in  equilibrium  with  y 

Integration  of  Eq.  (7-79)  is  difficult  because  very  little  information  is 
available  on  the  values  of  and  the  factors  involved  in  kA.  Assuming 
that  this  product  is  constant  and  making  the  usual  simplifying  assump- 
tions, equation  can  be  arranged  as  follows: 

n  _  Is  [v=VT_dy__  =  0.    I*'-"     dx 

Ic  A     I         »   ti ^  ~—  ti        1f^  A     /          i  i*  -I.-  '1** 
A//I  jy^yff  y          y        K  /i  J  x**Xf'  **        •*" 

Below  the  feed, 

1^  A     i  tt^  —  11         Jf*  A     I  /   'T  — —  o* 

A/ /I    Jy—ys       y  y  ft/  /I    J  x=*XW      •"  * 

where  x/,  t//  are  the  liquid  and  vapor  compositions  in  the  tower  at  the 
level  at  which  the  feed  is  introduced,  and  x'w  is  the  liquid  concentra- 
tion at  the  bottom  of  the  packed  section. 
A  material  balance  above  the  feed  gives 


V  =  \?)*  +  \^)*»  <7-82) 

By  assuming  values  of  y,  values  of  x  can  be  calculated  by  Eq.  (7-82), 
and  these  used  in  Eq.  (7-80)  together  with  equilibrium  data  to  evaluate 
the  integrals.  A  similar  material  balance  below  the  feed  can  be  used 
with  Eq.  (7-81).  The  equilibrium  curve  and  the  material-balance 
equation  can  be  plotted  on  the  y,x  diagram,  and  x*  —  x  or  y  —  y* 
read  directly. 

Consider  the  separation  of  an  equimolal  mixture  of  benzene  and 
toluene  into  a  product  containing  95  mol  per  cent  benzene  and  a  residue 
containing  5  mol  per  cent  benzene.  An  0/D  of  3  will  be  used,  and  the 
feed  will  enter  heated  such  that  the  mols  of  vapor  above  and  below  the 
feed  are  the  same.  The  usual  simplifying  assumptions  will  be  made. 
The  equilibrium  curve  is  given  in  Fig.  7-30. 

The  operating  lines  are  identical  with  those  for  the  stepwise  dia- 
gram. The  vertical  distance  between  the  equilibrium  line  and  the 
operating  line  is  y*  —  y,  and  the  horizontal  distance  between  the 
operating  line  and  the  equilibrium  line  is  x  —  x*.  In  general,  the  inte- 
gration must  be  performed  graphically,  although  in  cases  where  the 
equilibrium  curve  can  be  expressed  as  an  algebraic  relation  between  y 


186 


FRACTIONAL  DISTILLATION 


and  x  the  integration  can  be  carried  out  algebraically,  but  the  resulting 
equations  are  often  complex  and  involved. 

I.Or 


Xw 


0.2 


0.4  0.6 

x 

FIG.  7-30. 


0.8 


1.0' 


From  Fig.  7-30,  values  of  y*  —  y  are  read  at  various  values  of  y. 
Such  values  are  tabulated  in  Table  7-8. 

TABLE  7-8 


y 

y*  -y 

1 

y*  -y 

0.615 

0.098 

10  2 

0.7 

0.100 

10.0 

0.8 

0.090 

11.1 

0.9 

0.043 

23.2 

0.95 

0.028 

35.7 

Values  of  l/(y*  —  y)  are  then  calculated 
and  plotted  vs.  y.  The  area  under  this  curve 
from  y  =  0.615  to  y  =  0.95  is  equal  to 
kAn/V]  so  that  if  kA  is  known,  n  can  be  cal- 
culated. The  values  l/(y*  —  y)  vs.  y  are 
plotted  in  Fig.  7-31.  The  value  of 
dy/(y*  —  y)  is  the  shaded  area,  which  is 
equal  to  5.35.  A  similar  procedure  is  used 
below  the  feed,  and  alternately  the  x  —  x*  values  may  be  used. 

The  difficulty  in  using  this  procedure  is  in  evaluating  kA  and  k'A. 


RECTIFICATION  OF  BINARY  MIXTURES  187 

There  is  a  real  question  as  to  whether  Eq.  (7-79)  is  a  proper  formula- 
tion for  the  rate  of  transfer.     This  doubt  arises  from  several  factors: 

1.  There  is  no  adequate  proof  that  the  rate  of  transfer  is  directly 
proportional  to  the  difference  in  concentration,  but  there  are  theoretical 
considerations  that  would  indicate  this  is  not  the  proper  driving  force. 

2.  The  distribution  and  ratio  of  the  liquid  and  gas  are  not  the  same 
at  all  points  of  a  given  cross  section  resulting  in  differences  of  concen- 
tration, and  it  is  doubtful  whether  a  rate  equation  of  this  type  is  satis- 
factory with  average  values. 

3.  The  wetted-area  factor  is  a  nebulous  quantity  because  all  the  area 
of  the  packing  is  not  wetted  and  because  the  wet  areas  are  not  all 
equivalent.     It  is  a  factor  that  varies  widely  with  the  system  being 
distilled  and  with  the  column  construction. 

As  a  result  of  these  factors,  no  satisfactory  correlations  have  been 
presented  for  the  kA  and  k'A.  If  data  on  a  tower  operating  under 
essentially  identical  conditions  to  the  one  being  designed  are  available, 
a  reasonable  extrapolation  can  usually  be  made.  However,  any  major 
extrapolation  will  make  the  results  questionable. 

A  number  of  attempts  have  been  made  to  relate  the  values  of  kA 
obtained  for  one  system  with  those  for  other  systems.  The  coefficients 
involve  resistances  to  mass  transfer  for  both  the  vapor  and  the  liquid 
phases,  and  it  has  been  customary  to  apply  relations  based  on  the 
Lewis  and  Whitman  two-film  theory.  It  is  doubtful  that  such  a 
theory  is  applicable  in  this  case  since  it  is  difficult  to  visualize  the  con- 
ditions inherent  in  this  theory  for  a  liquid  phase  in  a  packed  tower. 
Further  studies  of  the  mechanism  of  mass  transfer  between  the  vapor 
and  the  liquid  for  systems  approximating  the  conditions  in  a  packed 
tower  are  needed  to  furnish  a  sound  basis  for  correlating  the  over-all 
mass-transfer  coefficients. 

Height  Equivalent  to  a  Theoretical  Plate.  In  general,  the  graphical 
calculation  involved  in  the  design  of  a  packed  tower  is  more  tedious 
and  time  consuming  than  the  stepwise  procedure  used  for  plate  towers. 
Actually,  the  equilibrium  curve  and  operating  lines  on  the  ytx  diagram 
are  identical  for  the  two  cases,  and  one  of  the  most  common  methods 
of  designing  packed  towers  has  been  to  determine  the  number  of 
theoretical  plates  required  for  the  separation  by  the  usual  stepwise 
method  and  then  to  convert  to  the  height  of  the  corresponding  packed 
tower,  by  multiplying  the  number  of  theoretical  plates  by  the  height  of 
packing  equivalent  to  one  theoretical  plate.  This  is  abbreviated  to 
H.E.T.P.  (Ref.  13)  and  is  a  height  of  packing  such  that  the  vapor 
leaving  the  top  of  the  section  will  have  the  same  composition  as  the 


188  FRACTIONAL  DISTILLATION 

vapor  in  equilibrium  with  the  liquid  leaving  the  bottom  of  the  section. 
The  use  of  the  H.E.T.P.  substitutes  a  stepwise  countercurrent  proce- 
dure for  the  true  countercurrent  operation  and  is  therefore  theoretically 
unsound;  but  when  the  concentration  change  between  plates  is  small 
and  the  number  of  plates  is  large,  the  error  introduced  by  its  use  will 
be  small.  Values  of  H.E.T.P.  are  determined  experimentally  by  cal- 
culating the  number  of  theoretical  plates  necessary  to  be  equivalent  to 
some  actual  packed  tower;  the  height  of  the  packed  tower  divided  by 
the  number  of  theoretical  plates  is  then  the  H.E.T.P.  Values  of 
H.E.T.P.  will  be  considered  in  Chap.  17  on  Column  Performance 
(page  465). 

Height  of  a  Transfer  Unit.  Chilton  and  Colburn  (Ref.  3)  have 
proposed  that  the  integrals  of  Eqs.  (7-80)  and  (7-81)  be  termed  the 
number  of  transfer  units,  N.T.U.  The  height  of  the  packed  section 
divided  by  the  number  of  transfer  units  is  termed  the  height  of  a  trans- 
fer unit,  H.T.U.  This  latter  unit  being  defined  above  the  feed  plate  as 

TT  «T>  TT      _  U 

H.l.U.  -  j^rjr^ 

=  JL  =  J5L 

kA       k'A 

The  transfer  unit  consists  of  a  step  on  the  operating  line  such  that 
the  change  in  y  or  x  is  equal  to  the  average  difference  between  the 
equilibrium  curve  and  the  operating  line  over  the  region  of  the  step. 
If  the  equilibrium  and  operating  lines  are  parallel,  the  step  will  be 
exactly  equal  to  that  for  a  theoretical  plate,  and  the  H.T.U.  value  will 
equal  the  H.E.T.P.  value.  If  the  slope  of  the  equilibrium  curve  is  less 
than  that  of  the  operating  line,  the  two  curves  will  tend  to  converge 
with  increasing  concentration,  and  the  initial  difference,  which  corre- 
sponds to  the  step  taken  for  a  theoretical  plate,  will  be  greater  than  the 
average  difference  corresponding  to  the  transfer  unit.  Thus,  in  this 
case  one  equilibrium  plate  will  give  a  greater  concentration  change 
than  one  transfer  unit.  If  the  two  curves  diverge  with  increasing 
concentration,  the  reverse  is  true. 

Over  the  height  of  one  transfer  unit,  the  value  of  y*  —  y  does  not 
ordinarily  vary  widely,  and  the  arithmetic  average  may  be  used. 
Baker  (Ref.  1)  has  developed  a  relatively  simple  stepwise  method  for 
estimating  the  number  of  transfer  units  under  these  conditions.  In 
the  usual  y,x  diagram  (Fig.  7-32),  a  line  ab  is  drawn  at  the  arithmetic 
mean  of  the  equilibrium  curve  and  the  operating  line.  The  H.T.U. 
corresponds  to  a  step  giving  a  change  in  y  equal  to  the  average  value  of 


RECTIFICATION  OF  BINARY  MIXTURES 


189 


*  —  y  over  the  step.  Starting  at  A,  one  proceeds  not  to  B  but  to  C, 
uch  that  AD  =  DC,  and  then  steps  from  C  to  the  operating  line, 
f  the  curvature  of  the  equilibrium  curve  is  not  too  great,  C(r,  the 
hange  in  y  which  is  numerically  equivalent  to  the  EF,  will  be  approxi- 
nately  equal  to  the  average  value  of  y*  —  y  between  A  and  (?;  there- 
ore,  the  steps  correspond  to  one  transfer  unit.  This  stepwise  proce- 
lure  is  continued  to  the  terminals  of  the  tower,  giving  the  number  of 
/ransfer  units  in  the  tower.  Values  of  H.T.U.  are  multiplied  by  the 
lumber  of  transfer  units  required  to  determine  the  height  of  packing 
lesired. 


FIG.  7-32.     Diagram  for  H.T.U. 

Because  the  transfer  unit  is  defined  on  a  differential  countercurrent 
basis,  it  is  usually  assumed  to  be  more  correct  for  the  design  of  packed 
towers  than  the  stepwise  countercurrent  procedure  of  the  theoretical 
plate.  This  is  probably  true,  but  there  is  a  serious  question  whether 
the  transfer  unit  is  on  a  sound  theoretical  basis.  Most  distillation 
operations  are  of  a  degree  that  requires  a  number  of  theoretical  plates, 
and  for  suph  cases  it  is  doubtful  whether  at  the  present  stage  of  devel- 
opment the  transfer-unit  concept  has  any  advantage  over  the  the- 
oretical-plate basis  for  the  design  of  packed  towers.  The  latter  is 
easier  to  employ. 

In  practice  the  packed  tower  has  been  losing  out  relative  to  the 
bubble  tower.  The  development  of  corrosion-resistant  alloys  and  of 
bubble-plate  columns  made  of  ceramic,  glass,  and  plastic  has  made  it 
possible  to  rectify  corrosive  mixtures  in  such  units.  The  development 
of  efficient  laboratory  bubble-plate  columns  as  small  as  1  in.  in  diam- 
eter has  made  it  possible  to  carry  out  such  distillation  in  the  laboratory, 
and  experience  has  indicated  that  these  columns  give  data  that  are 


190  FRACTIONAL  DISTILLATION 

much  more  suitable  for  extrapolation  in  the  design  of  large-scale  bub- 
ble-plate towers  than  those  obtained  in  packed  towers.  These  small 
columns  give  plate  efficiencies  that  are  comparable  to  those  of  large 
towers,  and  their  effectiveness  is  not  a  major  function  of  the  wetting 
characteristics  of  the  liquid  as  it  is  in  a  packed  tower.  The  one  major 
remaining  advantage  of  the  packed  tower  is  low  pressure  drop;  even 
in  this  case,  new  types  of  columns  are  being  developed  which  give 
definite  and  controlled  contact  between  the  liquid  and  the  vapor  and 
which  have  pressure  drop  as  low  as  in  packed  towers.  If  large-sized 
packed  towers  can  be  developed  to  give  efficiencies  commensurate  with 
those  obtained  in  small  laboratory  packed  towers  and  if  their  design 
can  be  made  reliable  and  reproducible  such  that  their  performance  can 
be  predicted  with  reasonable  accuracy,  they  could  become  a  major 
factor  in  vapor-liquid  interchange  processes. 

Nomenclature 

a  •»  relative  volatility 
C  **  specific  heat 

D  «  mols  of  distillate  withdrawn  as  overhead  products  per  unit  of  time 
F  **  mols  of  mixture  fed  to  column  per  unit  of  time 
//  «  enthalpy,  or  heat  content  of  vapor 
h  «•  enthalpy,  or  heat  content  of  liquid 
H.K.T.P.  »  height  equivalent  to  a  theoretical  plate 
H.T.U.  ««  height  of  a  transfer  unit 

m  »  number  of  plate  under  consideration,  counting  up  from  still 
n  =  number  of  plate  under  consideration,  counting  up  from  feed  plate 
0  •»  total  mols  of  overflow  from  one  plate  to  next,  per  unit  of  time 
p  as  vapor  pressure 
p  -  (0/+1  -  0,)/P 
Q  w  heat  added  or  removed 
T,t,  »  temperature 
W  •»  mols  of  residue  per  unit  of  time 
x  «•  rnol  fraction  of  more  volatile  component  in  liquid 
re7  «•  pseudo  mol  fraction  in  liquid 
y  «  mol  fraction  of  more  volatile  component  in  vapor 
yf  **  pseudo  mol  fraction  in  vapor 
z  «•  mol  fraction  in  feed  mixture 
y  a«  activity  coefficient 

Subscripts 

n  refers  to  nth  plate;  *".«.,  0«  and  Vn  refer  to  mols  of  liquid  and  vapor  leaving  nth 

plate,  respectively 
m  refers  to  wth  plate 
/  refers  to  feed  plate;  i.e.,  Xf  is  mol  fraction  of  more  volatile  component  in  overflow 

from  feed  plate 
F  refers  to  feed;  i.e.,  xp  is  mol  fraction  of  more  volatile  component  in  feed  mixture 


RECTIFICATION  OF  BINARY  MIXTURES  191 

D  refers  to  distillate 
R  refers  to  reflux 
W  refers  to  bottoms 
L  refers  to  liquid 
F  refers  to  vapor 
t  refers  to  top  plate 

References 

1.  BAKER,  Ind.  Eng.  Chem.,  27,~  977  (1935). 

2.  BOSNJAKOVIC,  "Technische  Thermodynamik  II,"  diagrams,  T.   Steinkopf, 
Dresden,  1937. 

3.  CHILTON  and  COLBURN,  Ind.  Eng.  Chem.,  27,  255,  904  (1935). 

4.  DODGE,  Chem.  Met.  Eng.,  35,  622  (1928). 

5.  FENSKE,  Ind.  Eng.  Chem.,  24,  482  (1932). 

6.  GUNNESS,  Sc.D.  thesis  in  chemical  engineering,  M.T.T.,  1936. 

7.  KEESOM,  Bull.  Intern.  Inst.  Refrig.,  15  (1934). 

8.  LEWIS,  Ind.  Eng.  Chem.,  14,  492  (1922). 

9.  LEWIS,  "Unit  Operation  Notes/'  M.I.T.,  1920. 

10.  McADAMS,  "Heat  Transmission/'  2d  ed.,  McGraw-Hill  Book  Company,  Inc., 
New  York,  1942. 

11.  McCABE  and  THIELE,  Ind.  Eng.  Chem.,  17,  605  (1925). 

12.  OTHMER,  Ind.  Eng.  Chem.t  28,  1435  (1936). 

13.  PETERS,  Ind.  Eng.  Chem.,  14,  476  (1922). 

14.  PONCHON,  Tech.  moderne,  13,  20  (1921). 

15.  SAVARIT,  Arts  et  metiers,  pp.  65,  142,  178,  241,  266,  307  (1922). 

16.  SMOKER,  Trans.  Am.  Inst.  Chem.  Engrs.,  34,  165  (1938). 

17.  SOREL,  "La  rectification  de  Falcool,"  Paris,  1893. 

18.  UNDERWOOD,  /.  Inst.  Petroleum,  29,  147  (1943). 

19.  WALKER,  LEWIS,  MCADAMS,  and  GILLILAND,  "Principles  of  Chemical  Engi- 
neering/' 3d  ed.,  McGraw-Hill  Book  Company,  Inc.,  New  York,  1937. 


CHAPTER  8 
SPECIAL  BINARY  MIXTURES 

This  chapter  covers  special  fractional  distillation  systems  for  binary 
mixtures.  The  examples  illustrate  the  flexibility  of  the  fractional  dis- 
tillation process  and  the  broad  applicability  of  the  design  methods. 
The  first  section  will  consider  operating  conditions  that  lead  to  unusual 
operating  lines,  and  the  last  section  will  consider  the  separation  of 
binary  azeotropic  mixtures. 

Special  Operating  Lines.  The  operating  lines  so  far  considered 
intersected  the  y  ~  x  line  at  values  between  x  =  0  and  x  =  1.0  and, 
with  the  exception  of  the  bottom  portion  of  the  lower  operating  line 
for  the  steam  distillation  case,  they  were  between  the  equilibrium  curve 
and  the  y  =  x  diagonal.  Although  most  operating  lines  lie  in  this 
region,  it  is  not  necessary  that  they  do  so.  The  operating  line  is  a 
combination  of  the  over-all  material  balance  and  a  component  balance. 
Consider  the  case  of  the  section  above  the  feed  plate  : 

By  over-all  balance, 

Vn   =   On+1  +  R 

By  component  balance, 


Vnyn  =  On+iXn+i 

where  R  is  the  net  molal  withdrawal  from  the  section  other  than  in  Vn 
and  On+i,  and  RxR  is  the  net  molal  withdrawal  of  the  component  from 
the  section  other  than  in  Fn  and  On+i> 

In  the  cases  considered  in  Chap.  7,  it  was  assumed  that  -there  was  a 
net  withdrawal  at  the  top  of  the  column  equal  to  Z),  but  it  may  be  that 
material  is  also  added  to  the  section  such  that  R  is  positive,  zero,  or 
negative.  If  R  is  positive,  On+i/Vn  will  be  less  than  1.0;  if  R  =  0, 
On+i/Vn  will  equal  1.0;  if  R  is  negative,  On+i/Vn  will  be  greater  than 
1.0.  In  these  cases  XR  may  be  positive  or  negative.  The  following 
example  will  illustrate  this  case. 

Leaky  Condenser  Example.  An  old  ethyl  alcohol  distillation  column  has  been 
tested,  and  it  is  concluded  that  there  is  a  leak  into  the  condensate  in  the  condenser 
that  amounts  to  5  mols  of  water  per  100  mols  of  feed.  The  tower  is  operating  on  a 
feed  containing  3.5  mol  per  cent  ethanol  and  96.5  mol  per  cent  water  and  produces 

192 


SPECIAL  BINARY  MIXTURES 


193 


a  distillate  containing  70  mol  per  cent  ethanol  and  a  bottoms  containing  0.001  mol 
per  cent  ethanol.    The  feed  enters  preheated  such  that  Vn  **  Fw.     Make  the  usual 
simplifying  assumptions. 
Solution.     As  a  basis  take  100  mols  of  feed.     Then,  by  over-all  material  balance, 


By  alcohol  balance, 


For  the  upper  section, 


D  +  W  -  105 

0.7D  +  0.00001  W  =  3.5 

•      D  -  5.0 
TF  -  100 

5  +  Fn  -  On4.!  -f  D 

#  =o" 


The  alcohol-enriching  line  is 

Vny«  »  On+lxn+l  H- 

»    On+l*»+I    +  3.5 

Since  Vn  =  On+i,  this  line  intersects  the  y  «  x  line  at  «  •*  eo  and  is  parallel  to  the 
diagonal.  With  Vn  =  Fm,  this  line  will  intersect  the  lower  operating  line  at  the 
usual  value,  i.e.,  at  x  =  0.035. 


0.1     0.2     03     0.4     05     06     0.7     0.8     0.9     1.0 
Mol  fraction  ethanol  in  liquid 

FIG.  8-1. 

A  possible  position  for  the  operating  line  is  shown  in  Fig.  8-1.  (The  exact  posi- 
tion is  not  known  since  the  reflux  ratio  was  not  specified.)  While  the  line  extends 
outside  of  the  diagram,  it  is  utilized  only  in  the  region  below  the  y,x  equilibrium 
curve.  The  composition  of  the  top  vapor  and  the  liquid  reflux  corresponds  to  a 
point  on  the  operating  line,  and  the  step  for  the  top  plate  should  terminate  at  this 
point. 


194  FRACTIONAL  DISTILLATION 

For  the  case  just  considered,  R  *=  D  —  leak.  If  the  leak  is  less  than  Z>,  R  would 
be  positive,  0/V  would  be  less  than  1.0,  and  the  intersection  with  y  **  x  would  be 
positive  and  greater  than  0.7.  As  the  size  of  the  leak  increased  relative  to  £>, 
0/V  would  become  nearer  to  1.0  and  the  intersection  would  increase  and  tend  to 
infinity.  If  the  leak  became  greater  than  D,  R  would  become  negative,  0/V 
would  be  greater  than  1.0,  and  the  intersection  with  the  y  *•  x  line  would  be 
negative. 

It  appears  that  the  leak  makes  the  slope  of  the  upper  operating  line  more  desir- 
able (i.e.,  nearer  to  1.0),  but  a  little  consideration  will  show  that,  when  this  line 
keeps  the  same  intersection  with  the  lower  operating  line,  the  smaller  the  value  of 
(0/V)n,  the  fewer  the  plates  required.  Common  sense  also  indicates  that  the  leak 
is  undesirable. 

In  this  example  the  operating  lines  were  above  the  y  =  x  diagonal 
but  had  intersections  with  the  diagonal  outside  of  the  usual  diagram. 
The  following  example  illustrates  a  case  for  which  the  operating  lines 
are  also  below  the  diagonal. 

Isopropyl  Alcohol  Stripping  Example.  In  the  manufacture  of  isopropyl  alcohol, 
propylene  is  dissolved  in  sulfuric  acid  to  give  an  extract  of  mono-isopropyl  sulfate. 
To  avoid  excessive  decomposition  this  extract  must  be  diluted  until  the  ratio 

Lb.H2S04 


Lb.  H20  +  lb.  H2S04 

before  the  alcohol  can  be  distilled.  In  this  ratio  the  H2S04  is  the  total  acid, 
whether  free  or  combined  with  propylene,  and  H2O  is  the  total  water,  free  or  com- 
bined with  propylene  as  isopropyl  alcohol.  A  typical  extract  containing  H2S04, 
propylene,  and  H20  in  equimolal  proportions  is  first  diluted  with  water,  and  is  then 
stripped  of  isopropyl  alcohol.  Usually  live  steam  is  used  in  the  stripping,  and  this 
results  in  dilute  alcohol  and  very  dilute  bottom  acid  which  must  be  reconcentrated 
The  stripped  acid  contains  negligible  alcohol. 

It  has  been  proposed  to  modify  this  operation  by  diluting  the  extract  in  the 
stripping  tower  with  the  overflow  and  thereby  produce  more  concentrated  alcohol 
and  at  the  time  obtain  45  per  cent  II 2804  as  bottoms.  To  avoid  polymerization 
during  dilution,  it  is  estimated  that  the  overflow  with  which  the  extract  is  mixed 
must  not  contain  over  10  mol  per  cent  isopropyl  alcohol.  Isopropyl  alcohol  and 
water  form  an  azeotrope  containing  68  mol  per  cent  alcohol.  The  accompanying 
diagram  (Fig.  8-2)  shows  the  flow  sheet  for  the  proposed  modification  to  produce  a 
66  mol  per  cent  alcohol  product.  The  tower  will  operate  with  live  steam  and  a 
total  condenser.  It  is  assumed  that  the  feed  will  enter  such  that  Vn  «  Vm,  and 
the  usual  simplifying  assumptions  will  be  made.  It  is  assumed  that  sulfuric  acid  is 
nonvolatile,  and  the  calculations  will  be  made  on  a  ' '  sulfuric  acid-free  "  basis.  It  is 
also  assumed  that  all  of  the  propylene  is  in  the  solution  as  isopropyl  alcohol  and 
that  the  vapor-liquid  equilibria  for  the  isopropyl  alcohol-water  system  will  be  used. 

Solution.    Basis :  1  mol  of  propylene  in  feed  (also  1  mol  each  of  HaO  and 

By  a  sulfuric  acid  balance  the  water  in  the  bottom  equals 

98/0.55 


SPECIAL  BINARY  MIXTURES 


195 


Less  than 


f  r 

fresh      W 
extract        Mixer 

j 

1 

i 

i 

66% 
isopropanol 


45% 


FIG.  8-2.     Isopropyl  alcohol  distillation  unit. 

Uncombined  water  in  distillate  =  ^!~  =  0.515  mol 

O.oo 

Combined  water  as  isopropyl  alcohol  in  distillate  —  1.0  mol 
Water  with  feed  »  1.0  mol 

Mols  of  water  added  as  steam  S  =  6.65  +  1.515  —  1.0  «  7.165  mols. 
Operating  lines: 

By  an  alcohol  balance, 


Vnya    «  On+lXa 

D          l 


On+iXa   -f   1.0 


1.515 


'  0.66 

Vn  =  Vm  -  S  -  7.165 
Om  SB  6.65  (sulfuric  acid-free  basis) 
On  -  5.65 


Operating  line  above  feed  plate, 

5.65         ,     1.0 


"  7.165" 


7.165 


=  0.789z0  +  0.139 


This  line  crosses  the  y  »  x  line  at  0.66  =  XD. 
Operating  line  below  feed  plate, 


Wxw 


The  two  operating  lines  intersect  at  y  »  0.928,  x  «  1.0.  These  lines  are  shown 
in  Fig.  8-3.  The  equilibrium  curve  shown  is  for  water-isopropanol  and  should  be 
suitable  above  the  feed  plate;  below  the  feed  plate,  with  the  sulfuric  acid  present 
the  separation  should  be  easier  than  shown.  The  steps  on  the  diagram  would  start 
at  XD'  "•  0.66  and  continue  down  the  upper  operating  line  until  the  mol  fraction  of 
alcohol  in  the  liquid  is  less  than  0.10;  then  the  shift  to  the  lower  operating  line  would 


196 


FRACTIONAL  DISTILLATION 


be  made  and  the  step  wise  procedure  carried  on  to  the  bottom.  In  the  diagram  as 
drawn,  the  first  plate  with  an  overflow  containing  less  than  10  mol  per  cent  alcohol 
has  a  concentration  of  5  per  cent  alcohol.  This  would  be  mixed  with  the  fresh 
feed,  and  the  mixture  would  be  the  liquid  to  the  feed  plate.  As  drawn  in  Fig.  8-3, 
the  liquid  leaving  the  feed  plate  contains  approximately  1  mol  per  cent  alcohol, 
and  from  this  concentration  on  down,  the  steps  would  be  made  between  the  equilib- 
rium curve  and  the  lower  operating  line  which  is  below  the  y  =  x  line. 


0      01      02     03     04     05     06     07     08     09      10 
Mol  fraction  of  isopropanol  in  liquid,  sulfunc-acid  -free  basis 

FIG.  8-3.     y,x  diagram  for  isopropyl  alcohol  example. 

In  this  case  the  diagram  is  unusual  in  two  respects:  (1)  The  distillate  has  a  lower 
ratio  of  alcohol  to  water  than  the  feed  and  (2),  the  feed-plate  composition  is 
purposely  fixed  at  a  value  different  from  that  corresponding  to  the  fewest  plates. 

Separation  of  Binary  Azeotropic  Mixtures.  A  large  number  of  two- 
component  systems  form  azeotropic  mixtures,  and  it  is  frequently 
necessary  to  separate  them  into  their  components.  Regular  frac- 
tional distillation  will  not  separate  such  mixtures  into  the  components 
in  high  purity,  but  by  suitable  modifications  it  is  frequently  possible  to 
obtain  the  desired  separation.  At  the  azeotropic  composition  the 
relative  volatility  is  unity,  and  rectification  is  not  possible.  The 
methods  employed  for  separating  such  systems  involve  using  either  (1) 
distillation  plus  other  separation  processes  to  get  past  the  azeotropic 
composition  or  (2)  a  modification  of  the  relative  volatility. 

1.  Distillation  plus  Other  Separation  Processes.  The  techniques 
most  commonly  employed  for  moving  by  the  azeotropic  composition 
are  decantation,  extraction,  crystallization,  or  absorption.  A  compo- 


SPECIAL  BINARY  MIXTURES 


197 


sition  near  the  azeotrope  can  be  produced  by  distillation  and  then 
separated  into  two  fractions  having  compositions  on  each  side  of  the 
azeotrope  by  one  of  these  techniques.  These  two  fractions  can  be 
fractionally  distilled  separately  to  give  the  essentially  pure  components 
and  fractions  of  approximately  azeotropic  composition  which  can  be 
recycled  through  the  process. 

The  method  of  utilizing  Recantation  to  aid  in  the  separation  of  the 
azeotrope  is  illustrated  by  the  system  phenol-water  which  at  atmos- 
pheric pressure  forms  an  azeotrope  containing  1.92  mol  per  cent 


Greater  than  1.68% 
phenol  — 

Less  than  192% 
phenol 


JJ/%  phenol 
168%  phenol 


*  Water  (low  in  phenol) 
FIG.  8-4.     Fractionating  system  to  concentrate  dilute  aqueous  phenol. 

phenol.  If  a  mixture  of  phenol  and  water  is  subjected  to  a  continuous 
f  ractionation,  the  overhead  will  tend  to  the  azeotropic  composition  and 
the  bottoms  will  approach  either  phenol  or  water  depending  on  the 
composition  of  the  feed.  Thus,  if  a  feed  containing  1  mol  per  cent 
phenol  is  given  such  a  distillation,  an  overhead  product  approaching 
the  azeotropic  composition  can  be  made  with  water,  low  in  phenol,  as 
bottoms.  In  this  case,  the  azeotrope  can  be  separated  because  phenol 
and  water  are  only  partly  miscible.  For  example,  at  20°C.  the  two 
saturated  liquid  phases  contain  1.68  and  33.1  mol  per  cent  phenol, 
respectively.  Thus  if  the  fractionation  gives  an  overhead  vapor  con- 
taining more  than  1.68  mol  per  cent  phenol,  it  will  break  into  two 
liquid  layers  on  cooling  to  20°C.,  the  water  layer  containing  1.68  mol 
per  cent  phenol  can  be  used  as  reflux,  and  the  phenol  layer  can  be 
withdrawn  as  product.  This  system  of  fractionation,  cooling,  and 
decantation  is  shown  in  Fig.  8-4.  Such  a  system  will  separate  dilute 
solutions  of  phenol  into  water  and  33  per  cent  phenol.  The  conden- 
sate  can  be  cooled  to  temperatures  other  than  20°C.,  but  for  the  system 
to  be  effective  two  liquid  phases  must  be  formed,  one  of  which  must 


198 


FRACTIONAL  DISTILLATION 


have  a  composition  less  than  the  azeotrope  and  the  other  a  composition 
greater.  The  system  of  Fig.  8-4  does  not  give  complete  separation  of 
the  two  components,  but  by  a  similar  arrangement  concentrated  solu- 
tions of  phenol  can  be  fractionated  to  give  an  overhead  condensate  that 
will  separate  into  two  liquid  phases.  Figure  8-5  shows  a  schematic 
flow  sheet  for  a  feed  containing  50  mol  per  cent  phenol.  The  column 
operates  to  produce  an  overhead  vapor  containing  between  1.92  and 
33.1  per  cent  phenol  and,  on  cooling  to  20°C.,  the  condensate  would 
give  the  same  liquid  compositions  as  obtained  in  Fig.  8-4.  In  this 
case  the  water  layer  would  be  the  product  and  the  phenol  layer  would 
be  employed  as  reflux. 


Less  than  53 1  % 
phenol  —  • 

Greater  than  132% 
phenol 


168%  phenol 
35.1%  phenol 


FIG.  8-5. 


50%phenol 
50%  wafer 


Phenol  (low  in  water) 
Fractionating  system  to  dehydrate  aqueous  phenol. 


The  systems  shown  in  Figs.  8-4  and  8-5  can  be  combined  to  give 
complete  separation.  The  arrangement  of  Fig.  8-6  shows  a  two-tower 
system  for  a  dilute  phenol  feed.  Both  towers  give  the  same  con- 
densate compositions,  and  the  overhead  vapors  are  fed  to  the  same 
condenser  and  decanter.  Tower  1  operates  in  the  same  manner  as 
the  single  tower  of  Fig.  8-4  producing  an  overhead  product  containing 
33.1  per  cent  phenol  and  a  bottoms  low  in  phenol.  Tower  2  is  only 
a  stripping  section  because  the  reflux  stream  is  its  only  feed.  If  the 
fresh  feed  contains  less  than  1.68  per  cent  phenol,  it  should  be  intro- 
duced into  tower  1;  from  1.68  to  33.1  per  cent  phenol,  it  should  be 
added  to  the  decantation  system;  if  the  feed  contains  more  than  33.1 
per  cent  phenol,  it  should  be  introduced  into  tower  2.  The  arrange- 
ment can  give  a  high  degree  of  separation  between  phenol  and  water. 
The  quantitative  calculations  for  such  a  system  are  illustrated  by  the 
following  example. 


SPECIAL  BINARY  MIXTURES 


199 


Phenol- Water  Rectification  Example.  A  plant  using  phenol  as  a  solvent  desires 
to  rectify  a  mixture  containing  1.0  mol  per  cent  phenol  in  water.  As  the  bottoms 
are  to  be  discarded  in  a  nearby  river,  they  must  not  exceed  0.001  mol  per  cent 
phenol.  A  system  similar  to  that  shown  in  Fig.  8-6  will  be  employed.  The  over- 
head vapors  from  the  bubble-plate  columns  to  be  used  are  condensed  and  cooled  to 
20°C.  i  Under  these  conditions  the  condensate  separates  into  two  saturated  layers. 
The  water  layer  is  reheated  to  its  boiling  point  and  refluxed  to  column  1,  and  the 


Water  Phenol 

FIG.  8-6.     Two-tower  fractionating  system  to  separate  aqueous  phenol. 


phenol  layer  is  reheated  and  sent  to  stripping  tower  2,  to  recover  a  99.99  per  cent 
phenol  as  bottoms.  At  20°C.  the  saturated  water  layer  contains  1.68  mol  per  cent 
phenol,  and  the  saturated  phenol  layer  contains  33-ljnol  per  cent  pEenol.  Assum- 
ing that  Vn  ~  Vm  and  making  the  usual  simplifying  assumptions,  determine: 

1.  The  minimum  mols  of  vapor  per  100  mols  of  feed  for  each  tower. 

2.  The  number  of  theoretical  plates  required,  using  a  vapor  rate  %  times  the 
minimum  rate. 

3.  The  minimum  number  of  plates  at  total  reflux.     (Only  the  water  layer  is 
refluxed  to  the  water  tower.) 

Solution  of  Part  1.     Basis :  100  mols  of  fresh  feed.     Referring  to  Fig.  8-6, 

1  -  0.00001  Wi  +  0.9999 W 2 
W i  +  W*  -  100;        Wz  -  1.0;        Wl  -  99.0 

Operating  lines  for  tower  1 : 

Above  the  feed  plate,  a  balance  between  plates  n  and  n  -f  1  and  around  tower  2, 

Vn  -  ft,*,  -f  W* 

+  0.9999JT , 

This  line  intersects  the  y  «  x  line  at  x  «  0.9999. 


200 


FRACTIONAL  DISTILLATION 


Below  the  feed  plate, 

Om+l  -  Fm  +  Wi 
Vmym  -  0»+iafcH4  -  0.00001  Wi 

which  intersects  the  y  «  x  diagonal  at  x  «•  0.00001. 

VAPOK-LIQUID  EQUILIBRIUM  DATA  FOR  PHENOL- WATER  AT  1  ATM. 

(Ref.  5) 


X 

y 

x 

y 

0 

0 

0.10 

0.029 

0.001 

0.002 

0.20 

0.032 

0.002 

0.004 

0.30 

0.038 

0.004 

0.0072 

0.40 

0.048 

0.006 

0.0098 

0.50 

0.065 

0.008 

0.012 

0.60 

0.090 

0.010 

0.0138 

0.70 

0.150 

0.015 

0.0172 

0.80 

0.270 

0.017 

0.0182 

0.85 

0.370 

„  0.018 

0.0186 

0.90 

0.55 

rO.019 

0.0191 

0.95 

0.77 

0.020 

0.0195 

1.00 

1.00 

Minimum  vapor  rate  for  tower  1  : 

The  minimum  vapor  corresponds  to  the  minimum  reflux  ratio,  and  there  are  two 
possibilities  in  this  case:  (1)  the  pinched-in  region  could  occur  at  the  feed  plate, 
or  (2)  it  could  occur  at  the  top  of  the  tower. 

If  the  feed  plate  is  the  limiting  condition,  then 


0.0138(Fn)mm 
(F.U, 


0.010n+i  +  0.9999TF2 
0.01  Vn  +0.9899TT2 


260TF,  -  260 


(0.0138  is  the  vapor  in  equilibrium  with  a  liquid  of  0.01). 

If  the  top  plate  is  the  limiting  condition,  then  the  concentration  on  this  plate 
must  be  equal  to  the  reflux  concentration,  and  the  top  vapor  will  be  in  equilibrium 
with  this  composition,  thus, 


0.0181  (Vn)mm 


0.01680n+i  +  0.9999TF2 
•  0.0168  Vn  +  0.9831  JF2 
0.9831TF2 


7K- 
755 


Since  the  top  condition  requires  more  vapor,  it  is  the  limiting  condition  and 

(Fn)min   -  755. 

Minimum  vapor  rate  for  tower  2: 

Tower  2  will  pinch  at  top,  the  liquid  on  the  top  plate  will  have  a  composition  of 
0.331,  and  the  equilibrium  vapor  composition  will  equal  0.0403. 


SPECIAL  BINARY  MIXTURES  201 

0.0403(F)min  -  0.3310  -  0.9999JF2 
-  0.331  F  ~  0.6689  JF2 


Solution  of  Part  2 

Tower  1: 

(Fn)aet  -  ^(755)  -  1,007;        0»+i  -  1,006 
Fw  -  1,007;        Om+i  «  1,106 

Above  the  feed  plate,          * 

l,007?/n  -  1,006*.+!  +  0.9999 
0.9999  -  yn  «  0.999(0.9999  -  sn+i) 

Below  the  feed  plate, 

l,007yw  -  l,10tew+i  -  0.00001(99) 
y«  -  0.00001  =  1.098(3m+i  -  0.00001) 
Tower  2: 

Fact  -  «(2.3)  -  3.06 


Operating  line  for  phenol, 

3.06?/m  =  4.06zm+i  -  0.9999 
Operating  line  for  water, 

3.06yM  »  4.06xw+i  -  0.0001 

These  operating  lines  are  plotted  in  Figs.  8-7  and  8-8.  Logarithmic  plotting  is 
used  to  facilitate  the  calculations  at  the  low  concentrations.  In  the  case  of  tower 
1,  the  operating  lines  for  phenol  are  used;  for  tower  2,  the  operating  line  for  water 
is  used. 

According  to  the  diagram,  tower  1  requires  a  still  and  15  theoretical  plates. 
Since  the  bottoms  of  this  tower  are  essentially  water,  live  steam  could  be  used,  but 
if  the  same  phenol  recovery  (99.99  per  cent)  were  obtained,  a  total  of  20  theoretical 
plates  would  be  required.  The  large  increase  in  the  number  of  plates  is  due  to 
dilution  by  the  large  amount  of  vapor  used.  If  the  same  bottoms  composition 
had  been  maintained  instead  of  the  same  recovery,  16  theoretical  plates  would  be 
required  and  the  recovery  would  be  98.9  per  cfent. 

The  steps  for  the  region  x  =  0.00001  to  0.001  of  this  tower  could  be  calculated 
by  Eq.  (7-77)  since  the  equilibrium  curve  is  y  »  2x. 

In  the  case  of  tower  2  using  mol  fractions  of  water  instead  of  phenol,  the  plates 
are  stepped  up  the  operating  line  from  xw  -  0.0001  to  XF  **  0.669.  In  Fig.  8-8 
a  still  and  six  theoretical  plates  would  give  a  reflux  composition  of  0.54,  and  seven 
theoretical  plates  give  0.71.  Thus  a  still  and  seven  theoretical  plates  would  give  a 
slightly  better  separation  than  desired. 

Solution  of  Part  3.  At  total  reflux,  the  plates  for  tower  1  correspond  to  the 
steps  between  the  equilibrium  curve  and  the  y  —  x  line  from  xw  »  0.00001  to  the 
reflux  composition,  XR  «  0.0168.  From  the  diagram  it  is  found  that  a  still  and 
between  11  and  12  theoretical  plates  are  needed. 

For  tower  2,  the  steps  at  total  reflux  go  from  x  «  0.0001  to  0.669.  In  this  case, 
a  still  and  five  theoretical  plates  are  required. 


202 


FRACTIONAL  DISTILLATION 


It  might  be  thought  that  the  effectiveness  of  the  fractionation  would  be  improved 
by  refluxing  to  tower  1  not  only  the  1.68  per  cent  phenol  layer  but  also  a  portion 
of  the  33.1  per  cent  layer.  This  would  result  in  an  increased  overhead  vapor  com- 
position but  would  require  additional  plates  for  the  same  vapor  rate.  Thus  for 
the  same  separation  and  vapor  rate,  more  plates  are  required,  indicating  that  only 


O.I 
0.08 
0.06 

0.04 
0.02 
001 

1 

y 

^cj^r  operating  line  •• 

--^ 

^/ 

irf 

0.008 

_^J_ 

^2|__^ 

0.006 
|^     0.004 
§ 

£      0002 
i 

"*•     0.00! 

_ 

^ 

~^ 

*r 

^x 

^ 

2 

^/ 

7 

? 

*jf 

i( 

$? 

\\^ 

Z^W 

?/• 

operaf 

//7^  ///7tf 

Z__ 

./,'„.  ..„ 

.1     0.0008 
g     0.0006 

A  (\f\f\A 

—  j7 

^~ 

__    ^ 

^  

/  ~~ 

TSr 

y 

^ 

|J/ 

/^ 

ji/1 

0      U.UUU4 

0.0002 
0.0001 

/ 

, 

JS 

/ 

^ 

"7 

/ 

•'7 

£    _ 

0.00008 

-7 

/? 

t  . 

0.00006 
0.00004 

000002 

0.00001 
O.OC 

•-•••-^ 

"™ 

2 

x^ 

y 

s/ 

> 

^ 

/ 

X 

Z_ 

00!             000004 

0000!              0.0004 

000!                 0004 

001        002 

Mo!  fraction  phenol  in  liquid 
FIG.  8-7.     Diagram  for  tower  1. 

the  1.68  per  cent  layer  should  be  refluxed.  For  compositions  between  1.68  and 
33.1  per  cent  "phenol,  decantation  is  a  more  efficient  method  of  separation  than 
distillation. 

This  two-tower  system  can  be  used  to  separate  a  number  of  binary 
systems,  such  as  isobutanol-water,  aniline-water,  or  benzene-water. 
In  the  last  system  the  solubility  of  water  in  benzene  is  so  low  that  a 
single-tower  system  is  usually  used  for  the  dehydration  of  benzene, 


SPECIAL  BINARY  MIXTURES 


203 


and  the  water  layer  is  discarded  without  treating  it  in  a  stripping 
tower. 

Some  partially  miscible  mixtures  cannot  be  separated  directly  by 
this  method.  For  example,  methyl  ethyl  ketone  and  water  are  par- 
tially miscible  and  form  an  azeotrope  at  atmospheric  pressure,  but 


1.0 

-i4 

"-X 

^ 

^  ^ 

V 

^ 

0.4 

/ 

X 

/ 

^ 

^ 

,/ 

0.1 

/ 

x 

_  ?t 

/^ 

rrri 

X 

4r* 

j-    /\  t\i 

,16 

/ 

y 

/ 

j 

CL  0.04 

o^ 

/ 

/ 

> 

c 

^ 

V 

/ 

k. 
« 

*    001 

1 

^  V 

/^ 

^ 

^ 

ra\ 

'/>3 

^  ///5 

« 

.2 
g 

"""71 

-r    — 

—  - 

*~t* 

/ 

/y 

? 

^  0.004 
"o 

/ 

/ 

^ 

y 

F- 

X^ 

0.001 

/ 

X 

5* 

x^ 

T 

0.0004 

'v' 

<^x 

f\  ftOAl 

^ 

0.0001 
0.0 

001 

00 

004 

aooi 

00 

34 

00! 

0 

04 

01 

0 

^ 

U 

Mol  fraction  water  in  liquid 
FIG.  8-8.     Diagram  for  tower  2. 

the  compositions  of  the  two  liquid  phases  do  not  bracket  the  constant- 
boiling  mixture  composition.  Cooling  and  decantation  will  not  take 
the  composition  across  the  azeotrope.  In  this  case,  if  salt  is  added  to 
the  decantation  system,  the  solubility  limits  can  be  made  to  overlap 
the  azeotrope  and  the  fractionating-decantation  system  will  make  the 
separation. 

2.  Modification  of  Relative  Volatility.  The  two  most  common  meth- 
ods of  modifying  the  relative  volatility  bf  azeotropic  mixtures  involve 
(1)  changing  the  total  pressure  and  (2)  adding  other  components  to 


204  FRACTIONAL  DISTILLATION 

the  mixture.  The  effect  of  pressure  on  the  azeotropic  composition  will 
be  considered  in  the  following  section  and  the  second  method  will  be 
analyzed  in  Chap.  10. 

The  effect  of  pressure  on  the  azeotropic  composition  is  the  result  of 
(1)  the  change  in  the  ratio  of  the  vapor  pressures  and  (2)  the  change  in 
the  activity  coefficients.  At  a  given  composition  the  change  in  the 
activity  coefficients  is  usually  small  in  comparison  to  the  effect  of  the 
Vapor-pressure  ratio.  Thus,  the  qualitative  effect  of  pressure  on  the 
azeotropic  composition  can  be  predicted  from  the  vapor-pressure  ratio. 
In  the  case  of  ethanol  and  water  mixtures  at  atmospheric  pressure, 
ethanol  is  the  more  volatile  component  for  mixtures  containing  less 
than  89.5  mol  per  cent  alcohol,  and  the  less  volatile  component  for 
more  concentrated  solutions.  The  ratio  of  the  vapor  pressure  of 
ethanol  to  water  decreases  with  increasing  temperature  and,  assuming 
that  the  activity  coefficients  do  not  change,  the  azeotropic  composition 
should  decrease  in  alcohol  content  as  the  total  pressure  increases. 
This  conclusion  is  in  agreement  with  the  experimental  data. 

A  more  quantitative  prediction  can  be  obtained  by  combining  the 
Margules  equation  for  the  activity  coefficients  with  equations  for  the 
vapor  pressures. 

At  the  azeotropic  composition, 


v  = 
7      Px 


and  Eq.  (3-34a)  becomes 


T°-26(ln  TT  -  In  P«)  =  xftV  +  c'(xz  +  0.5)]  (8-2) 

The  variation  of  the  azeotrope  composition  with  temperature  can 
be  obtained  by  subtracting  Eq.  (8-1)  from  Eq.  (8-2). 

yo.26  ln       «  v(2Xl  -  1)  -  c'(l  -  3aa  +  l.6xl)          (8-3) 


If  cr  and  6'  have  been  determined  for  one  temperature,  the  value  of 
#i  can  be  calculated  as  a  function  of  the  ratio  of  the  vapor  pressures. 
The  total  pressure  can  then  be  calculated  by  either  Eq.  (8-1)  or  (8-2). 

Equations  (8-1)  and  (8-2)  can  be  combined  with  empirical  equations 
for  the  vapor  pressure  as  a  function  of  the  temperature  to  give  a  rela- 
tion between  the  total  pressure  and  the  azeotropic  composition,  but 
the  procedure  outlined  in  the  preceding  paragraph  will  be  found  simpler 
in  general. 


SPECIAL  BINARY  MIXTURES 


205 


Equation  (8-3)  was  applied  to  the  system  ethanol-water,  using  T  as 
degree  Kelvin,  V  =  0.605,  and  c'  =  6.01.  These  values  were  obtained 
by  fitting  Eq.  (8-1)  and  (8-2)  to  the  azeotrope  data  for  atmospheric 
pressure.  The  calculated  results  are  compared  with  the  experimental 
data  (Ref .  2)  in  Fig.  8-9.  Some  of  the  difference  shown  in  this  figure 
is  due  to  the  fact  that  the  constants  used  in  the  equation  were  based  on 


£WV 

1000 
a?  800 

1    600 
£  500 

1   40° 

*•  300 
</> 

ja 

EOO 

100 

xf> 
*C* 

perit 

1/CU/C 

rjsnfi 
v  feet  I 

,/ 
*fi 

8-3 

) 

0 

/• 

A 

r   o 

9 

/ 

^ 

j 

V 

O 

y 

/ 

^ 

^* 

^> 

)       1       2       3      4       5       6       7       8       9      10      11      12 

Mol  percent  water  in  azeotrope 
FIG.  8-9.     Effect  of  pressure  on  azeotrope  composition  for  system,  ethanol-water. 

an  azeotropic  composition  at  atmospheric  pressure  of  89.4  mol  per  cent 
alcohol  (Ref.  4),  while  the  experimental  data  plotted  give  90  per  cent 
alcohol. 

Example  of  Fractionation  at  Two  Pressures  to  Separate  Azeotrope.  Lewis 
(Ref.  3)  suggested  rectification  at  two  different  pressures  to  produce  absolute 
alcohol  from  aqueous  solutions.  As  an  example  of  the  use  of  this  system,  consider 
the  production  of  99.9  mol  per  cent  alcohol  from  an  aqueous  feed  containing  30  mol 
per  cent  alcohol.  The  azeotrope  in  this  case  increases  in  alcohol  content  as  the 
pressure  is  reduced.  Thus,  by  operating  at  reduced  pressure,  an  overhead  product 
can  be  produced  which  contains  a  higher  percentage  of  alcohol  than  corresponds  to 
the  azeotrope  at  some  higher  pressure.  By  redistilling  this  overhead  product  at  a 
higher  pret  <»re,  it  can  be  separated  into  a  high-concentration  alcohol  as  bottoms 
and  an  overhead  which  can  be  recycled  to  the  low-pressure  tower.  Figure  8-10 
shows  such  an  arrangement.  The  30  mol  per  cent  feed  is  introduced  into  the  low- 
pressure  tower  which  operates  at  95  mm.  Hg  abs.  and  produces  an  overhead  con- 
taining 95  mol  per  cent  alcohol  and  a  bottoms  containing  0.0001  per  cent  alcohol. 
The  95  per  cent  overhead  product  is  pumped  into  the  higher  pressure  tower  which 
will  operate  at  atmospheric  pressure  and  produce  an  overhead  containing  92.5 


206 


FRACTIONAL  DISTILLATION 


per  cent  alcohol  and  a  bottoms  containing  99.9  per  cent  alcohol.  The  overhead  in 
this  column  will  be  returned  to  the  low-pressure  column  for  further  rectification. 
This  stream  may  be  returned  either  as  vapor  or  as  liquid;  however,  the  condensing 
temperature  of  the  atmospheric  column  is  sufficiently  high  that  it  will  serve  as  the 
heat  supply  for  the  low-pressure  column,  and  in  such  a  case  heat  economy  will  be 
obtained  by  totally  condensing  the  overhead  and  recycling  the  92.5  per  cent  stream 
as  a  liquid.  In  order  to  simplify  the  calculations,  it  is  assumed  that  all  three  feed 


30% 
alcohol 


0, 

95%  alcohol 


0.000/%  alcohol 


Fio.  8-10. 


99.3 '%  alcohol 
Production  of  absolute  alcohol. 


streams  enter  such  that  there  is  no  change  in  feed  rate  across  the  feed  plates.  The 
95  per  cent  feed  will  be  slightly  below  the  temperature  necessary  for  this  condition, 
but  it  could  be  preheated  by  countercurrent  heat  exchange  with  the  steam  con- 
densate  from  the  reboiler  of  the  atmospheric  tower.  The  92.5  per  cent  stream 
will  be  slightly  superheated  but  will  closely  approximate  the  assumption  made. 
The  usual  simplifying  assumptions  are  made,  and  it  is  assumed  that  the  vapor  rates 
in  the  two  towers  are  equal.  The  vapor  rate  will  be  calculated  for  each  tower  at  a 
reflux  ratio,  0/D,  equal  to  1.5  times  the  minimum  0/D.  The  larger  vapor  rate 
for  the  two  towers  will  be  employed. 

The  equilibrium  data  of  Lewis  and  Carey  (Ref.  4)  will  be  used  for  the  atmos- 
pheric conditions  and  the  data  of  Beebee  et  al  (Ref.  1)  for  95  mm.  Hg.  The  frac- 
tionation  will  be  most  difficult  in  the  region  containing  more  than  80  mol  per  cent 


SPECIAL  BINARY  MIXTURES 


207 


alcohol,  and  the  relative  volatilities  for  this  range  are  shown  in  Fig.  8-11.  The 
atmospheric  pressure  data  were  extrapolated  past  the  azeotrope  by  the  use  of  the 
Margules  equation.  The  data  for  95  ram.  Hg  were  drawn  to  agree  with  the 
azeotrope  composition  reported  in  the  literature  (Ref.  2). 


1.4 


j    Data 
o  Lewis  and  Carey 
KBeebee  etal.  • 


"0.7  0.8  09  1.0 

x,Mol  fraction  alcohol  in  licjuid 

FIG.  8-11.     Equilibrium  data  for  system,  ethanol-water. 

Solution.     Basis:  100  mols  of  original  feed.     (See  Fig.  8-10  for  nomenclature.) 
By  over-all  alcohol  balances, 


30  «  0.000001  Wi  +  0.999  JF2 


By  total  balance, 


For  tower  2: 

By  alcohol  balance, 

By  total  balance, 


100  = 
2  »  30, 


0.95Z)i 


i  +  Wz 
ft7!  -  70 


0.925jD2 


30 


and 

D2  *  60,        Dl  -  90 
Minimum  reflux  ratios: 

Tower  2:  F«  *  Vm  and,  from  a  study  of  the  equilibrium  data,  it  is  apparent  that 
the  pinched-in  condition  will  occur  at  the  feed  plate.    At  x  **  0.95,  the  equilibrium 


208 


FRACTIONAL  DISTILLATION 


vapor  is  0.948  and  by  Eq.  (7-10), 

=  0.925  -  0.948 
i  0.948  -  0.95 
n  _  0  +  D  -  12.5D 


-  11.5 

750  mols 


Tower  1 :  The  minimum  reflux  ratio  in  this  case  could  be  limited  by  either  feed- 
plate  condition  or  by  a  tangent  contact  between  the  equilibrium  curve  and  the 
operating  line.  All  three  possibilities  will  be  checked: 

1.  Pinch  at  0.925  feed. 


VT 


0.925,  y  in  equilibrium 
.  0.95  -  0.9277 
"  0.9277  -  0.925 

9.25(90)  -  832  mols 


0.9277 


2.  Pinch  at  0.30  feed. 


x  =  0.30,  y  in  equilibrium  ~  0.5875 
Using  the  operating  line  below  the  feed  plate, 

0.5875  -  0.000001 


0.3  -  0.000001 
-  1.96 

Om  -  Vm  +  70 
Vm  -  72.9  mols 

This  is  much  less  vapor  than  required  for  the  feed  at  x  =  0.925,  and  the  minimum 
reflux  ratio  will  not  be  determined  by  a  contact  at  #  =  0.3. 

3.  Tangency.  For  this  case,  the  832  mols  of  vapor  required  for  x  =  0.925  will 
be  used  to  determine  whether  the  operating  line  between  x  =  0.3  and  0.925  is 
below  the  equilibrium  curve.  In  this  section  the  overflow  will  be 

O  -  832  -  90  +  60  -  802 
V  =  832 

and  the  equation  of  the  operating  line  will  be 

0  _  802  _  0.9277  -  y 
V  ~  832  ~  0.925  -  x 

y  values  were  calculated  for  a  series  of  values  of  x  and  are  given  in  Table  8-1. 

TABLE  8-1 


X 

3/op    line 

^/equilibrium 

0.925 

0  9277 

0  9277 

0.92 

0.9229 

0  923 

0  91 

0  9132 

0.9139 

0  90 

0.9036 

0  9048 

0.88 

0.8843 

0  887 

0.86 

0.8651 

0.87 

0.84 

0.8458 

0.854 

0.82 

0  8265 

0  838 

SPECIAL  BINARY  MIXTURES 


209 


In  all  cases  the  equilibrium  curve  is  above  the  operating  line  except  at  X' »  0.925. 
Thus  the  minimum  reflux  ratio  for  tower  1  is  determined  by  a  pinched-in  condition 
at  x  =  0.925.  The  minimum  vapor  corresponding  to  this  condition  is  832  mols 
which  is  greater  than  the  750  mols  calculated  for  tower  2.  This  larger  value  will  be 
taken  as  the  minimum  vapor  rate  for  the  system.  Thus, 


Tower  1: 


Operating  lines  i 
Above  x  =  0.925. 


(§)- =  8-25(1-5)  =  12-38 


OR  -  1,115 
VT  -  1,205 


1,115 


90 


-  0.9254zn+i  +  0.071 


10 

•?J- 

O.I 

Hl^'TTyr 
:>!     1^ 

£ 

/ 

/ 

^ 

O.Oi 

^ 

/ 

* 

^ 

itiil/b 

rium  cu 

m 

/ 

4/ 

r 

s 
2 

8       0.001 
o 

/ 

f 

yL 

...J- 

/ 

fflr  i 

*ope 

rating 

//>7 

e 

Mol  fraction 
<=> 

^ 

^ 

2 

# 

/\ 

/ 

.00001 

_  .,_ 

...  -, 

^ 

^i 

, 

# 

/ 

Ssgs 

==• 

:  %.     ,,  ..,, 

.000001 
QOO 

v 

>  — 

il 

/ 

x; 

0001 

0.00001 

aoooi 

0.001 

0.01           at            u 

Mol  fraction  alcohol  in  liquid 
Fia.  8-12. 


210 


FRACTIONAL  DISTILLATION 


SPECIAL  BINARY  MIXTURES 


211 


From  x  -  0.3  to  x  -  0.925, 

'       M7^' 


90(0.95)  -  60(0.925) 


1,205  ~n+1  1,205 

=  0.9751o£+1  +  0.0249 

From  x  -  0.000001  to  x  -  0.3, 

-  ^275  -  70(0.000001) 
^m  ~"  1  205-a?m"1"1  1  205 

-  l'.058l£m+i  -  5.81 'x  10-8 


Tower  2: 


Operating  lines, 
Above  feed  plate, 


Below  feed  plate, 


V  -  1,205 
0*  -  1,145 


1,145 


60(0.925) 

1^05  *w+1  '   1,205 
»  0.9502xn+i  +  0.0461 


iCn-fl 


1,235  30(0.999) 

l^OS3"^1  1,205 

1.0249a;m+1  -  0.0249 


0.91 


088 


0.88 


0.89  OSO 

Mol  fraction  alcohol  in  liquid 
FIG.  8-14. 


To  Fig.  8-15 


212 


FRACTIONAL  DISTILLATION 


The  various  operating  lines  are  plotted  in  Figs.  8-12  to  8-15.  For  the  atmos- 
pheric pressure  tower  the  plates  in  the  bottom  of  the  tower  up  to  1  per  cent  water 
were  calculated  by  Eq.  (7-77). 


Equilibrium  curve 
atmospheric  pressure 


XF  for  tower  No  2 
xn  for  tower  No  / 


XD  for  fewer  No  2 
xf  for  tower  No  I 


0910   0915  0920  0925  0930  0935  0940  0945  0950  0955  0960  0965  0970  0975  0980  0985  0990  0995    10 
Mol  fraction  alcohol  in  liquid 
FIG.  8-15. 

The  low-pressure  tower  requires  a  still  and  65  theoretical  plates.  The  atmos- 
pheric pressure  tower  needs  a  still  plus  120  theoretical  plates.  In  view  of  the  high 
heat  consumption  and  the  large  number  of  plates  required,  other  methods  of 
producing  absolute  alcohol  are  more  economical.  Pressures  above  atmospheric 
would  be  advantageous  for  tower  2  and  result  in  fewer  theoretical  plates. 

Nomenclature 

6',  c'  «*  constants  in  Margules  equation 
D  «•  distillate  rate,  mols  per  unit  time 
Q  as  overflow  rate,  mols  per  unit  time 
P  »  vapor  pressure 

R  «  net  molal  withdrawal  rate  from  section  other  than  Vn  and  On+i 
8  »  steam  rate,  mols  per  unit  time 


SPECIAL  BINARY  MIXTURES  213 

T  «•  temperature 

V  «•  vapor  rate,  mols  per  unit  time 
W  «  bottoms  rate,  mols  per  unit  time 
x  »  mol  fraction  in  liquid 
#  «  mol  fraction  in  vapor 
7  «  activity  coefficient 
IT  «•  total  pressure 

References 

1.  BEEBEE,  COULTER,  LINDSAY,  and  BAKER,  Ind.  Eng.  Chem.,  34, 1501  (1942). 

2.  "International  Critical  Tables,"  Vol.  Ill,  p.  322,  McGraw-Hill  Book  Company, 
Inc.,  New  York,  1928. 

3.  LEWIS,  U.S.  Patent  1,676,700  (1928). 

4.  LEWIS  and  CAREY,  Ind.  Eng.  Chem.,  24,  882  (1932). 

5    SIMS,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1933. 


CHAPTER  9 
RECTIFICATION  OF  MULTICOMPONENT  MIXTURES 

Multicomponent  mixtures  are  those  containing  more  than  two  com- 
ponents in  significant  amounts.  In  commercial  operations,  they  are 
encountered  more  generally  than  are  binary  mixtures,  and  as  with 
binary  mixtures,  they  can  be  treated  in  batch  or  continuous  opera- 
tions, in  bubble-plate  or  packed  towers.  Since  the  continuous  opera- 
tion is  much  more  amenable  to  mathematical  analysis,  owing  to  the 
steady  conditions  of  concentration  and  operation,  it  will  be  considered 
first. 

Fundamentally,  the  estimation  of  the  number  of  theoretical  plates 
involved  for  the  continuous  separation  of  a  multicomponent  mixture 
involves  exactly  the  same  principles  as  those  given  for  binary  mixtures. 
Thus,  the  operating-line  equations  for  each  component  in  a  multicom- 
ponent mixture  are  identical  in  form  with  those  given  for  binary  mix- 
tures (see  page  119).  The  procedure  is  exactly  the  same;  i.e.,  starting 
with  the  composition  of  the  liquid  at  any  position  in  the  tower,  the 
vapor  in  equilibrium  with  this  liquid  is  calculated;  and  then  by  apply- 
ing the  appropriate  operating  line  for  the  section  of  the  tower  in  ques- 
tion to  each  component,  the  liquid  composition  on  the  plate  above  is 
determined,  and  the  operation  repeated  from  plate  to  plate  up  the 
column.  However,  actually  the  estimation  of  the  number  of  theoreti- 
cal plates  required  for  the  separation  of  a  complex  mixture  is  more 
difficult  than  for  a  binary  mixture.  When  considering  binary  mix- 
tures, fixing  the  total  pressure  and  one  component  in  either  the  liquid 
or  vapor  immediately  fixes  the  temperature  and  composition  of  the 
other  phase;  i.e.,  at  a  given  total  pressure,  a  unique  or  definite  relation 
between  y  and  x  allows  the  construction  of  the  y,x  curve.  In  the  case 
of  a  multicomponent  mixture  of  n  components,  in  addition  to  the 
pressure,  it  is  necessary  to  fix  (n  —  1)  concentrations  before  the  system 
is  completely  defined.  This  means  that  for  a  given  component  in  such 
a  mixture  the  y,x  curve  is  a  function  not  only  of  the  physical  character- 
istics of  the  other  components  but  also  of  their  relative  amounts. 
Therefore,  instead  of  a  single  y,x  curve  for  a  given  component,  there  are 
an  infinite  number  of  such  curves  depending  on  the  relative  amounts  of 

214 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        215 

the  other  components  present.  This  necessitates  a  large  amount  of 
equilibrium  data  for  each  component  in  the  presence  of  varying  pro- 
'portions  of  the  others,  and,  except  in  the  special  cases  in  which  some 
generalized  rule  (such  as  Raoult's  law)  applies,  these  are  not  usually 
available,  and  it  is  very  laborious  to  obtain  them.  One  of  the  greatest 
uses  of  multicomponent  rectification  has  been  in  the  petroleum  indus- 
try; for  a  large  number  of  the  hydrocarbon  mixtures  encountered  in 
these  rectifications,  generalized  rules  have  been  developed  which  give 
multicomponent  vapor-liquid  equilibria  with  precision  sufficient  for 
design  calculations.  Such  data  are  usually  presented  in  the  form 
y  =  Kxj  where  K  is  a  function  of  the  pressure,  temperature,  and  com- 
ponent. The  use  of  equilibrium  data  in  such  a  form  requires  a  trial- 
and-error  calculation  to  estimate  the  vapor  in  equilibrium  with  a  given 
liquid  at  a  known  pressure.  This  results  from  the  fact  that  the  tem- 
perature is  not  known,  so  a  temperature  is  assumed,  and  the  various 
equilibrium  constants  at  the  known  pressure  jju4  .assumed  temperature 
are  used  to  estimate  the  vapor  composition.  If  the_sum  of  the  mol 
fractions  of  all  the  components  in  the  vapor,  so  calculated,  add  up  to  1». 
the  assumed  temperature  was  correct.  If  the  sum  is  not  equal  to  1,  a 
new  temperature  must  be  assumed,  and  the  calculation  repeated  until 
the  sum  is  unity.  Such  a  procedure  is  much  more  laborious  than  that 
involved  in  a  binary  mixture  where  the  composition  of  the  liquid  and 
the  pressure  together  with  equilibrium  data  immediately  gives  the 
vapor  composition  without  trial  and  error. 

In  the  foregoing  discussion  of  multicomponent  systems,  it  was 
assumed  that  the  complete  composition  of  the  liquid  at  some  position 
in  the  column  was  known  as  a  starting  point  for  the  calculation.  _The 
determination  of  this  complete  composition  as  a  starting  point  is  often 
t£e  most  difficult  part  of  the  whole  multicomponent  design.  This 
difficulty  arises  from  the  fact  that  there  are  a  limited  number  of  inde- 
pendent variables  which  will  completely  define  the  distillation  process; 
therefore,  it  is  not  possible  to  select  arbitrarily  the  complete  composi- 
tion of  a  liquid  or  vapor  at  some  position  in  the  distillation  system. 
The  deggpgesjDO  system  can  be 

evaluated  (Ref.  3)  by  applyinglTTthe  law  of  conservation  of  matter, 
(2)  tEe  law  of  conservation  of  energy,  and  (3)  the  second  Jaw  of  thexmo- 
.  These  laws  together  with  the  phase  rule  can  be  applied  to 
,  the  still,  and  the  condenser  in  a  distillation  unit  and  the 
over-M  degrees  of  freedom  for  the  system  determined.  For  the  case  of 
a  reeling  column  consisting  of  a  total  condenser,  a  reboiler,  a  feed 
plate,lf  theoretical  plates  above  the  feed  plate,  and  m  theoretical 


216  FRACTIONAL  DISTILLATION 

plates  below  the  feed  plate,  the  degrees  of  freedom  for  a  system  involv- 
ing C  components  is 

C  +  2m  +  2n  +  10  (9-1) 

The  variables  used  for  these  degrees  of  freedom  are  usually  chosen  from 
the  ones  summarized  in  Table  9-1.  Theoretically,  the  choice  of 
variables  is  completely  independent,  but  in  practically  all  distillation 
calculations  certain  of  those  given  in  the  tables  are  ordinarily  fixed. 
For  example,  it  is  usual  to  define  the  composition  and  condition  of  the 
feed,  the  operating  pressure  of  each  plate,  and  the  heat  gain  or  loss  to 
or  from  each  plate  and  the  condenser.  Referring  to  the  table,  these 
four  items  add  up  to  C  +  2m  +  2n  +  6,  leaving  four  variables  that 
can  still  be  assigned.  In  most  cases,  to  facilitate  the  design  calcula- 
tions the  reflux  ratio  is  fixed,  and  in  general  it  is  desirable  to  carry  out 
the  separation  specified  with  the  minimum  number  of  theoretical 
plates;  i.e.,  the  ratio  of  n/m  is  such  that  the  total  number  of  plates 
shall  be  a  minimum  and  this  effectively  fixes  one  additional  variable. 
There  are  thus  only  two  remaining  variables  which  can  be  fixed,  and 

TABLE  9-1.     RECTIFYING  COLUMN  VARIABLES 

No.  of 
Type  of  Variable  Variables  Fixed 

Complete  composition  of  feed (C  —  1 ) 

Condition  of  feed .     .  2 

Operating  pressure  over  each  plate  and  in  still  and  condenser  . .  w  +  n  +  3 

Operating  temperature  on  each  plate  and  in  still  and  condenser .  m  +  n  -f  3 

Heat  gain  or  loss  to  or  from  each  plate  and  condenser             .    .  m  +  n  -f  2 

Heat  supplied'  to  still .         .  1 

Composition  of  product  streams  . .            .                           ,  2(C  —  1) 

Relative  quantity  of  two  product  streams .                      ...  1 

No.  plates  above  feed .          . .             ....  1 

No.  plates  below  feed ...           1 

Relative  quantity  of  liquid  returned  to  top  plate  to  overhead 

product 1 

the  choice  of  these  is  dictated  by  the  essential  nature  of  the  operation 
to  be  performed  in  the  column.  In  the  case  of  a  binary  mixture,  the 
choice  of  these  two  independent  terminal  concentrations  obviously 
gives  the  complete  compositions  of  the  distillate  and  residue  and  makes 
the  design  calculations  easy  and  straightforward.  However,  in  the 
case  of  multicomponent  mixtures,  the  problem  is  more  complex  and, 
in  general,  the  complete  composition  of  neither  the  residue  nor  the  dis- 
tillate can  be  determined  by  using  the  two  additional  factors  to  fix  two 
terminal  conditions.  In  this  case,  it  is  necessary  to  estimate  the  com- 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        217 

plete  composition  of  either  the  product  or  the  residue  and  then  proceed 
with  the  calculations  as  before  until  the  desired  degree  of  separation  is 
attained.  If,  then,  the  calculated  product  and  residue  compositions 
satisfy  a  material  balance  for  each  component,  the  estimated  composi- 
tion was  correct.  However,  if  a  material  balance  is  not  satisfied  by 
any  one  of  the  components,  it  is  necessary  to  readjust  the  composition 
and  repeat  the  calculation  until  the  material  balances  are  all  satisfied 
simultaneously.  This  estimation  is  often  simplified  because  the  degree 
of  separation  is  so  high  that  the  heavier  components  will  appear  in  the 
product  in  quantities  so  small  as  to  be  negligible.  The  same  will  be 
true  for  the  lighter  components  in  the  residue. 

In  selecting  the  two  terminal  concentrations,  it  is  desirable  to  choose 
components  that  will  give  a  significant  control  of  the  separation  desired 
and,  at  the  same  time,  be  components  that  appear  in  appreciable 
amounts  in  both  the  bottoms  and  the  distillate.  Because  these  con- 
trolling components  are  so  important  in  determining  the  design  calcu- 
lations, they  have  been  termed  the  "key  components."  In  other 
words,  they  are  the  key  to  the  design  problem. 

In  the  development  of  design  equations,  it  has  been  found  convenient 
to  pick  two  key  components:  the  light  key  component  and  the  heavy 
key  component.  The  former  is  the  more  volatile  component  whose 
concentration  it  is  desired  to  control  in  the  bottoms;  the  latter  is  the 
less"  volatile  component  whose  concentration  is  specified  in  the  distil- 
late. Thus,  in  the  stabilization  of  gasoline  it  is  often  desired  to  have 
only  a  sniall  concentration  of  propane  in  the  bottoms  in  order  that  the 
vapor  pressure  of  the  finished  product  will  meet  the  desired  specifica- 
tions and  also  to  limit  the  butane  in  the  distillate  so  as  to  retain  this 
component  in  the  gasoline.  In  such  a  case,  propane  would  be  the 
light  key  component  and  butane  the  heavy  key  component. 

TfheYerminal  concentrations  of  the  two  key  components  are  impor- 
tant because  most  of  the  practical  equations  which  have  been  developed 
for  the  minimum  number  of  theoretical  plates  at  total  reflux,  the 
optimum  feed-plate  location,  and  the  minimum  reflux  ratio  have 
involved  these  concentrations.  However,  certain  difficulties  are 
involved:  (1)  the  design  specifications  may  be  such  that  the  key  com- 
ponents are  not^bvious  and  (2)  these  design  equations  often  require 
the  concentrations  of  both  key  components  in  the  distillate  and  bot- 
toms as  well  as  the  concentration  of  some  of  the  other  components. 
But  as  demonstrated  in  the  foregoing  analysis,  only  two  of  these 
terminal  concentrations  are  independent  and  can  be  arbitrarily  fixed 
as  design  conditions. 


218  FRACTIONAL  DISTILLATION 

The  difficulties  of  choosing  the  key  components  and  estimating  the 
complete  distillate  and  bottoms  compositions  are  often  the  most  diffi- 
cult parts  of  a  multicomponent  design  calculation.  The  problem  can 
generally  be  simplified  if  the  design  conditions  are  chosen  with  this 
problem  in  mind.  Thus,  in  cases  where  the  separation  between 
adjacent  components  is  essentially  complete,  the  two  independent 
variables  can  be  chosen  as  the  concentration  of  the  more  volatile  of 
these  two  in  the  bottoms  and  as  the  concentration  of  the  less  volatile 
component  in  the  distillate.  These  adjacent  components  then  become 
the  key  components,  and  the  composition  of  the  distillate  and  bottoms 
can  be  determined  completely  enough  for  design  calculations  by  simple 
material  balances.  Components  more  volatile  than  the  light  key  com- 
ponent will  be  almost  negligible  in  the  bottom,  and  components 
heavier  than  the  heavy  key  component  will  be  negligible  in  the  dis- 
tillate. For  rectifications  in  which  there  is  an  appreciable  difference 
in  volatility  between  adjacent  components  and  in  which  a  fairly  high 
degree  of  separation  is  being  carried  out,  the  design  condition  can  gen- 
erally be  specified  in  this  manner  and  thereby  simplify  the  problem. 
If  the  degree  of  separation  is  low  and/or  there  are  several  components 
of  nearly  the  same  volatility  in  the  range  in  which  the  separation  is 
being  made,  the  selection  of  the  two  key  terminal  concentrations  will 
generally  not  give  enough  information  to  allow  the  complete  terminal 
compositions  to  be  calculated  by  simple  material  balances.  In  such 
a  case,  it  is  necessary  to  estimate  the  terminal  concentration  of  the 
other  distributed  components  and  then  check  this  estimation  by  pro- 
ceeding with  the  usual  stepwise  plate-to-plate  calculations.  If  such 
plate-to-plate  calculations  give  a  consistent  over-all  result,  the  esti- 
mated values  are  satisfactory;  if  the  results  are  inconsistent,  new  values 
must  be  estimated  and  the  calculation  repeated.  It  should  be  empha- 
sized that  even  in  this  latter  case,  although  a  large  number  of  the 
terminal  concentrations  may  not  be  known,  only  two  of  these  are 
independent  after  the  other  variables  selected  have  been  fixed.  ,  All 
the  other  terminal  concentrations  are  fixed  when  these  two  independent 
ones  are  chosen  and  therefore  cannot  be  given  values  arbitrarily.  The 
necessity  of  having  the  concentrations  of  these  other  components,  that 
are  fixed  but  not  known,  offers  the  main  difficulty  in  setting  up  a  multi- 
component  distillation  example. 

The  above  procedure  will  give  the  optimum,  design  for  the,  operating 
conditions  chosen.  However,  occasionally  it  is  desirable  not  only  to 
obtain  the  desired  separation  between  the  two  components  but  at  the 
same  time  to  control  the  amount  of  one  of  the  other  components  in  one 
of  the  products.  Frequently,  it  is  possible  to  accomplish  this  result 


RECTIFICATION  OF  MVLTICOMPONENT  MIXTURES        219 

by  shifting  the  feed-plate  location.  Thus  the  tower  is  not  designed  for 
the  minimum  number  of  plates  to  separate  two  key  components  but 
will  employ  a  larger  number  of  plates  to  accomplish  a  desired  result. 

Lewis  and  Matheson  Method.  Several  methods  have  been  pro- 
posed for  the  design  of  multicomponent  mixtures,  but  fundamentally 
they  are  based  on  Sorel's  method.  One  of  the  best  is  that  due  to 
Lewis  and  Matheson  (Ref.  5).  This  is  the  application  of  Sorel's 
method  together  with  the  usual  simplifying  assumptions  to  multicom- 
ponent mixtures.  The  same  operating  lines  as  used  on  page  119  for 
binary  mixtures  are  employed  to  determine  the  relation  between  the 
vapor  composition  and  the  composition  of  liquid  on  the  plate  above, 
this  calculation  together  with  vapor-liquid  equilibrium  data  being 
sufficient  for  the  determination  of  the  number  of  theoretical  plates  for 
given  conditions.  The  use  of  this  method  will  be  illustrated  by  the 
fractionation  of  a  mixture  of  benzene,  toluene,  and  xylene  under  condi- 
tions where  the  separation  will  be  sufficiently  good  so  that  the  determi- 
nation of  the  terminal  conditions  will  not  be  difficult. 

Benzene-Toluene-Xylene  Example.  Consider  the  rectification  of  a 
mixture  containing  60  mol  per  cent  of  benzene,  30  mol  per  cent  of 
toluene,  and  10  mol  per  cent  of  xylene  into  a  distillate  or  product  con- 
taining not  over  0.5  mol  per  cent  of  toluen^ntfid  a  bottoms  or  residue 
containing  0.5  mol  per  cent  of  benzene  V^.  reflux  ratio  0/D  equal  to 
2  will  be  used,  and  the  feed  will  enter  preheated  so  that  the  change  in 
mols  of  overflow  across  the  feed  plate  will  be  equal  to  the  mols  of  feed. 
The  usual  simplifying  assumptions  will  be  made,  and  Raoult's  law  will 
be  used.  The  distillation  will  be  carried  out  at  1  atm.  abs.  pressure. 

Since  the  concentration  of  toluene  in  the  distillate  D  is  low,  the 
xylene  will  be  practically  zero  and  therefore  will  be  essentially  all  in  the 
residue  W.  Taking  as  a  basis  100  mols  of  feed,  a  benzene  material 
balance,  input  equals  output,  gives  the  following: 

60  «  DxDB  +  0.005TF  =  (100  -  W)xDB  +  0.005TF 

-  WQxDB  +  (0.005  - 
XDB  =  0.995 
60  «  99.5  -  TF(0.99) 


D  -  60.1 

where  D  =  mols  of  product 
W  =  mols  of  residue 

mol  fraction  of  benzene  in  liquid  distillate 


220 


FRACTIONAL  DISTILLATION 


The  terminal  conditions  are  then 


Distillate 

Residue 

Mols 

Mol 
per  cent 

Mols 

Mol 
per  cent 

Benzene  

59  8 
0  30 
0 

99  5 
0  5 

0  20 
29  7 
10.0 

0  5 

74  4 
25  1 

Toluene 
Xylene 
Total  

60  1 

100.0 

39.9 

100  0 

Since  0/D  =  2,  in  the  top  part  of  the  tower 
On  «  2  X  60.1  -  120.2, 

and   Vn  =  On  +  D  =  180.3,   and  Om  =  On  +  F  =  220.2,  giving   Vm 
=  180.3.       ' :i'    l   '    *     ' 

For  the  part  of  the  column  below  the  feed  plate,  the  operating  lines 
are 
For  benzene: 


0\  /W 

ymB  =     ).  X(m+  1)B  ~~  VT 

=  1.221a?(m+i)B  -  0.0011 
For  toluene  : 


/220.2\ 

=  \i8o3y 


__ 

180.3 


(0.005) 


For  xylene  : 


ymx  = 


-  0.164 


-  0.0555 


Beginning  with  the  composition  of  the  liquid  in  the  still,  a  tempera- 
ture is  assumed,  and  the  partial  pressure  of  each  component  is  calcu- 
lated using  Raoult's  law.  If  the  sum  of  the  partial  pressure  is  760 
mm.  Hg  (the  total  pressure),  the  assumed  temperature  was  correct. 
The  vapor  pressure  of  these  components  is  given  in  Fig.  9-1.  Assume 
T  =  115°C. 


C6 
C7 
C8 

xw 

P 

xwP 

yw  =  xwP/  2xwP 

0  005 
0.744 
0.251 

1990 
850 
390 

10 
632 
98 

0  0135 
0  854 
0  1325 

740 

1  0000 

Since  the  total  is  740  instead  of  760,  the  assumed  temperature  was 
too  low,  but  a  more  nearly  correct  temperature  is  easily  found  by  deter- 
mining the  temperature  from  Fig.  9-1  at  which  the  vapor  pressure  of 
toluene  is  (76!^4o)(850)  *  873,  giving  T  «  116.0°C.  The  calculation 
is  then  repeated  for  116°C. 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        221 


Xw 

P 

xwP 

yw  «  xwPHW  4 

Ce 

0.005 

2000 

10 

0  0131 

C7 

0  744 

873 

650 

0  855 

C8 

0  261 

400 

100  4 

0  132 

760  4 

1  0000 

1800 


1400 
1300 
1200 
1100 


E  900 

fsoo 


With  this  assumed  temperature,  the  sum  of  xwP  is  seen  to  be  very 
close  to  760,  and  the  temperature  is  satisfactory.  Actually,  such  a 
recalculation  is  not  necessary,  since 
the  values  desired  are  ywy  and 
these  values  may  be  obtained  from 
yw  =  XwP/^xwP  where  2xwP  is 
the  sum  of  the  xwP  values.  Thus, 
in  the  fourth  column  of  the  table 
for  the  first  assumed  temperature 
are  given  values  of  yw  =  x^P/740. 
These  are  seen  to  correspond  closely 
to  the  values  in  the  corrected  table 
and  agree  well  within  the  accuracy 
of  such  factors  as  the  vapor  pres- 
sures, the  applicability  of  Raoult's 
law,  etc.  In  general,  such  a  simpli- 
fied procedure  is  satisfactory  when 
the  sum  of  XwP  is  within  10  per  cent 
of  the  desired  value;  however,  at 
times,  such  a  simplification  is  not 
justified,  and  a  preliminary  check 
on  the  system  in  question  should  be 
made  to  determine  the  satisfactory 
limit  of  the  sum  of  xwP. 

The  value  of  x:  is  obtained  from  yw  by  the  use  of  the  appropriate 
operating-lme  equation  applied  to  each  component,  yi  is  then  calcu- 
lated using  Raoult's  and  Dalton's  laws  at  an  assumed  temperature  of 
115°C.,  and  £2  is  obtained  from  the  values  of  y\  by  the  use  of  the 
operating-line  equation.  The  operation  is  repeated,  making  adjust- 
ments of  the  assumed  temperatures  such  that  2xP  stays  between  700 
and  820.  In  making  these  adjustments  of  temperature,  it  is  desirable 
to  continue  using  one  temperature  until  the  value  o^SzP  is  about  as 
much  greater  than  760  as  it  was  less  than  760  on  the  first  plate  on  which 
the  temperature  was  used.  Thus,  in  the  following  table,  the  values  of 


80 


90         100        no 

Temperature, deg  C. 
FIG.  9-1. 


Com- 
ponent 

T,  °C. 

as- 
sumed 

mm 

xw 

XWP 

yw  -  XwP/2xwP 

Xl 

Ce 
CT 
C8 

115 

1990 
850 
390 

0.005 
0.744 
0  251 

10 
632 

98 

0.0135 
0.854 
0.1325 

ZxP  -  740 

2y  *  1.0000 

Xi 

Ce 
C7 
C8 

116 

2000 
873 
400 

0  005 
0.744 
0.251 

10 
650 
100.4 

0.0131 
0.855 
0  132 

0  0116 
0.835 
0  153 

VxP  =760.4 

2y  -  1.0000 

Xi 

*lP 

2/i 

rc2 

Ce 
C7 
Cs 

115 

1990 
850 
390 

0.0116 
0.835 
0.153 

23  1 
709 
59.7 

0.0292 
0.895 
0.0755 

0  0248 
0  868 
0  1065 

2zP  «  791  8 

2y  -  1 

xz 

x*P 

2/2 

X3 

Ce 

C7 
C8 

110 

1740 
740 
330 

0.0248 
0.868 
0.1065 

43 
642 
35 

0  0597 
0  892 
0.0486 

0  0498 
0  865 
0.085 

2zP  -  720 

38 

x*P 

2/3 

X* 

Ce 

C7 
C8 

110 

1740 
740 
330 

0  0498 
0  865 
0.085 

86.7 
640 

28 

0  115 
0  848 
0.037 

0  095 
0  830 
0.075 

S#P  -  754  7 

#4 

x*P 

2/4 

x& 

C« 
CT 
C8 

110 

1740 
740 
330 

0  095 
0  830 
0.075 

165 
614 
25 

0.205 
0  763 
0.031 

0.169 
0.759 
0.071 

SarP  -  804 

3« 

xj> 

y«  « 

Xt 

Ce 
CT 
Cs 

105 

1520 
645 
280 

0.169 
0.759 
0.071 

257 
489 
20 

a.  336 
0.638 
0.026 

0.276 
0.657 
0.067 

SxP  *  766 

222 


RECTIFICATION  OF  MULT  I  COMPONENT  MIXTURES  ,      223 


110°C.  were  continued  from  xf  =  720  to  xJP  =  804,  giving  approxi- 
mately an  equal  displacement  on  both  sides  of  760. 


Com- 
ponent 

J.    J         O. 

as- 
sumed 

K  -  P/760 

x* 

xK 

y,  -  xK/ZxK 

*, 

C6 

100 

1  745 

0  276 

0.482 

0.49 

0.402 

C7 

— 

0  735 

0  657 

0  482 

0  49 

0  535 

C8 

__ 

0  316 

0.067 

0.021 

0  021 

0.063 

ZxK  -  0.985 

x. 

x,K 

.        2/7 

x, 

C6 

95 

1  52 

0.402 

0.612 

0.635 

0.521 

C7 

— 

0  628 

0  535 

0  336 

0  348 

0  420 

C8 

— 

0.263 

0.063 

0  016 

0.017 

0.059 

VxK  ~  0  964 

#8 

x* 

If. 

**, 

C6 

95 

1  52 

0  521 

0  793 

0  738 

0  605 

C7 

— 

0.628 

0  420 

0.264 

0  246 

0  336 

C8 

— 

0.263 

0  059 

0  016 

0  015 

0.058 

VxK  =  1  073 

The  table  on  page  222  carries  these  calculations  up  to  the  sixth  plate, 
As  pointed  out  on  page  215,  the  vapor-liquid  data  are  more  often 
given  as  y  =  Kx  rather  than  as  Raoult's  law.  With  equilibrium  data 
in  such  form,  the  method  of  calculation  is  similar.  Given  the  values 
of  x  in  the  liquid  on  any  plate,  the  temperature  is  assumed,  and  a  value 
of  K  for  each  component  is  obtaijied  from  equilibrium  data  at  the 
assumed  temperature  and  the  operating  pressure.  The  value  of  y  in 
equilibrium  with  this  liquid  is  given  by  Kx.  If  the  sum  of  the  values 
of  y  is  equal  to  1,  the  assumed  temperature  was  correct,  and  the  x 
values  on  the  plate  above  may  be  obtained  from  the  y  values  just  cal- 
culated by  using  the  operating-line  equation.  If  the  sum  of  the  values 
of  Kx  does  not  equal  1,  the  temperature  should  be  readjusted;  but  as  in 
the  previous  case,  this  adjustment  is  usually  unnecessary  if  the  sum  of 
Kx  is  within  10  per  cent  of  1,  in  which  case  the  values  of  y  are  calculated 
by  ya  =  KaXa/I,Kx. 

The  benzene-toluene-xylene  calculations  will  be  continued,  using  the 
K  method.  For  this  particular  mixture,  where  Raoult's  and  Dalton's 
laws  are  assumed  to  apply,  the  equilibrium  constant  is  equal  to  the 
vapor  pressure  divided  by  the  total  pressure;  e.g.,  yair  »  Xa 


or 


224 


FRACTIONAL  DISTILLATION 


The  ratio  of  xC6  to  zC7  on  plate  9  is  approximately  that  in  the  feed, 
so  this  plate  will  be  used  as  the  feed  plate.  The  proper  feed-plate  loca- 
tion for  this  column  will  be  considered  in  a  later  section.  Above  the 
feed  plate,  the  procedure  is  the  same,  except  that  the  equation  for  the 
upper  portion  of  .the  tower  is  utilized. 

The  operating-line  equations  above  the  feed  are 
For  Oe;  " '"" 


=  (^  ) 

\V  /n 


0.332 


For  C7, 


For 


ynT  = 


0.0017 


Proceeding  as  before: 


Com- 
ponent 

T,  °C. 

as- 
sumed 

X 

X9 

xK 

y,  =  xKfZxK 

-BIO 

C6 
C7 
C8 

90 

1  33 
0  533 
0.221 

0  605 
0.336 
0  058 

0  805 
0.179 
0  013 

0  807 
0.180 
0  013 

0  712 
0  267 
0  020 

ZxK  =  0  997 

2lO 

XuK 

2/io 

Xn 

C6 
C7 
Cs 

85 

1.15 
0  452 
0  184 

0  712 
0.'207 
0.020 
*  * 

0.819 
0  121 
0  004 

0.867 
0  128 
0  004 

0  802 
0.189 
0.006 

VxK  -  0  944 

Xn 

XnK 

2/ii 

#12 

Ce 
C7 
C8 

85 

1  15 
0  452 
0.184 

0  802 
0  189 
0.006 

0  923 
0  085 
0  0012 

0.914 
0  084 
0.0012 

0  873 
0  123 
0.0018 

2xK  -  1.009 

#12 

XnK 

2/12 

#13 

C6 

C7 
C8 

85 

1.15 
0  452 
0.184 

0.873 
0.123 
0.0018 

1.005 
0.055 
0.0004 

0.947 
0  053 
0.0004 

0.922 
0.0765 
0.0006 

2xK  =  1.061 

RECTIFICATION  OF  MULTICOMPONENT  MIXTURES         225 


Com- 
ponent 

as- 
sumed 

K 

Xg 

xK 

_i 

£-"•" 

£10 

X* 

xnK 

2/13 

*u 

c« 

C7 
C8 

80 

0.995 
0.379 
0.153 

0  922 
0  0765 
0.0006 

0  917 
0  029 
0  0001 

0  968 
0.032 
0.0001 

0.953 
0.045 
0.00015 

SarJfC  ==  0  946 

1 

*14 

XuK 

014 

#16 

C8 
C7 
C8 

80 

0  995 
0  379 
0  153 

0  953 
0  045 
0  00015 

0  948 
0  017 
0.00002 

0  982 
0  018 
0  00002 

0  974 
0.024 
0  00003 

#15 

*i* 

2/15 

*, 

C6 
C7 

C8 

80 

0  995 
0  379 
0  153 

0  974 
0  024 
0  00003 

0  969 
0  0091 
0  000005 

0  99 
0  0093 
0  000005 

0.988 
0  0114 
0.000007 

*!, 

*16* 

2/16 

C6 
C7 
C8 

80 

0  995 
0.379 
0.153 

0  988 
0.0114 
7  X  10~6 

0  983 
0  0043 

io-6 

0  9956 
0  0044 
10-6 

The  vapor  leaving  the  sixteenth  plate,  on  being  liquefied  in  the  total 
condenser,  will  give  a  product  containing  slightly  more  than  99.5  per 
cent  benzene.  Thus,  approximately  16  theoretical  plates  together 
with  a  total  condenser  and  still  or  reboiler  are  required  to  effect  the 
desired  separation  under  the  operating  conditions  chosen. 

In  general,  it  is  instructive  to  plot  the  compositions  vs.  the  plates. 
This  type  of  figure  is  shown  in  Figs.  9-2  and  9-3  for  'the  example  just 
solved.  The  benzene  is  seen  to  rise  on  a  smooth  curve,  and  the  con- 
centration of  toluene  in  the  liquid  passes  through  a  maximum  two 
plates  above  the  still  and  then  falls  off  in  a  smooth  curve  with  the 
exception  of  a  slight  break  at  the  feed  plate;  the  xylene  drops  rapidly 
above  the  still  and  then  flattens  out  until  the  feed  plate  is  reached  and 
then  drops  rapidly  to  a  negligible  value.  The  maximum  in  the  toluene 
curve  is  a  result  of  the  fact  that,  in  the  still,  toluene  and  xylene  are  the 
main  components;  and  since  toluene  is  the  more  volatile  of  the  two,  it 


226 


FRACTIONAL  DISTILLATION 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        227 


228  FRACTIONAL  DISTILLATION 

tends  to  increase,  and  the  xylene  tends  to  decrease.  This  increase  in 
toluene  concentration  jefmtinues  until  the  benzene  concentration 
becomes  appreciable V^nd  since  this  latter  component  is  very  volatile, 
it  increases  rapidly  and  forces  the  toluene  to  decrease.  This  increase 
of  benzene  relative  to  the  toluene  continues  up  to  the  condenser.  The 
xylene  decreases  up  the  column  from  the  still  because  of  its  low  vola- 
tility but  cannot  decrease  below  a  certain  value,  since  the  10  mols  of 
xylene  in  the  feed  must  flow  down  the  column,  and  this  sets  a  minimum 
limit  on  the  concentration  of 

10        10 

OTm  -  220  =  0'0455 

actually  the  value  will  be  slightly  higher,  since  the  small  amount  of 
xylene  that  passes  upward  in  the  vapor  must  again  pass  down  the 
column.  This  is  due  to  the  fact  that  essentially  no  xylene  leaves  the 
top  of  the  column.  Above  the  feed  plate,  the  amount  of  xylene  passing 
with  the  vapor  to  a  plate  must  be  equal  to  the  xylene  in  the  overflow 
from  the  plate,  i.e.,  Vyn  =  Oxn+i,  since  DxD  is  essentially  zero.  How- 
ever, yn  =  Kxn,  giving  xn+i  =  (VK/On)xn\  for  a  heavy  component 
such  as  xylene  which  does  not  leave  the  top  of  the  column  in  appreciable 
amount,  the  composition  of  the  liquid  on  one  plate  is  related  to  that  on 
the  plate  below  VK/0.  In  general,  K  is  very  small  for  such  compo- 
nents and  the  concentration  decreases  rapidly  as  shown  by  the  straight 
line  in  Fig.  9-2. 

These  concentration-gradient  curves  are  typical  of  those  generally 
obtained.  The  two  main  components  between  which  the  rectification 
is  taking  place  tend  to  increase  and  decrease  up  the  column,  much  as  in 
a  binary  mixture.  They  are  often  called  the  key  components.  The 
concentrations  of  the  components  heavier  than  the  heavier  key  com- 
ponent decrease  rapidly  as  one  proceeds  from  the  still  up  the  column, 
but  they  tend  to  become  constant  because  of  the  necessity  of  their 
flowing  down  the  overflows  in  order  that  they  may  be  removed  at  the 
still.  These  components  then  decrease  rapidly  above  the  feed  plate, 
usually  dropping  to  negligible  values  a  few  plates  above  this  plate. 
The  concentrations  of  components  lighter  than  the  light  key  com- 
ponent give  the  same  type  of  curves  from  the  condenser  down  the 
column  as  the  heavier  components  do  from  the  still  upward.  Thus,  the 
concentrations  of  these  light  components  decrease  rapidly  for  a  few 
plates  down  from  the  condenser  but  then  flatten  out,  since  by  material 
balance  essentially  all  of  the  mols  of  these  components  that  are  in  the 
feed  must  flow  up  through  the  upper  part  of  the  column  to  be  removed 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES 


229 


at  the  condenser,  and  this  factor  sets  a  lower  limit  on  their  concentra- 
tion in  this  section.  Below  the  feed  plate,  these  light  components 
decrease  rapidly  and  generally  become  negligible  a  few  plates  below  the 
feed  plate. 

Lewis  and  Cope  Method.    Lewis  and  Cope  (Ref .  4)  applied  the  same 
method  graphically,  by  constructing  a  separate  y,x  plot  for  each  com- 


0   I         0.02  0.04          0.06  0.08 

'  Mol  Fraction  in  Liquid 

xw 

FIG.  9-4. 


0.10 


ponent.  On  these  plots,  the  y  =  x  line  and  the  operating  lines  are 
drawn  the  same  as  for  a  binary  mixture.  The  three  plots  for  the 
previous  examples  are  given  in  Figs.  9-4  to  9-6.  Only  the  lower  por- 
tions of  the  benzene  and  xylene  curves  are  given  in  order  to  increase  the 
graphical  accuracy.  It  is  interesting  to  note  that  for  xylene  and 
toluene  the  intersection  of  the  operating  line,  which  occurs  at  xj?,  just 
as  in  the  case  of  a  binary  mixture,  falls  below  the  y  =  x  diagonal. 
This  is  due  to  the  fact  that  xw  is  greater  than  XD,  and  the  components 
are  both  of  lower  volatility  than  the  benzene.  If  unique  equilibrium 


230 


FRACTIONAL  DISTILLATION 


0.2  0.4  0.6  0.8 

Mol  Fraction  in  Liquid 

FIG.  9-5. 


O.I  0.2  Q.3 

Mol  Fraction  in  Liquid 
FIG.  9-6. 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        231 

curves  could  be  drawn  on  the  diagrams,  the  problem  would  become 
similar  to  the  stepwise  procedure  for  a  binary  mixture.  However,  IB 
general,  such  curves  are  not  known.  The  Lewis  and  Cope  method 
was  to  draw  a  series  of  equilibrium  curves  of  the  type  y  ~  Kx  which 
at  constant  temperature  are  in  general  straight  lines  through  the  origin 
of  slope  K.  Thus,  in  the  present  example,  K  =  P/760,  where  P  of 
any  one  component  is  a  function  of  the  temperature  only.  Such 
equilibrium  curves  have  been  drawn  in  for  the  temperatures  of  105, 
110,  and  115°C.  Starting  at  xw  on  each  plot,  vertical  lines  are  drawn 
through  this  point  cutting  the  equilibrium  curves.  By  trial  and  error, 
temperatures  are  tried  until  the  sum  of  the  y  values  at  the  intersection 
of  the  vertical  line  through  Xw  and  the  equilibrium  curve  for  the 
assumed  temperature  adds  up  to  unity.  Thus,  if  115°C.  is  tried,  the 
sum  of  the  y  values  at  the  intersection  of  the  115°C.  curve  with  the  xw 
lines  is  0.013  +  0.837  +  0.13  =  0.98,  indicating  that  115°C.  is  too  low. 
By  interpolation  at  116°C.,  the  sum  becomes  0.013  +  0.855  4-  0.132 
=  1.00,  indicating  that  this  is  the  correct  temperature,  and  the  y 
values  give  the  composition  yw  of  the  vapor  in  equilibrium  with  Xw. 
Horizontal  lines  are  then  drawn  through  the  yw  values  to  the  operating 
line,  the  abscissa  of  the  intersection  with  the  operating  line  being  x\. 
Vertical  lines  are  drawn  through  the  #i's;  and  by  using  the  same  proce- 
dure as  for  xw  a  temperature  of  112.5°C.  is  found  to  give  I>y  equal  to 
unity,  and  the  step  is  then  completed  to  the  operating  line.  In  a  like 
manner,  steps  are  taken  up  the  column.  The  same  operating  line  is 
used  until  the  feed  plate  is  reached,  and  theh  the  change  is  made  to  the 
operating  line  for  the  upper  portion  of  the  column  simultaneously  for 
all  three  components. 

A  comparison  of  the  values  of  these  figures  with  those  obtained  in 
the  previous  algebraic  calculation  shows  the  close  agreement.  Actu- 
ally, they  have  to  give  the  same  result,  since  they  both  are  solutions  ol 
the  same  set  of  equations,  one  being  algebraic  and  the  other  graphical, 
Both  methodsjhiavejbheir  advantages;  in  the  algebraic  method,  as  a 
rule,TugEer  accuracy  can  be  obtained  "than  in  the  graphical  method 
this  {^especially  true  in  the  low-concentration  region  where  the  graphi- 
cal diagram  must  t>e  greatly  expanded  or  replotted  on.  logarithmic 
paper,  such  as  was  utilized  in  the  binary  mixtures.  The  advantage  o1 
tte  graphical  method  is  that  it  gives  a  visual  picture  of  the  concentra- 
tion gradients  and  operation  of  the  tower.  The  amount  of  labor  anc 
time  consumed  i£  approximately  the  same  for  the  two  methods. 

Numerous  analytical  methods  based  on  the  foregoing  methods  hav< 
been  proposed  to  simplify  the  trial  and  error  required  in  the  Lewis  anc 


232  FRACTIONAL  DISTILLATION 

Matheson  method.  Some  of  these  methods  will  be  considered  in  a 
later  section,  but  generally  the  stepwise  method  outlined  above  is  more 
satisfactory.  By  using  y  =  Kx/^Kx  instead  of  making  the  Kx's  add 
to  exactly  unity,  the  trial-and-error  work  of  the  Lewis  and  Matheson 
method  is  practically  eliminated.  In  the  example  just  solved,  when 
became  larger  than  1,  the  temperature  was  dropped,  making 
less  than  1;  and  this  temperature  was  used  until  2Kx  again 
became  greater  than  unity,  and  then  the  temperature  was  again 
dropped.  Thus,  no  actual  trial  and  error  was  required,  but  merely 
successive  drops  of  temperature  of  5  to  10°.  Such  calculations  require 
only  a  few  hours  more  than  the  simplest  of  the  approximate  methods 
and  only  two  tor  three  such  stepwise  calculations  at  different  reflux 
ratios  together  with  the  minimum  number  of  plates  at  total  reflux  and 
the  minimum  reflux  ratio  are  required  to  allow  the  construction  of  a 
curve  of  theoretical  plates  required  vs.  the  reflux  ratio.  In  general, 
the  added  confidence  that  may  be  placed  in  the  stepwise  calculations 
relative  to  the  approximate  methods  more  than  justifies  the  extra  work 
involved. 

In  using  the  stepwise  method  with  the  simplification  that 
y  =  Kx/%Kx,  the  problem  arises  as  to  how  much  IZKx  can  differ  from 
unity  and  still  not  appreciably  affect  the  values  of  y.  The  justification 
of  this  simplification  is  that  for  moderate  changes  in  temperature  the 
percentage  change  in  the  values  of  K  for  substances  that  do  not  differ 
too  widely  is  approximately  the  same.  A  little  consideration  will  show 
that  if  all  the  K  values  change  the  same  percentage  with  temperature, 
then  the  values  of  y  calculated  by  such  a  method  will  be  independent  of 
the  temperature  chosen.  This  relative  variation  in  the  K  values  is 
best  expressed  in  the  relative  volatility.  Thus,  if  yA  =  KAxA  and 
yB  =  KBxB,  then  yt/yu  =  (KA/KB)(XA/XB),  and  (KA/KB)  is  the  rela- 
tive volatility  of  A  to  B,  OMB  (see  page  30).  If  the  percentage  change 
in  both  KA  and  KB  is  the  same  with  temperature,  aAB  will  be  a  constant 
over  this  region,  and  a  plot  of  &AB  vs.  temperature  will  give  immediately 
the  region  over  which  IZKx  can  vary  without  appreciably  altering  the 
y  value.  Actually,  the  a's  can  be  introduced  into  the  equations,  and 
the  K  'a  eliminated.  Thus, 

yA  +  VB  +  yc  +  yo  +  •  •  •  =  l 


using  the  relative  volatility 

<XAB%A    ,    «    ,    QtCBXc    ,  XD 

.         -p    1    -f~  -    ~  -f-   Otj)B 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        233 
which  can  be  rearranged  to  give 


XB 


XB 


since 


+  XB  +  otcsXc  +  otDBXp  + 
XB 


Likewise, 


A  similar  analysis  starting  with 

XA  +  XB  +  Xc 
leads  to 


90  100  110  120 

Temperature,  deg .  C . 
FIG.  9-7. 


(9-2) 


(9-3) 


where  all  the  relative  volatilities  are  with  respect  to  the  B  component; 
and  in  the  case  of  the  y*  equation,  a  relative  volatility  does  not  appear 
with  xB,  since  aBs  is  L  Given  the  liquid  composition  on  any  plate,  the 


234 


FRACTIONAL  DISTILLATION 


values  of  x  are  multiplied  by  the  a  corresponding  to  the  component  in 
question,  and  the  values  of  ax  are  totaled  to  give  Sax,  then  the  value  of 
y  for  any  component  is  calculated  by  dividing  ax  for  the  component  by 
Sox.  In  general,  it  is  desirable  to  take  the  volatilities  relative  to  one 
of  the  key  components;  this  will  cause  a  to  be  greater  than  1  for  the 
components  that  are  lighter  and  less  than  1  for  the  heavier  components. 
This  method  will  be  most  clearly  brought  out  by  its  application  to 
actual  problems.  First  it  will  be  applied  to  the  benzene,  toluene,  and 
xylene  problem  previously  solved.  Figure  9-7  shows  the  volatilities 
relative  to  toluene  plotted  as  a  function  of  temperature  and  also  shows 
the  K  for  toluene  as  a  function  of  the  temperature.  It  will  be  noticed 
that  the  variation  in  the  relative  volatilities  with  temperature  is  very 
small  and  that  for  xylene  in  the  lower  part  of  the  column  a  constant 
value  of  a  equal  to  0.43  is  well  within  the  design  accuracy.  The  ben- 
zene volatility  relative  to  toluene  varies  more,  but  even  here  the  varia- 
tion is  small.  In  the  previous  example,  starting  at  the  still: 


xw 

ano 

aXw 

yw  -  <*3/0.869 

Xl 

an  0^1 

2/i  «  a£/0.932 

Cfl 

0.005 

2.36 

0.0118 

0  0136 

0  012 

0  0283 

0  030 

C7 

0.744 

1  0 

0  744 

0.856 

0.835 

0  835 

0  896 

Cg 

0.251 

0.45 

0.113 

0.130 

0.152 

0  0684 

0  074 

0.8688 

0.9317 

KT  -  0.856/0.744 

Tw  ~  H6°C. 


1.15         KT  -  0.896/0.835  =  1.07 
Ti  -  113.4°C. 


Xz 

«110#2 

2/2  »  <*c/0.976 

0-3 

<*110#3 

2/3  -  ax  /1.  02 

Ce 

CT 
Cs 

0.0254 
0,868 
0.106 

0.06 
0.868 
0.0477 

0.061 
0.890 
0.049 

f.*.* 

0.0508 
0  864 
0.086 

0.13 
0.864 
0  039 

0.117 
0.845 
0.038 

0.9757 

1.023 

KT  -  1/0.9757  -  1.025 


KT  -  0.978 
r3  -  110,2°C. 


*4 

«iio3U 

y<  «  ax/1.  086 

X6 

auo£« 

2/6  «•  aaJ/1.192 

C6 
C7 
C* 

0.096 
0.826 
0.076 

0.226 
0.826 
0.034 

0.208 
0.760 
0.031 

0.171 
0.757 
0.071 

0.403 
0.757 
0.032 

0.338 
0.636 
0.027 

1.086 

1.192 

K  -  0.922 
T4  «  108°C. 


K  *0.84 
f  •  -  104.8'C. 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        235 

A  comparison  of  the  values  calculated  above  with  those  previously 
obtained  shows  a  very  close  agreement,  as  must  be  the  case,  since  both 
calculations  are  fundamentally  identical.  In  these  calculations,  the 
temperature  has  been  determined  on  each  plate  by  taking  the  K  for 
toluene  corresponding  to  the  plate  and  determining  the  temperature 
from  Fig.  9-7.  Thus,  K's  were  determined  for  the  still  and  first  plate 
by  dividing  the  calculated  y  values  by  the  value  of  #,  KT  «  IJT/XT] 
thus  for  the  still,  XT  is  0.744,  and  yT  was  calculated  as  0.856,  so  that 
KT  =  0.856/0.744  «  1.15.  From  Fig.  9-7,  the  temperature  is  116°C. 
at  K  *=  1.15.  A  little  consideration  will  show  that  KT  also  is  equal 
to  I/ Sax,  since  y  =  ax/1,ax}  then  y/x  =  «/2ax;  but  for  the  com- 
ponent Relative  to  which  the  volatilities  are  taken,  a  is  equal  to  1,  and 
y/x  =  K  =  I /'Sax  for  this  component.  This  latter  method  was  used 
for  the  second  plate  upward. 

Continuing  in  this  manner,  the  a's  should  probably  be  shifted  when 
the  temperature  becomes  about  100°C.  Taking  the  new  values  of  a  at 
90°C.  should  be  satisfactory  for  finishing  the  column.  Thus,  no  trial 
and  error  is  needed,  and  only  two  sets  of  a  values  are  employed.  In 
order  to  speed  computations,  further  modifications  can  be  made.  If, 
instead  of  calculating  ym,  one  calculates  Vym,  where  V  is  the  mols  of 
vapor  per  mol  of  residue,  then  simply  adding  Xw  to  these  values  gives 
OmXm+i,  where  Om  is  the  mols  of  overflow  per  mol  of  residue.  Then 
Oaxm+i  is  calculated,  and  Vym+i  =  VOaxm+i/I,Oaxm+i,  which  mate- 
rially shortens  the  time  necessary  per  plate  but  has  the  disadvan- 
tage that  the  actual  x  and  y  values  do  not  appear.  Continuing 
by  this  method,  V  per  mol  of  bottoms  is  180.3/39.9  =  4.52,  and 
O  =  220.2/39.9  =  5.52. 

The  values  of  XQ  are  essentially  those  obtained  previously,  indicating 
that  the  trial-and-error  calculation  to  determine  plate  temperature  is, 
in  general,  not  necessary  and  that  the  Lewis  and  Matheson  method 
when  carried  out  in  such  a  manner  does  not  possess  any  great  obstacles. 
The  use  of  the  relative-volatility  method  also  offers  other  advantages 
than  the  ease  of  determining  the  plate  composition.  Consider  the 
fractionation  in  a  vacuum  column  in  which  the  overhead  pressure  is 
fixed  and  the  pressure  drop  per  plate  is  an  appreciable  percentage  of  the 
total  pressure  causing  the  absolute  pressure  to  vary  widely.  In  gen- 
eral, the  K  values  are  approximately  inversely  proportional  to  the 
pressure  and  therefore  would  vary  with  the  changes  in  pressure  as  well 
as  with  the  changes  in  temperature.  On  the  other  hand,  the  relative 
volatility  is  often  mainly  a  function  of  the  temperature  and  only 
slightly  affected  by  the  moderate  changes  in  pressure.  In  such  cases, 


236 


FRACTIONAL  DISTILLATION 


xw 

2/6 

J 

4.52t/j 

5.52x0  •  x^  4~ 
4.52^/6 

«« 

(5.52az8/7.411)(4.52) 

Ce 
C7 
C8 

0  005 
0.744 
0.251 

0.338 
0  636 
0  027 

1  55 
2.88 
0.12 

1  535 
3.624 
0.371 

3  62 
3.624 
0  167 

2  21 
2  21 
0.098 

7  411 

K  -  5.52/7,411  =  0.746        T6  -  100.7 


5.52z7 

«90 

5.52az7 

4.52^/7  ='(5.52<*z/8.56)4.52 

5.52s8 

c« 

2  215 

2  47 

5.4fr 

2  89 

2  895 

C7 

2.954 

1.0 

2.954 

1  56 

2.304 

C, 

0  349 

0  42 

0  147 

0  078 

0  329 

8  561 

K  =  5.52/8.56  -  0.646'       T7  =  95.8 


'5.52a$8 

4,522/8 

5,52z9 

c« 

C7 

Cs  * 

7  15 
2  304 

'  0  158 

3  36 
1  09 
0.065 

3  365 
1  834 
0  316 

9  592 

K  -  0.577        T*  -  92.2 


C6 
C7 

C8 


0  609 
0  334 
0  057 


it  is  therefore  possible  to  proceed  by  the  a  method,  as  in  the  constant- 
pressure  calculations,  without  troubling  with  the  pressure  variation. 

Ta^r  Acid  Fjractionati&n.  As  another  example  of  such  calculations, 
consider  the  Iractionation  of  a  35  mol  per  cent  phenol,  15  mol  per  cent 
o-djresol,  30  mol  pef  cent  m-cresol,  15  mol  per  cent  xylenols,  and  5  mol 
per  ce;ttt  heavier.  The  overhead  is  to  be  95  mol  per  cent  phenol,  and 
the  phenol  recovery  is  to  be  90  per  cent.  The  still  pressure  will  be  250 
mm.  Hg  abs.,  and  4  mm.  Hg  pressure  drop  will  be  allowed  per  theoreti- 
cal plate.  A  reflux  ratio  0/D  equal  to  10  will  be  employed. 

The  equilibriuAx  data  obtained  by  Rhodes,  Wells,  and  Murray 
(Ref  .  7)  for  this  type  of  system  indicate  that  RaoulVs  law  is  followed, 
and  thus  the  relative  volatilities  are  independent  of  the  pressure  and  a 
function  of  the  temperature  only.  Thus,  the  relative-volatility 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        237 


t- 


r 


363.5 


330.5 


430.5 


mols      % 

.  C6  31.50  95.30 
OC7  1.50  4.55 
mC7  0.05  OJ5 

33.05 


method  will  be  most  suitable  for  estimating  the  number  of  theoretical 
plates. 

The  result  of  over-all  material  bal- 
ances is  given  in  Fig.  9-8.  The  ratio 
of  o-C?  to  ra-C?  in  the  distillate  was 
assumed  as  30  to  1. 

Figure  9-9  gives  the  volatilities 
relative  to  o-cresol  as  well  as  the 
vapor  pressure  of  o-cresol.  In  the 
calculations,  the  temperature  is 
checked  occasionally  by  determining 
the  vapor  pressure  of  o-cresol,  P0,  on 
the  plate.  Since  y0ir  =  P0x0>  then 


where  TT  is  corrected  for  pressure  drop 
in  the  column.  Below  the  feed,  per 
mol  of  bottoms,  the  mols  of  vapor  are 


mols  % 

C6    3.50  524 

oC7    13.50  20.20  • 

mC7  29.95  44.70 

C8    1500  22.40 

R      5.00  7.50 

6695 
FIG.  9-8. 


80 


120  140          160 

Te  mperatu  re ,  deg .  C 

FIG.  9-9. 


238 


FRACTIONAL  DISTILLATION 


5.43,  and  the  mols  of  liquid  are  6.43;  above  the  feed,  the  corresponding 
figures  per  mol  of  distillate  are  11  and  10,  respectively. 


xw 

<*160 

axw 

5.43^  <-  a*iK5.43)/0.685 

6.43zi 

c, 

0.0524 

1  25 

0  0656 

0  521 

0.573 

o-C7 

0.202 

1.0 

0.202 

1  60 

1  80 

n-C7 

0  447 

0  7 

0.312 

2  48 

2  93 

C8 

0.224 

0.44 

0  099 

0.79 

1.01 

R 

0.075 

0.087 

0  006 

0.048 

0.123 

0  6846 

-  250/0,685  -  365  mm. 


165°C. 


6.43.x, 

5.4%, 

6.43*2 

6.43ax2 

5.432/2 

C9 

0.716 

0.775 

0  828 

1  035 

1  045 

o-Cr 

1  80 

1  95 

2.152 

2  152 

2  18 

Wl-C? 

2.05 

2  22 

2.667 

1  865 

1  89 

C8 

0  444 

0  48 

0  704 

0  310 

0  31 

R 

0  Oil 

0.012 

0.087 

0  008 

0  008 

5.021 

5  370 

6.43z3 

6.43ax3 

5.43^/3 

6.43x4 

6.43«x4 

c« 

1.097 

1.37 

1  32 

1  372 

1.72 

o-Cr 

2  38 

2  38 

2.30 

2  50 

2  50 

m-Gt 

2  34 

1  635 

1  575 

2  02 

1  41 

C8 

0.53 

0.233 

0  225 

0  45 

0  198 

R 

0.083 

0  007 

0  007 

0  082 

0  007 

5  625 

5  835 

5.43^4 

6.43z6 

6.43^X6 

5.432/s 

6.43x6 

Cfl 

1.60 

1.652 

2  07 

1  87 

1.92 

o-C7 

2  33 

2.53 

2  53 

2  28 

2.48 

w-C7 

1.31 

1  76 

1  23 

1  11 

1.56 

C8 

0.184 

0.41 

0.18 

0.16 

0.38 

R 

0.006 

0.081 

0.007 

0.006 

0.081 

6.017 

230(6.43)/6.017  -  246 


*  153°C. 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES 


239 


6.43«x6 

5.432/6 

6.43x7 

6.43ax7 

5.431/7 

C6 

2.40 

2.12 

2  17 

2  72 

2.36 

o-Cr 

2.48 

2.20 

2.40 

2  40 

2  08 

W-Cz 

1.09 

0  96 

1.41 

0  99 

0.86 

C8 

0.17 

0  15 

0  37 

0  163 

0.14 

R 

0  007 

0  006 

0.081 

0  007 

0.006 

6.147 

6.280 

6.43x8 

«140 

6.43az8 

5.43^/8 

6.43x9 

6.43<*Xj) 

Co 

2.41 

1  26 

3  04 

2.61 

2  66 

3  35 

0-C7 

2.28 

1  0 

2  28 

1  95 

2  15 

2  15 

w-C? 

1.31 

0  675 

0  88 

0  76 

1  21 

0  82 

C8 

0.36 

0  392 

0.14 

0  12 

0  34 

0  13 

R 

0.081 

0.087 

0.007 

0.006 

0  081 

0  007 

6  347 

6  457 

5.43i/9 

6.43xio 

6.43<*xio 

5.43?yio 

6.43xu 

6.4301X1! 

C6 

2.82 

2  87 

3  62 

3  01 

3.06 

3  86 

o-C7 

1  81 

2.01 

2.01 

1  67 

1  87 

1  87 

WI-C7 

0.69 

1  14 

0  77 

0.64 

1  09 

0  74 

Cs 

0  112 

0  34 

0  13 

0.11 

0.33 

0  13 

R 

0  006 

0.081 

0  007 

0.006 

0.081 

0  007 

6.537 

6  607 

5.43yii 

6.43xi2 

6.43«Xi, 

5.43t/i2 

Ce 
o-C7 

C8 
R 

3  17 
1  54 
0.61 
0  11 
0.006 

3.22 
1  74 
1.06 
0  33 
0.081 

4.06 
1.74 
0.71 
0.129 
0.007 

3.32 
1  42 
0.58 
0  105 
0  006 

6.646 

Po  -  202(6.43)/6.65  -  195; 


146.5°C. 


6.43x18 

6.43«xi3 

y™ 

Ce 

3.37 

4.25 

0  634 

o-C7 

1.62 

1  62 

0  242 

m-C? 

1.03 

0.70 

0.104 

C8 

0.33 

0.129 

0.019 

R 

0.081 

0.007 

0.001 

6.706 

240 


FRACTIONAL  DISTILLATION 


The  ratio  of  phenol  to  o-cresol  in  the  liquid  on  the  thirteenth  plate  is 
essentially  that  in  the  feed,  and  this  plate  was  used  as  the  feed  plate. 
The  calculations  are  then  completed  using  a  basis  of  one  mol  of  dis- 
tillate. On  such  a  basis,  the  operating  line  for  each  component 
I0xn  =  ll^n-i  — •  %D  and  the  remainder  of  the  table  is  set  up  in  this 
manner. 


2/13 

112/13 

10X14 

10c*X14 

Hi/14 

C« 

0  634 

6  974 

6  024 

7  59 

7  54 

o-C7 

0  242 

2  662 

2  617 

2.617 

2.60 

77t-C/7 

0.104 

1  144 

1  14 

0.77 

0  77 

C8 

0.019 

.209 

.209 

0  082 

0  081- 

R 

0.001 

Oil 

.011 

0  001 

0  001 

11  060 

10*» 

lOaXis 

102/15 

10*16 

lOorXie 

H2/16 

10*17 

lOttXiT 

H2/17 

c. 

6.59 

8  3 

8.0 

7  05 

8  89 

8.38 

7.42 

9  36 

8  73 

o-C7 

2  55 

2  55 

2  46 

2  41 

2.41 

2.27 

2.22 

2  22 

2  07 

m-C? 

0  765 

0  516 

0.498 

0.493 

0  333 

0.314 

0.31 

0  209 

0  195 

C8 

0.081 

0  032 

0  031 

0.031 

0  012 

0.011 

0  Oil 

0  004 

0  004 

R 

0.001 

9  X  10-* 

8  X  10~6 

8  X  10-6 

7  X10-* 

7  XlO-« 

7  X10-6 

6  X10-7 

6  X10"7 

11.398 

11.645 

11  793 

10^18 

lOaXis 

112/18 

10*u 

10aXi9 

112/19 

10X20 

10o:X2o 

112/20 

c, 

7  78 

9  81 

9  02 

8.07 

10.2 

9  3 

8  35 

10  55 

9  53 

o-C: 

2.02 

2  02 

1  86 

1.81 

1.81 

1  65 

1  60 

1  6 

1  44 

WrCj7 

0.19 

0  128 

0  118 

0.113 

0  076 

0  069 

0  064 

0  043 

0  0399 

Cs 

0.004 

0  002 

2  X  lO-3 

2  X  10-* 

8  X  10-* 

7  X  10~4 

7  X  10-' 

3  X  10-4 

3X10~4 

11.960 

12.086 

12.193 

P,  -  178(10/11.96)  =  149        Tit  =  138°C. 


10^21 

10o:£2l 

112/21 

10^22 

10a*22 

lly»* 

10x23 

10aX23 

112/23 

Ce 
o-C7 

W-C7 

8.58 
1.39 
0.035 

10.8 

1  39 
0  024 

9.74 
1  25 
0.022 

8.79 
1  20 
0  018 

11.08 
1.20 
0  012 

9.91 
1  07 
0.011 

8  96 
1.02 
0.007 

11.30 
1  02 
4  7  X  10~3 

10.1 
0.91 
4.2  X  10-s 

12  214 

12.292 

12  32 

RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        241 


10z24 

10aX24 

112/26 

10^26 

10«£25 

1  If/26 

10X26 

10tX#28 

2/26 

C6 

9  15 

11  52 

10  24 

9  29 

11  70 

10.37 

9  42 

11  88 

0  952 

o-C7 

0.86 

0.86 

0.76 

0.71 

0.71 

0  63 

0  58 

0  58 

0  047 

12  38 

12  41 

12  46 

154(10/12,38)  *  124 


133°C. 


The  concentration  of  m-cresol  in  the  distillate  is  less  than  the 
assumed  value,  but  a  recorrection  of  this  value  would  not  make  enough 
difference  to  be  significant,  and  a  material  balance  on  this  component 
is  in  essence  satisfied.  The  results  of  the  calculations  are  plotted  in 
Fig.  9-10. 

It  is  interesting  to  consider  what  would  happen  if  the  feed  had  not 
been  introduced  on  the  thirteenth  plate.  This  calculation  has  been 
carried  out  and  the  results  plotted  in  Fig.  9-1 1.  Up  to  the  thirteenth 
plate,  the  results  are  obviously  identical  with  those  given  in  Fig.  9-10; 
but  above  this  plate,  the  change  of  concentration  per  plate  is  much  less 
in  Fig.  9-11.  By  the  twenty-sixth  plate,  all  the  components  have 
become  almost  asymptotic,  and  increasing  the  plates  to  an  infinite 
number  would  make  little  difference  in  the  concentrations  from  those 
for  the  twenty-sixth  plate.  Thus  it  is  impossible  to  obtain  the  desired 
separation  without  having  plates  above  the  feed  plate,  since  the 
asymptotic  ratio  of  phenol  to  o-cresol  is  less  than  the  desired  ratio  in 
the  distillate.  The  limit  to  this  asymptotic  ratio  is  obvious  from  Fig. 
9-1 1 ;  since  the  o-cresol,  m-cresol,  xylol,  and  residue  must  all  flow  down 
the  column,  their  concentrations  cannot  decrease  below  the  value 
necessitated  by  material  balance  for  their  removal  from  the  still. 
Although  the  concentration  of  the  phenol  is  not  limited  by  the  same 
factor  as  the  heavier  components,  it  is  limited  by  the  fact  that  its 
value  cannot  exceed  1  minus  the  sum  of  the  concentration  of  the 
heavier  fractions;  and  since  a  minimum  limit  for  the  heavier  com- 
ponents is  fixed,  a  maximum  for  the  phenol  is  likewise  fixed.  The 
condition  illustrated  in  Fig.  9-11  around  twentieth  to  twenty-sixth 
plate  is  termed  " pinched  in";  i.e.,  conditions  are  so  pinched  that 
effective  rectification  is  not  obtained.  As  soon  as  the  feed  plate  is 
passed,  this  pinched-in  condition  would  be  relieved,  since  the  heavier 
components  would  decrease  rapidly,  as  in  Fig.  9-10,  thereby  allowing 
the  phenol  to  increase. 

At  a  lower  reflux  ratio,  the  enrichment  per  plate  would  be  reduced, 
and  more  plates  required  for  a  given  separation  Figure  9-12  gives  the 


242 


FRACTIONAL  DISTILLATION 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        243 

results  for  the  same  design  conditions  as  Fig.  9-10,  except  that  0/D  was 
7  instead  of  10.  The  number  of  plates  increases  from  26  to  35.  It  is 
to  be  noted  that  the  asymptotic  values  of  the  concentrations  of  the 
heavier  components  in  the  lower  part  of  the  column  also  increase; 
this  results  from  the  fact  that  there  is  less  overflow  in  this  section  of 
the  tower. 


24 
22 
20 
18 
16 
Sl4 

4- 

£  12 

Q. 

10 
8 
6 
4 
Z 

c? 

i 

-Creso/ 

ry/eA 

o/ 

\ 

Residue 

) 

L 

J 

\ 

1 

\ 

/ 

/ 

7-C 

resol 

\ 

/ 

\ 

V 

/Phenol 

^ 

V  A 

\ 

X 

\ 

^ 

X 

^/ 

'  V 

sswh- 

"•"*»"    .-». 

bam 

.c^ 

*^ 

^^ 

^/ 

} 

0.01         0;02         0.04  0.06      O.I  0.2 

Mol  Fraction  in  Liquid 
FIG.  9-11. 


0.4     0.6 


c 


Minimum  Theoretical  Plates  at  Total  Reflux.  The  minimum  num- 
ber of  theoretical  plates  for  a  given  separation  is  obtained  at  total 
reflux,  the  same  as  for  a  binary  mixture.  This  minimum  can  be  calcu- 
lated by  the  stepwise  method,  using  the  operating  line  y  =  x  for  each 
component  for  both  sections  ,of  the  column.  Such  a  calculation  will 
give  the  concentration  and  conditions  through  the  tower.  The 
assumptions  made  on  page  175  for  the  calculation  of  the  minimum 
number  of  plates  in  the  development  of  Eq.  (7-52)  apply  to  any  two  of 
the  components  of  a  multicomponent  mixture,  and  by  its  application 
this  limiting  condition  can  be  calculated.  In  general,  when  applying 
the  total  reflux  equation  to  a  multicomponent  mixture,  it  is  desirable 
to  use  the  two  components  whose  concentrations  are  most  accurately 
known  in  the  distillate  and  residue,  and  most  often  these  two  com- 
ponents are  the  kev  components.  However,  the  equation  applies  to 


244 


FRACTIONAL  DISTILLATION 


^       CM 
fO       fO 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        245 

any  two  components  in  the  mixture.  The  same  limitations  as  to  the 
constancy  of  relative  volatility  would  apply  as  for  the  binary  mixtures* 
Feed-plate  Location.  The  criterion  for  the  optimum  location 
of  the  feed  plate  is  that  the  relative  enrichment  of  the  key  components 
should  be  a  maximum.  As  with  binary  mixtures,  the  feed  plate  corre- 
sponds to  the  step  that  passes  from  one  operating  line  to  the  other. 
The  change  from  one  operating  line  to  the  other  should  be  made  just 
as  soon  as  it  will  give  a  greater  enrichment  than  continuing  on  the  same 
operating  line.  In  coming  up  from  the  still,  the  feed  plate  is  the  last 
step  on  the  lower  operating  line;  calling  this  the  nth  plate,  and  the  key 
component  vapors  entering,  yik(n~\}  and  yhk(n-i),  for  the  light,  or  more 
volatile,  and  heavy,  or  less  volatile,  components,  respectively,  the  fore- 
going criterion  states  that  if  the  nth  plate  is  the  optimum  position  for 
the  feed,  then  the  x  ratio  on  this  plate  should  be  greater  when  calcu- 
lated by  the  lower  operating  line  from  the  yn~i  values  than  by  the 
upper  operating  line,  or 


\Xhk/n          yhk(n-l)   -T    (W/Vm)Xwhk          yhk(n-l)    — 

where  yn-i  =  mol  fraction  in  vapor  from  plate  n  —  1  when  feed  is 

added  to  plate  n 

y'n-i  =  raol  fraction  in  vapor  from  plate  n  —  1  when  feed  is 
added  to  plate  n  —  1 

For  a  binary  mixture,  yn-i,  would  equal  y'n_^  assuming  that  the 
bottoms  concentration  is  kept  constant.  However,  the  two  values  are 
not  equal  for  multicomponent  mixtures  because  of  the  presence  of 
the  components  lighter  and  heavier  than  the  key  components.  Owing 
to  the  presence  of  these  components,  the  total  vapor  is  not  available  for 
fractionating  the  key  components.  In  order  to  reduce  the  interference 
of  the  components  lighter  than  the  light  key  component,  it  is  desirable 
to  utilize  the  lower  operating  line  to  a  higher  value  of  the  ratio  of  the 
key  components  than  would  be  the  case  for  yn~i  =  y'n-i-  Owing  to 
the  interference  of  the  heavy  components,  it  would  be  desirable  to 
change  to  the  upper  operating  line  at  a  lower  ratio  of  the  key  com- 
ponents. However,  as  a  first  approximation,  it  will  be  assumed  that 
i/n_i  ss  y^  and  a  correction  for  this  assumption  will  be  made  later. 

Combining  Eq.  (9-4)  with 

Fn  -  Vm  +  (P  +  1)F 
and 

Wxw  +  DXD  —  FZP 


246  FRACTIONAL  DISTILLATION 

gives 


~*  -  1  ) 
— 


Also,  by  this  criterion  the  x  ratio  on  the  (n  +  l)th  plate  should  be 
greater  when  calculated  by  the  upper  than  by  the  lower  opera  ting  line: 


(Xlk 
Xhk 


—    (D/Vn)XDhk          y'nhk  +   (W/Vm)Xwhh 

where  yn  =  mol  fraction  in  vapor  from  plate  n  when  feed  is  added  to 

plate  n 
y'n  =  mol  fraction  in  vapor  from  pla/te  n  when  feed  is  added  to 

plate  n  +  1 
Again  using  yn  =  y'n  gives 

ZFlk  + 


The  right-hand  sides  of  Eqs.  (9-5)  and  (9-7)  are  equivalent  and 
equal  to  (xik/Xhk)t  as  given  by  the  intersection  of  the  operating  line; 
thus,  since  n  is  now  the  optimum  feed  plate,  the  subscript  may  be 
changed,  and  the  criterion  for  the  feed-plate  step  becomes 


^  > 

~  W/  v    ; 

where  (xik/xhk)i  is  the  ratio  of  the  key  components  as  given  by  the 
intersections  of  the  operating  lines.  However,  it  should  be  emphasized 
that  the  feed  plate  does  not  necessarily  step  across  the  intersection  of 
the  operating  lines,  as  it  does  for  a  binary,  but  simply  that  the  ratio  of 
the  keys  for  the  optimum  feed-plate  step  passes  over  the  ratio  of  the 
values  given  by  the  operating-line  intersections.  The  absolute  value 
of  both  key  components  may  be  several  times  the  values  given  at  the 
intersection,  provided  the  ratio  satisfies  Eq.  (9-8). 

The  derivations  of  Eqs.  (9-5)  and  (9-7)  neglected  the  effect  of  the 
changing  concentrations  of  the  light  and  heavy  components.  The 
light  components  have  a  relative  constant  concentration  above  the 
feed  plate,  and  this  fact  can  be  used  to  calculate  their  value.  Thus, 


and,  assuming 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES         247 
giving 


n          Vn  -   (On/Kl) 

As  shown  on  page  228,  the  concentration  of  a  light  component 
changes  approximately  by  a  factor  of  Om/VmKi  per  plate  below  the 
feed  plate.  Thus  the  change  in  the  concentration  of  a  light  component 
per  plate  at  the  feed  plate  is  * 

-  (Q./F.JC,) 


VJt 

and  for  most  cases  DXDI  = 

Let  Aj  equal  the  sum  of  the  changes  per  plate  for  all  light  compo- 
nents, then 

(9-10) 


where  ^   is  the  sum  of  the  term  for  all  components  more  volatile  than 


the  light  key  components. 

In  a  similar  manner  the  sum  of  the  changes  in  the  vapor  concentra- 
tions for  the  heavy  components  per  plate  at  the  feed  plate,  AA,  can  be 
calculated. 


where   ^   is  the  summation  for  all  terms  less  volatile  than  the  heavy 

**+ 
key  components, 

The  corrections  due  to  these  changes  can  be  utilized  (1)  to  calculate 
terms  to  be  added  to  the  intersection  ratio  (#«,)/#**)*  °r  (2)  to  modify 
the  expression  for  the  intersection  ratio.  The  relations  can  be  formu- 
lated so  that  both  methods  give  essentially  the  same  optimum  ratio  for 
the  key  components  at  the  feed  plate.  The  latter  method  is  believed 
to  be  the  more  convenient,  and  using  <£  as  the  optimum  ratio,  an 
approximate  expression  is 


248 


FRACTIONAL  DISTILLATION 


In  using  this  expression,  it  is  recommended  that  K  i  and  KH  be  cal- 
culated as  (oti/aik)  and  («/»/«*&),  respectively.  This  expression  is  simi- 
lar to  that  for  the  intersection  ratio,  but  if  A*  is  large,  <£  will  be  larger 
than  (xik)/(Xhk)%y  if  A*  is  large,  <f>  will  be  smaller.  Using  these  correc- 
tions, the  optimum  feed-plate  location  is  such  that 


(**)   ^  +  <;  (V) 

\W/  \W/ 


Optimum  Feed-plate  Location.     The  use  of  Eq.  (9-12)  will  be  illustrated  by  the 
examples  already  considered. 

1.  Benzene-Toluene-Xylene  example 


Kh 


~  -  0.18; 


-  120.2; 

-  180.3 


O 


220.2 


220.2 


0593  ~~  0  29> 


29-7 


30 


In  this  case  <#>  is  essentially  the  same  as  the  intersection  ratio  of  the  operating 
line  which  was  2.0. 


2.  Phenol-Cresol 
D  -  33.05; 
mC7:    #,  =9^-0.54 


330.5;        On  -  430.5; 


3.6 


1.26 
0.392 
1.26 

_  0.087 
R.    Kh  -    |  2Q 


C8:     #* 


430.5 


-  0.31 

-  0.07 

0.54(30) 


1  - 


0.54(363.6) 
330.5 


1  - 


0.54(363.6) 


+ 


430.5 


4-  0.07(5) 


0.31(363.6) 

1 330.5 

0.31(363.6) 

430.5 

0.07(363.6) 
330.5 


1  - 


0.07(363.6) 
430.5      . 


0.039 


35 


^  » 


(sm  ~  0 3-5 


15  +  (sm  ~  0 13-5 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        249 


As  compared  to  the  intersection  ratio  of  8Ms  *"  2.33.  In  the  stepwise  calcula- 
tion, the  ratio  of  the  key  components  in  the  feed  plate  was  2,08  and  2.31  on  the 
plate  above.  Thus  the  feed-plate  composition  utilized  agrees  with  Eq.  (9-12) 
satisfactorily. 

In  the  foregoing  derivations  the  feed  could  be  vapor,  liquid,  or  a 
mixture  of  the  two,  but  it  was  assumed  that  the  vapor  and  liquid  leav- 
ing the  feed  plate  were  in  equilibrium.  In  the  case  of  an  all-vapor  feed 
that  mixes  with  the  vapor  from  the  feed  plate  but  does  not  react  with 
the  liquid  on  the  plate,  a  similar  derivation  gives  (Ref  .  1) 


{Xhk/f+i         \Xhk/%         \XhkJ 

This  indicates  the  ratios  of  the  concentrations  of  the  key  components 
on  the  plate  above  the  feed  plate,  and  the  plate  above  that  should 
straddle  the  intersection  ratio  for  the  key  components.  Or,  if  the 
nomenclature  is  changed  such  that  the  feed  plate  is  the  first  plate  with 
which  this  feed  vapor  reacts  (i.e.j  the  plate  above)  and  not  the  plate 
into  which  it  enters,  then  the  criterion  becomes  the  same  as  before. 

Minimum  Reflux  Ratio.  As  in  the  case  of  binary  mixtures,  there 
is  a  reflux  ratio  below  which  it  is  not  possible  to  obtain  the  desired 
separation  of  a  multicomponent  mixture  even  when  an  infinite  number 
of  plates  is  used.  The  calculation  of  this  minimum  for  a  multicom- 
ponent mixture  is  much  more  involved  than  for  the  corresponding 
binary  mixture.  The  condition  of  the  minimum  reflux  ratio  requires 
that  to  perform  the  given  separation  an  infinite  number  of  plates  must 
be  needed,  which  means  that  there  must  be  a  pinched-in  region  where 
there  are  a  large  number  of  plates  having  the  same  composition,  but 
for  multicomponent  mixtures  of  normal  volatility  this  region  usually 
does  not  occur  at  the  feed  plate  as  it  does  in  the  case  of  binary  mix- 
tures. Under  this  condition  usually  a  relatively  few  plates  above  the 
feed  serve  to  reduce  the  concentrations  of  the  components  less  volatile 
than  the  heavy  key  component  to  negligible  values,  and  then  a  true 
pinched-in  condition  does  occur  with  only  the  heavy  key  and  more 
volatile  components  present.  Likewise,  below  the  feed  plate  a  rela- 
tively few  plates  reduce  the  concentrations  of  the  components  more 
volatile  than  the  light  key  component  to  negligible  values,  and  a  true 
pinched-in  condition  occurs  with  only  the  light  key  and  less  volatile 
components  present.  Thus  the  tower  operating  at  the  minimum 
reflux  ratio  might  be  considered  as  being  composed  of  five  sections 
(Ref.  2):  -  v 

(1)  Starting  at  the  still  the  only  rtrmrmruients  present  in  significant 


250  FRACTIONAL  DISTILLATION 

amounts  are  the  light  key  and  the  less  volatile  components,  and  in 
proceeding  up  the  tower  in  this  bottom  section,  the  concentration  of 
the  light  key  component  increases  relative  to  the  concentrations  of  the 
heavy  key  and  heavier  components. 

2.  Above  section  1  is  a  pinched-in  region  where  the  concentrations  of 
the  light  key  and  the  less  volatile  components  are  all  con^iant,  and  an 
infinite  number  of  plates  is  required  to  produce  a  finite  change  in  the 
plate  composition. 

3.  Next  there  is  an  intermediate  region  where  the  concentrations  of 
the  components  less  volatile  than  the  heavy  key  component  decrease 
to  negligible  values  and  where  the  concentrations  of  the  components 
more  volatile  than  the  light  key  component  increase  to  significant 
values. 

4.  Above  section  3  is  another  pinched-in  region  where  the  concen- 
trations of  the  heavy  key  and  the  more  volatile  components  are  all  con- 
stant, and  an  infinite  number  of  the  plates  is  required  to  produce  a  finite 
change  in  plate  composition. 

5.  A  section  exists  where  the  concentration  of  the  heavy  key  com- 
ponent decreases  relative  to  the  concentration  of  the  more  volatile 
components  until  the  overhead  composition  is  obtained. 

Actually  there  is  no  sharp  line  of  demarcation  between  these  five 
sections,  but  this  division  serves  as  a  useful  picture  for  considering  the 
case  of  the  minimum  reflux  ratio.  The  feed  to  the  fractionating  column 
would  be  introduced  on  some  plate  in  intermediate  section  3,  and  the 
true  criterion  for  the  minimum  reflux  ratio  should  be  based  on  matching 
the  ratio  of  the  concentrations  of  the  key  components  above  and  below 
the  feed  plate  under  conditions  such  that  a  pinched-in  section  occurs 
both  above  and  below  the  feed  plate.  For  mixtures  of  normal  vola- 
tility, a  pinched-in  region  in  only  one  section  does  not  necessarily  mean 
that  an  infinite  number  of  plates  would  be  required  to  perform  the 
desired  separation  at  the  reflux  ratio  under  consideration,  since  by 
relocating  the  feed  plate,  such  as  to  shift  the  ratio  of  the  concentra- 
tions of  the  key  components  at  this  plate,  the  section  that  was  not 
limited  could  be  made  to  do  more  separation  and  thereby  relieve  the 
load  on  the  pinched-in  section.  In  other  words,  for  mixtures  of  normal 
volatilities  the  condition  of  the  minimum  reflux  ratio  is  not  determined 
by  either  the  fractionation  above  or  below  the  feed  plate  alone,  but  is 
determined  such  that  the  separation  is  limited  both  above  and  below 
the  feed.1  The  conditions  in  the  intermediate  feed  section  lead  to 

1  For  mixtures  with  abnormal  volatilities  the  pinched-in  condition  may  be  due  to 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        251 

calculational  difficulties  because  the  concentrations  are  not  constant  in 
this  region.  Actually  in  this  region  the  ratios  of  the  concentrations  of 
the  key  components  generally  change  in  the  opposite  direction  from 
that  desired  for  the  separation;  thus,  proceeding  up  the  column  from 
the  feed  plate  at  the  minimum  reflux  ratio,  the  ratio  of  the  concentra- 
tions of  the  light  key  to  the  heavy  key  component  decreases  instead  of 
increases.  No  satisfactory  method  of  estimating  the  extent  of  this 
"  retrograde"  rectification  has  been  developed.  However,  it  is  rela- 
tively simple  to  calculate  the  plate  composition  for  a  region  where  the 
concentrations  are  the  same  on  successive  plates,  and  such  calculations 
can  be  made  the  basis  for  estimating  the  minimum  reflux  ratio. 

Case  I.  It  can  be  assumed  that  the  concentrations  of  all  com- 
ponents are  constant  for  a  number  of  plates  above  and  below  the  feed 
plate.  This  requires  that  the  pinched-in  condition  between  the  key 
components  must  occur  with  all  the  components  present  in  significant 
amounts;  actually,  as  pointed  out  before,  only  certain  of  the  compo- 
nents are  present  at  the  pinched-in  condition.  The  presence  of  these 
extra  components  makes  the  separation  more  difficult,  and  for  that 
reason  the  minimum  reflux  ratio  calculated  on  the  basis  of  the  assump- 
tions of  Case  I  will  be  equal  to  or  greater  than  the  true  minimum 
reflux  ratio. 

Case  II.  Alternately  it  can  be  assumed  that  the  ratio  of  the  concen- 
trations of  the  key  components  for  Jbhe  pinched-in  region  below  the  feed 
plate  (sectio^JjIc^  the  same  ratio  for  tlie  pinc£ed-in 


region  above  the  feed  plate  (section  4),  which  amounts  to  neglecting 
tfieTSlerme^at^  region^section  3)  .  However,  for  the  actual  operation 
sectiori"Sn^usTBepresent  and  in  this  section  the  ratio  of  the  concentra- 
tions of  the  key  components  changes  in  the  opposite  direction  to  that 
desired  for  the  separation;  for  that  reason,  the  assumptions  of  Case  II 
correspond  to  an  easier  separation  than  the  actual  case,  and  the  mini- 
mum reflux  ratio  calculated  by  these  assumptions  will  be  equal  to  or 
less  than  the  true  minimum  reflux  ratio. 

Calculation  of  the  minimum  reflux  ratio  for  these  two  cases  will  give 
limits  for  the  true  minimum  reflux  ratio.  To  evaluate  these  cases,  it 
is  necessary  to  calculate  the  concentrations  of  the  various  components 
for  the  pineh<gd-m  or  constant  composition  regions.  Such  composi- 
tions are"easlly  cjH<^  tKe  reTaHve^volatility,  a,  and 

the  fact  that  the  ratio  of  the  concentrations  of  any  two  components  are 

a  tangent  contact  between  the  equilibrium  curve  and  the  operating  line,  and  the 
separation  will  not  be  limited  both  above  and  below  the  feed  plate. 


252  FRACTIONAL  DISTILLATION 

the  same  on  successive  plates;  thus  for  the  volatile  components  above 
the  feed  plate, 

J2L  £L  =  !L  =  (°/D)Xi  +  XDI  -  2i  \  (v/D)yi 

othk  xhk  ™  yhk      (0/D)xhk  +  xDhk      <*hk  [  (V/D)yhk  -  > 

giving 

_  ahkxDi(D/0)  _ 


^  -  «»  +  a, 
and 


01  -  <*A* 
and,  for  less  volatile  components  below  the  feed  plate, 

Xwtk 


-  xwh         ah((Vm/W)yh  +  xwh] 
gvng 


'  \0m)  xlk 


and 

.  -  (9_ig) 


In  these  equations  the  last  term  of  the  denominator  is  usually  small 
relative  to  the  other  factors  and  can  be  neglected  for  the  first  estima- 
tions, making  only  the  flow  quantities,  relative  volatilities,  and  termi- 
nal concentrations  necessary  for  the  calculation  of  the  asymptotic 
values.  Corrections  can  then  be  made  for  the  last  term  in  the  denomi- 
nators, but  usually  this  is  not  necessary. 

These  equations  can  be  used  to  calculate  the  concentrations  of  the 
key  components  for  evaluating  the  minimum  reflux  ratios  for  Cases  I 
and  II,  which  would  involve  matching  the  ratio  of  the  concentrations 
of  the  key  components  above  the  feed  plate  with  the  same  ratio  below 
the  feed  plate.  This  leads  to  a  quadratic  solution  for  the  minimum 
reflux  ratio.  Thus  for  Case  I  the  ratio  of  the  key  components  above 
and  below  the  feed  plate  are  equated,  and  allowance  is  made  for  all 
components  at  their  asymptotic  values  in  both  sections. 


xn 


y  XK  -  y 


Xhk  /n       1  -  xn  -       xh  - 


RECTIFICATION  Of1  MULTICOMPONENT  MIXTURES         253 
On  substituting  the  values  of  Eqs.  9-13  and  9-15  and  rearranging, 


0\  --  (9-17) 


In  Case  II,  below  the  feed  plate  only  the  light  key  and  less  volatile 
components  are  considered;  above  the  feed  plate,  only  the  heavy  key 
and  more  volatile  components  are  considered  : 


Xh 


—  Xhk  —  y 


_   Xlk   - 


Y  Xi 


The  solution  of  this  equation  is  the  same  as  Eq.  (9-17),  but  instead  of 
/   and   J,    I'    and   «/',    respectively,    are   used.     Where 


**    i  r       JL  (X»**\(D\\ 

C*  =  -     aik  —  oihk  +  an  [  —  -  J  I  77  j 
XDikl  \XM/\U/I 

D  =  mols  of  distillate  per  unit  time 


F  —  mols  of  feed  per  unit  time 


T       pF    .    aik 
I  =        +- 


•2[- 

»    I   C^ZA  """"" 

h  l_ 


/  =  ^  +  aWl^)l^r 


(o*\  V  r 

i»/WZ/ 

*  r*~ 


a, 


254  FRACTIONAL  DISTILLATION 

Xwh 


(See  nomenclature  at  the  end  of  this  chapter.) 

These  equations  should  not  be  applied  to  mixtures  having  minimum 
reflux  ratios  determined  by  a  tangent  contact  in  one  section  of  the 
tower  only.  It  is  customary  to  use  relative  volatility  values  corre- 
sponding to  approximately  the  feed-plate  temperature.  Actually 
some  of  the  values  correspond  to  the  pinched-in  region  above  the  feed 
and  the  others  to  the  lower  section,  but  usually  this  refinement  is  not 
made  although  it  can  be  done  if  necessary. 

In  addition  to  Eq.  (9-17),  it  is  possible  to  develop  a  large  number  of 
alternate  relations.  The  most  helpful  of  those  are  obtained  by  equat- 
ing the  ratio  of  the  key  components  at  the  pinched-in  region  to  the 
optimum  feed-plate  ratio. 

Matching  the  optimum  feed-plate  ratio,  4,  with  the  pinched-in 
region, 

yik\  _  [  (0/D)xik  +  XDIJC  1  aw 
L  (0/D)xhk  +  xDhk]  alk 


and  solving  for  0/D, 


0   __    (cthkXplk/<t>)    •"" 

D 


Equation  (9-18)  is  applicable  to  either  binary  or  multicomponent 
nixtures,  and  the  problem  is  one  of  evaluating  XM. 
Tor  Case  I: 


V 

-  xik  -  2,  xi  - 


1  -  Ssi  - 

Xh  =  — 


For  Case  II: 

1  - 


-  1  ~  *»  -        Xl  "" 


TTT 
or,  in  general, 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES         255 

where  ft  =  1  f  or  a  binary 

-  1  -  2xi  for  Case  II 

=  1  —  2xi  —  Zxh  for  Case  I 

The  terms  2xt  and  2xA  can  be  evaluated  by  Eqs.  (9-13)  and  (9-15)  to 
give 

Oihk%Dlk 


( 
B 


(9-20) 


The  first  term  is  the  minimum  reflux  ratio  for  a  binary  mixture  with 
the  same  ratio  of  key  components  in  the  distillate  and  <f>.  Equation 
(9-20)  can  be  written 


(g)  -( 

\/>/min         \D 


Equation  (9-20)  involves  no  trial  and  error  if  On  =  Ow,  but  it  does 
for  other  cases.  By  equating  the  liquid-phase  ratios  instead  of  the 
vapor-phase  ratios  of  Eq.  (9-18),  one  obtains 


(n 
g 
-'J 


(9.22) 

y 


This  equation  requires  no  trial  and  ^error  for  F»  =  Fm.  The  first 
bracket  again  corresponds  to  the  binary  mixture  case  for  the  same  <f> 
and  ratio  of  key  components  in  the  distillate. 

Equations  (9-20)  and  (9-22)  are  given  for  Case  I,  and  for  Case  II  the 
last  term  involving  a;^  is  eliminated. 

Similar  equations  can  be  derived  by  equating  the  ratio  of  key  com- 
ponents below  the  feed  plate  to  0.  Thus, 


^\0ni          <1  + 


256  FRACTIONAL  DISTILLATION 

and 
0\  VnW 


xwlk\  /1 

~  -T)  V 


These  two  equations  correspond  to  Case  I;  Case  II  is  obtained  by 
omitting  the  last  term  involving  XDI  in  each  equation.  The  bracketed 
terms  are  again  the  minimum  reflux  ratios  for  the  binary  for  the  same 
<t>  and  ratio  of  key  components. 

If  there  are  no  components  heavier  than  the  heavy  key  component, 
Eqs.  (9-20)  and  (9-22)  should  give  essentially  the  true  minimum  reflux 
ratio;  if  there  are  no  components  lighter  than  the  light  key,  Eqs.  (9-23) 
and  (9-24)  should  give  the  desired  answer.  In  case  both  light  and 
heavy  components  are  present,  the  choice  should  be  based  on  the  rela- 
tive size  of  the  terms  involving  the  heavy  and  light  components.  If 
the  term  involving  the  summation  of  the  heavy  terms  is  greater  than 
the  summation  of  the  light  terms,  then  Eqs.  (9-23)  and  (9-24)  are  pre- 
ferred to  Eqs  (9-20)  and  (9-22),  and  vice  versa.  The  choice  between 
Eqs.  (9-20)  and  (9-22)  and  between  Eqs.  (9-23)  and  (9-24)  is  purely  a 
matter  of  convenience. 

In  the  use  of  Eqs.  (9-17)  to  (9-24)  the  values  of  the  concentration 
needed  were  calculated  by  difference.  For  example,  in  deriving  Eq. 
(9-17),  the  concentration  of  the  heavy  key  component  was  calculated 
as, 


Instead  of  this  relation,  it  is  possible  to  use 


and  other  equations  would  be  obtained  but  they  give  essentially  the 
same  results  (Ref.  8). 

As  a  result  of  the  various  approximations  involved,  a  large  number 
of  approximate  equations  can  be  derived.  For  the  general  case,  none 
of  them  gives  the  exact  answer.  The  ones  given  here  have  proved  to 
be  of  real  utility.  Equation  (9-17)  is  probably  the  most  general  and, 
using  the  proper  relative  volatility,  it  should  give  answers  for  Cases  I 
and  II  that  are  above  and  below  the  true  minimum  reflux  ratio.  The 
other  equations  presented  involve  an  approximate  expression  for  the 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES         257 

ratio  of  the  key  components  at  the  feed  plate.  It  is  therefore  not  so 
certain  that  the  upper  and  lower  values  will  bracket  the  true  minimum 
reflux  ratio  but  in  most  cases  they  do.  Equations  (9-20)  to  (9-24)  are 
in  general  considerably  easier  to  employ  than  Eq.  (9-17)  and,  for  that 
reason,  are  more  commonly  used.  When  it  becomes  possible  to  pre- 
dict exactly  the  extent  of  the  retrograde  fractionation  in  the  feed-plate 
section,  it  will  be  possible  to  develop  exact  equations  for  the  minimum 
reflux  ratio. 

The  foregoing  derivations  were  made  on  the  assumption  that  there 
were  no  components  heavier  than  the  heavy  key  components  at  the 
upper  pinched-in  region  and  no  components  lighter  than  the  light  key 
components  at  the  lower  pinched-in  region.  This  is  not  completely 
true.  If  there  are  components  only  slightly  heavier  than  the  heavy 
key  components  or  only  slightly  more  volatile  than  the  light  key,  it  is 
possible  for  them  to  be  present  at  the  pinched-in  region.  This  can  be 
shown  as  follows: 

Consider  the  pinched-in  section  below  the  feed  plate,  then  xm  =  xm+i 
for  all  components  present  and 

=  Om+ixm+i  -  Wxw 
=  Kxm 


and  combining  and  solving  for  xm  the  value  at  the  pinched-in  region, 

-       (w/°^x^ 
Xm  ""  1  -  (VmK/Om) 

For  the  light  key  and  the  heavier  components  the  value  of  xm  is  posi- 
tive and  finite,  which  necessitates  VmK/Om  being  less  than  1.0.  For 
the  light  key  components  Wxw/0m  is  usually  small  in  comparison  to 
Xm  which  results  from  VmK/Om  being  only  slightly  less  than  1.0.  If 
a  more  volatile  component  has  a  K  value  significantly  larger  than  K^, 
then  for  that  constituent  VmK/Om  would  be  greater  than  1.0  and  the 
only  solution  for  the  above  equation  is  xw  ~  0  (assuming  that  this 
component  is  not  being  added  below  this  region  which  could  give  nega- 
tive values  of  xw).  The  zero  value  for  xw  was  the  condition  assumed 
for  the  light  components  in  the  derivations  of  Eqs.  (9-17)  to  (9-24), 
and  in  most  cases  it  applies  to  all  components  having  K  values  20  per 
cent  or  more  greater  than  Km.  However,  it  is  possible  for  a  compo- 
nent to  have  a  K  value  only  slightly  greater  than  Kik  and  such  that 
VmK/Om  is  still  less  than  1.0.  This  type  of  component  could  appear 
at  the  lower  pinched-in  region  in  significant  amounts,  and  its  concen- 


258  FRACTIONAL  DISTILLA  TION 

tration  should  be  included  in  calculations  for  Case  II.  Case  I  equa- 
tions have  already  included  a  provision  for  such  a  component. 

The  derivation  of  the  minimum  reflux  ratio  equations  does  not  take 
into  account  the  possibility  of  components  intermediate  in  volatility 
to  the  two  key  components.  When  such  components  are  present,  they 
distribute  themselves  between  the  distillate  and  the  bottoms,  and  dis- 
tribution is  a  function  of  the  reflux  ratio;  i.e.,  the  percentage  of  such  a 
component  that  goes  overhead  will  be  different  at  total  reflux  than  at 
the  minimum  reflux  ratio.  Such  components  can  be  included  in  the 
minimum  reflux  ratio  equations  either  as  light  or  heavy  components 
but  to  do  so  requires  their  concentration  in  the  distillate  or  bottoms. 
It  is  suggested  that  this  distribution  be  made  such  that  the  values  of 
their  mol  fraction  as  calculated  by  Eqfc.  (9-13)  and  (9-15)  are  equal. 

These  equations  will  be  applied  to  some  of  the  previous  examples. 

Minimum  Reflux  Ratio  Examples.  1.  Benzene-Toluene-Xylene  (page  219).  The 
key  components  are  benzene  and  toluene,  and  the  design  conditions  are  given 
below.  Since  there  are  no  components  lighter  than  the  light  key  and  since 
Vn  »  Fm,  Eq.  (9-24)  will  be  employed. 

F  -  100  mols;        W  -  39.9  mols;        D  -  60.1  mols 
otik  «  2.5;        ahk  »  1.0;        ah  =-  0.45 
4>  =  1.98  (see  page  248) 


rio/071i       0>005\ 

/o\  39.9  LOV°-744     Lfl8; 

Wmin   +         ~"  60.1  L  2.5    -   1.0 


[1  -f  2.5(1.98)] 
39.9  ["0.45(0.251)1 

1.0 

This  value  should  be  nearly  exact  because,  in  addition  to  the  key  components 
only  heavy  components  were  present  and  in  small  amounts.  The  value  of  <f>  used 
was  for  a  reflux  ratio  (0/D)  »  2.0  and  should  be  rechecked. 

_  0-18(120.2) 
0.18  „,.,  60.1 


160.1 


30-f  (K-K  -1)29 


0.992       V 


so  the  assumed  value  was  satisfactory. 

As  a  comparison,  use  Eq.  (9-17)  and  neglect  the  small  correction  terms.    Assume 
-  1-0. 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES 
I 


259 


E  - 
G  - 

Tf     ess 


0.744 
2.5  _ 
2.02  ~ 

1.0 
1.505 
-100 

39.9 


(2.5  -  1.0) 
«  1.24 
*  0.665 


2.02 


/-0665+25r39-9(6(U)'U     °'251    ^-077 

«/     — •   \j, \j\jo    —p  tu,^j  I    rti-k'"'  1  / -t  r»rk~"T~\    I    \  o    r  i    n  f     "~" 


I7  =  -0.84 
J'  -  0.665 


Case  I: 


39.9 
60.1 


(-0.84)  4-0.77  -  0.213 


-OH  -  -0.84(0.77)  -  1.24(0.665)  -  -1.47 
0.213 


/ON 
\D) 


(-1.47) 


1.09 


The  calculated  value  is  close  enough  to  the  assumed  value  of  1.0,  and  the  trial 
will  not  be  repeated. 
Case  II: 

The  calculation  is  the  same  except  that  /'  and  J'  are  employed. 


1.01 


In  this  example,  the  two  cases  give  approximately  the  same  result,  but  the 
Case  II  value  should  be  essentially  correct  because  no  light  components  were 
present. 

Values  obtained  with  the  other  equations  are  summarized  in  the  following  table: 


(O/ 
Case  I 

Case  II 

Eq.  (9-24) 

1.0 

1  0 

Eq.  (9-17) 

1.09 

1.01 

Eq.  (9-23) 

1.02 

1  02 

Eq.  (9-20) 

1.07 

0  97 

Eq.  (9-22) 

1  0 

0.96 

2.  Phenol-Cresol  (page  236). 

F,«  100  W  « 

F«  -  Vm,        aik  -  1.26,  aM 

a»  -  0.087;        <A 


66.84        D  -  33.16 

-  1.0,        awC7  -  0.66, 

-  2.26  (for  ^  - 


-  0.39 


260 

Using  Eq.  (9-24), 


'FRACTIONAL  DISTILLATION 


Q\  5^84  (l.O  (0.202  -|g)  [1+2.26(1.26)]) 

D/min  *  33.16  I  1.26  -  1  J        l 

66.84  [-0.66(0.444)        0.39(0.224)        0.087(0.075)  1  _  _ 

33.16  Li.  26  -  0.66  "*"  1.26  -  0.39  "*"  1.26  -  0.087  J  ~~  "*"  ~ 


Rechecking  4  at  (0/D)  «  5.6, 

A  _  0.54(219) 
0.54(30) 


JL_ 
286 


186 


1  


0.54(219) 


286 


1  - 


+  0.31(15) 


0.31(219) 
186 


1  - 


0.31(219) 


0.07(5)  - 


286 

35  +  (eg  -  0  3'5 


1  - 


0.07(219) 

186 
0.07(219) 

286 


0.05 


15  + 


016  ~0 


2.24 


13-5 


This  is  close  enough  to  the  value  of  2.26  employed  so  that  no  recalculation  is 
necessary.  In  other  cases,  such  as  an  all  vapor  feed,  the  value  of  <£  varies  much 
more  with  the  reflux  ratio. 

3.  If,  in  the  preceding  example,  the  feed  condition  had  been  such  that  0,,  «  Om, 
a  higher  reflux  ratio  would  have  been  required.  As  a  first  assumption,  (0/D)mm 
will  be  taken  as  7.0 


On 


Calculating  < 


232 


232  »  Om; 
0.54(265) 


265;    Vm  =  165 
0.31(265) 


Using  Eq.  (9-23), 

66.84 
i  "  33.16  I 


)*           232        1  03K 

1  ^               232 

0.07(265)' 
232 

'          0.54(165)    '  U'JU 

D;          0.31(165) 

232 
35    1   f      265 

232 

1  - 

4-  0  07(5") 

i 

0.07(165) 

1     1    0    C 

232  J 
0.075  ' 

,        d°    '    L  165(0.94) 

1   J  o.O 
-J                    1    eo 

15  i  r  265 

10  e 

'   L  165(0.94) 
1"6(0"0")       °-052l 

+  1.52) 
,         °'224         , 

*.-v,Vw.-w-/        1.52   1  (1 

1.26  -  1          J 
/66.84\  /      0.447 

006 


^33.16/  V1.26  -  0.66  ^  1.26  -  0.39 


6.9 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        261 


This  value  is  near  enough  to  the  original  assumption  and  no  additional  retrials 
will  be  made.  » 

Gasoline-stabilization  Example.  As  a  further  illustration  of  the  use 
of  these  methods,  a  gasoline-stabilization  operation  will  be  con- 
sidered. The  feed  composition  is  given  in  the  table  on  this  page  and  the 
tower  is  to  operate  at  250  p.s.i.g.  A  reflux  ratio  of  2  will  be  used  in  the 
upper  portion  of  the  tower,  arid  the  feed  will  enter  such  that  £0/F)m 
below  the  feed  will  be  1.5.  It  is  desired  to  recover  96  per  cent  of  the 
normal  butane  with  the  stabilized  gasoline,  but  this  bottoms  product 
is  to  contain  not  over  0.25  mol  per  cent  propane. 

In  preparing  this  table,  it  was  assumed  that  the  concentrations  of  all 
components  lighter  than  propane  were  negligible  in  the  residue  and 
that  all  components  heavier  than  n-C*  were  negligible  in  the  distillate. 
The  isobutane  is  intermediate  to  the  propane  and  n-butane  and  there- 
fore will  appear  in  appreciable  quantities  in  both  the  distillate  and 
residue.  Since  the  i-C4  is  more  volatile  than  n-C4,  the  following  table 
was  prepared  on  the  assumption  that  20  per  cent  of  the  i-Ci  in  the  feed 
would  appear  in  the  overhead.  The  volatilities  relative  to  n-C4  are 
given  in  Fig.  9-13.  These  relative  volatilities  are  based  on  the  fugacity 
data  of  Lewis  and  coworkers  (Ref.  6).  The  equilibrium  constant  K 


Feed 

Residue 

Distillate 

Mol, 
per  cent 

Mols/100  feed 

Mol, 
per  cent 

Mols/100  feed 

Mol, 
per  cent 

CH4 

2.0 



— 

2.0 

6  33 

C2H6 

10.0 

— 

— 

10  0 

31  60 

C8H6 

6.0 

— 

— 

6  0 

19  00 

C3H8 

12.5 

0  0025TF 

0.25 

12.5  -0  0025T7 

39  00 

i-C4Hio 

3  5 

2  8 

4  10 

0  7 

2  2 

n-CiELio 

15.0 

14.4 

21  10 

0.6 

1.9 

C6 

15.2 

15  2 

22.20 

C« 

11.3 

11  3 

16.50 

C7 

9.0 

9.0 

13.20 

C8 

8.5 

8  5 

12.40 

360°F. 

7.0 

7.0 

10.20 

68.2  +0.002517 

31.8  -0.0025TF 

68,2  4-  0.0025TF  -  W 
W  -  68.4;        D  -  31.6, 

for  n-C4  is  also  plotted  in  this  figure.     Since  the  overhead  is  very 
volatile,  it  will  be  removed  as  a  vapor,  only  enough  liquid  being  pro- 


262 


FRACTIONAL  DISTILLATION 


0.004 


0.002 


0,00? 


150 


200  250  300 

Temperature  ,deg.F. 

Fia.  9-13. 


350 


400 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        263 

duced  in  the  partial  condenser  to  furnish  reflux.  It  is  assumed  that  the 
reflux  from  the  condenser  leaves  in  equilibrium  with  the  product  vapor. 

Starting  at  the  composition  of  the  overhead  vapor,  the  calculations 
are  carried  down  the  column  by  the  use  of  the  equations  given  on  page 
232  and  the  results  are  summarized  in  Table  9-2.  The  first  coulmn  of 
this  table  gives  the  components,  the  second  column  lists  the  vapor  con- 
centrations for  the  plate  in  question,  and  the  third  column  gives  the  a 
values  at  the  assumed  temperature.  The  next  column  gives  the  values 
of  the  vapor  concentrations  divided  by  the  relative  volatility,  and  by 
using  Eq.  (9-3)  on  page  233,  the  liquid  concentrations  for  this  plate  are 
obtained  by  dividing  the  values  of  the  fourth  column  by  the  sum  of  all  of 
the  values  in  the  fourth  column.  On  the  basis  of  1  mol  of  overhead 
vapor  or  product,  there  are  2  mols  of  reflux,  and  for  this  reason  the  fifth 
column  listst  twice  the  concentrations  obtained  from  column  four  and  is 
therefore  the  actual  mols  of  overflow  for  the  basis  chosen.  There  will 
be  3  mols  of  vapor  to  the  plates,  and  the  mols  of  each  component'in  the 
vapor  to  any  plate  above  the  feed  plate  must  equal  the  sum  of  the  mols 
of  that  component  in  the  product  and  in  the  overflow  from  that  plate; 
i.e.,  the  sum  of  the  values  in  column  five  plus  the  values  in  y0  H..  These 
vapor  values  for  the  plate  below  are  listed  in  the  last  column  of  the 
table.  In  Table  9-3,  for  the  calculation  beginning  at  the  still,  a  similar 
procedure  was  used  employing  Eq.  (9-2)  on  page  233  and  using  a  basis 
of  1  mol  of  residue. 

A  temperature  of  100°F.  was  assumed  for  the  partial  condenser,  and 
the  calculated  temperatures  based  on  K n-c4  are  given  for  each  plate. 
At  the  second  plate  below  the  top  plate,  the  a  values  are  shifted  to 
150°F.  The  liquid  on  the  second  plate  below  the  top  plate  has  a  ratio 
of  CaHs/n-C^  a  little  higher  than  the  feed  ratio,  and  this  plate  will  be 
made  the  last  plate  above  the  feed;  i.e.,  the  feed  plate  will  be  the  fourth 
from  the.toj)  of  the  column.  If  an  attempt  is  made  to  carry  the  calcu- 
lations farther  down  the  tower,  a  serious  difficulty  will  be  met  in  that 
no  components  heavier  tha$  ^Q±  have  been  considered,  but  they  are 
much  too  large  to  be  neglected  below  the  feed  plate.  The  most  satis- 
factory solution  to  this  difficulty  is  to  carry  out  calculations  starting  at 
the  still  and  continuing  up  to  the  feed  plate.  These  calculations  are 
presented  in  Table  9-3.  Such  calculations  are  continued  until  the 
ratio  of  C3H8/n-C4  in  the  vapor  from  some  plate  is  approximately  the 
same  as  the  ratio  in  the  vapor  calculated  from  the  feed  plate  in  Table 
9-2.  Thus,  it  is  found  that  the  vapor  from  plate  8  of  Table  9-3  gives  a 
ratio  approximately  equal  to  the  ratio  on  the  T— 3  plate  of  Table  9-2. 


264 


FRACTIONAL  DISTILLATION 


TABLE  9-2 
(Basis:  1  mol  overhead  vapor;  0/D  =  2) 


2/O.H. 

<*100 

2/0  H 

2yo  H 

3yT 

o         _        a 

"XR       *2y/ct 

Ci 

0.0633 

36.5 

0  00173 

0.012 

0.075 

C2 

0.316 

7.4 

0  0427 

0.296 

0.612 

C8  — 

0.190 

3.0 

0  0633 

0.440 

0.630 

C8  + 

0  390 

2.7 

0  144 

1.000 

1  390 

i-C4 

0.022 

1.3 

0  0169 

0.117 

0.139 

rc-C4 

0.019 

1.0 

0  019 

0.132 

0.151 

0  2876 

Kn-c<  -  0.2876         T  =  98°F. 


3i/r 

«ioo 

3i/r/<* 

2xr 

32/r-i 

Oi 

0.075 

36  5 

0.0021 

0.004 

0  067 

C2 

0.612 

7  4 

0  0826 

0  155 

0  471 

C8~ 

0.630 

3.0 

0.210 

0.394 

0.584 

C8+ 

1.390 

2  7 

0  515 

0.965 

1  355 

*-C4 

0.139 

1  3 

0  107 

0.200 

0  222 

7l-C4 

0.151 

1.0 

0  515 

0  283 

0  302 

1.067 

1.067/3  =  0.356         T  =  120°F. 


SVT-I 

<*100 

3yT-i/a 

2XT-1 

3?/r-2 

Ci 

0  067 

36  5 

0  0018 

0  0029 

0.066 

C2 

0.471 

7.4 

0.0637 

0.103 

0.419 

C3- 

0.584 

3  0 

0  195 

0.316 

0.506 

c,+ 

1.355 

2.7 

0.501 

0.810 

1.200 

^-€4 

0.222 

1.3 

0  171 

0.277 

0.299 

w-C4 

0.302 

1.0 

0.302 

0.487 

0.508 

1.2345 

n-c4  -  0.41         T  -  130°F. 


3l/r-2 

«150 

32/r~2/« 

2zr-2 

32/r.-8 

Ci 

0.066 

26 

0  0025 

0.003 

0  066 

C2 

0.419 

6 

0.070 

0.091 

0  407 

C3- 

0.506 

2.6 

0.195 

0.253 

0.443 

C84- 

1  200 

2.3 

0.522 

0.678 

1  068 

i-C4 

0,299 

1.23 

0.243 

0.315 

p.  337 

n-C* 

0.508 

1.0 

0.508 

0.660 

0.679 

1.5405 

c<  -  0.51 


150°F. 


RECTIFICATION  OF  MULTIGOMPONENT  MIXTURES         265 

Approximately  eleven  theoretical  plates  in  addition  to  the  still  and 
partial  condenser  are  required. 

Feed-plate  Matching.  In  the  gasoline  stabilization  problem  just 
considered,  the  vapor  compositions  obtained  by  calculating  down  from 
the  condenser  do  not  appear  to  match  very  satisfactorily  with  those 
obtained  by  calculations  up  from  the  still.  This  is  due  to  the  fact  that, 
with  the  exception  of  propane,^  isobutane,  and  n-butane,  the  compo- 
nents involved  in  the  two  cases  are  different.  In  general,  this  condi- 
tion is  always  encountered  when  there  are  components  both  more  vola- 
tile than  the  light  component  and  less  volatile  than  the  heavy  compo- 
nent. In  order  to  make  the  compositions  match  more  exactly,  it  is 
necessary  to  introduce  the  light  components  into  the  calculations 
below  the  plate  and  the  heavy  components  into  the  calculations  above 
the  feed  plate.  It  is  not  necessary  to  repeat  the  calculations  all  the 
way  from  the  still  with  all  the  light  components.  It  is  only  necessary 
to  drop  back  a  sufficient  number  of  plates  such  that  the  concentration 
of  the  component  or  components  to  be  added  will  be  small  and  can  be 
added  without  altering  the  accuracy  of  the  material  balance. 

Thus,  in  Table  9-4  the  results  of  Table  9-3  are  dropped  back  to  the 
seventh  plate  and  small  concentrations  of  C2  and  C3  are  added;  and 
then  on  the  eighth  plate  the  Ci  is  introduced.  It  is  obvious  that  the 
concentration  of  these  light  components  should  be  added  such  that  the 
vapor  from  the  eighth  plate  will  match  the  composition  of  these  com- 
ponents as  determined  by  the  calculations  for  the  top  section. 

This  matching  is  complicated  by  the  fact  that  the  values  of  light 
components  obtained  for  the  top  section  of  the  tower  in  the  first  calcu- 
lations will  be  reduced  somewhat  by  the  introduction  of  the  heavy 
components  into  this  section,  and  it  is  therefore  a  matter  of  successive 
approximations  to  obtain  an  exact  match.  The  quantity  of  the 
components  to  be  added  on  a  given  plate  to  obtain  a  desired  value 
requires  trial  and  error  but  can  be  simplified  by  the  fact  that  for 
a  light  component  below  the  feed  plate  the  operating  line  is  essentially 
VnJUm  =  (WiSm+i,  since  the  value  of  Wxw  is  negligible.  This  can  be 
combined  with  the  equilibrium  constant  to  give  own  =  Km(V/O)mXm. 
Thus  the  increase  in  concentration  per  plate  for  a  light  component  in 
the  lower  portion  of  the  tower  is  essentially  equal  to  the  equilibrium 
constant  times  the  7/0  ratio.  Instead  of  K  the  value  of  a/(2ax)  can 
be  used.  These  relations  make  it  relatively  easy  to  estimate  the 
number  of  plates  that  should  be  recalculated  to  obtain  the  desired 
values.  Above  the  feed  plate,  a  similar  relationship  can  be  developed 
for  the  heavy  components,  and  the  decrease  in  concentration  of  such  a 


^  ^  „•  ____  T 


266 


FRACTIONAL  DISTILLATION 


TABLE  9-3 
Basis:  1  mol  residue;  (0/F)n 


1.5 


Xw 

<*300 

aXw 

2yw  *»  2<xxw/2<xx 

3zi 

C.+ 

0  0025 

2.0 

0  005 

0  020 

0.0225 

z-C4 

0  041 

1  18 

0  048 

0  191 

0.232 

n-C4 

0  211 

1  0 

0  211 

0.840 

1.051 

C* 

0  222 

0  58 

0.129 

0.513 

0  735 

C6 

0.165 

0.38 

0  0627 

0.249 

0.414 

C7 

0  132 

0  215 

0  0284 

0  115 

0  247 

C8 

0  125 

0.12 

0.0150 

0  060 

0.185 

360°F. 

0.102 

0.038 

0.0039 

0.016 

0  118 

0.503 

«  1/0.503  »  1.99 


333°F. 


3zi 

«300 

Saxi 

2yi 

3z2 

C8  + 

0  0225 

2  0 

0  045 

0  044 

0.0465 

*-C4 

0.232 

1.18 

0  273 

0.269 

0  310 

w-C4 

1  051 

1.0 

1  051 

1.035 

1  246 

C6 

0  735 

0.58 

0  426 

0.420 

0  642 

c. 

0  414 

0  38 

0  157 

0  155 

0.320 

C7 

0.247 

0  215 

0  053 

0  052 

0  184 

C8 

0.185 

0  12 

0  022 

0  022 

0.147 

360°F. 

0  118 

0.038 

0.005 

0  005 

0  107 

2  032 

3/2.032  =  1.48 


285°F. 


3*2 

&300 

3o#2 

22/2 

3#3 

C8  + 

0.0465 

2.0 

0.093 

0  082 

0  0845 

t-C4 

0.310 

1  18 

0  365 

0  323 

0  364 

n-C4 

1.246 

1  0 

1  246 

1  103 

1  314 

C6 

0.642 

0  58 

0.372 

0.329 

0  551 

C6 

0.320 

0  38 

0.121 

0  107 

0  272 

C7 

0.184 

0  215 

0.040 

0.035 

0  167 

C8 

0.147 

0  12 

0  018 

0  016 

0.141 

360°F. 

0.107 

0.038 

0.004 

0.004 

0.106 

2.259 

#n.C4  -  3/2.259  »  1.33        T  -  273°F. 


RECTIFICATION  OF  MVLTICOMPONENT  MIXTURES        267 
TABLE  9-3  (Continued) 


3#8 

a  2  50 

3ax3 

2t/3 

3£4 

c,+ 

0.0845 

2.03 

0  172 

0  147 

0  1495 

*'-C4 

0.364 

1  18 

0  430 

0.367 

0.408 

n-C4 

1  314 

1.0 

1.314 

1.120 

1.331 

C6 

0.551 

0  55  , 

0.302 

0  258 

0.480 

cfl 

0.272 

0.30 

0.082 

0.070 

0.235 

C7 

0  167 

0  172 

0  029 

0  025 

0.157 

C8 

0.141 

0  090 

0  013 

0  Oil 

0  136 

360°F. 

0.106 

0.022 

0  002 

0  002 

0  104 

2  344 

1.28    T  -  265°F. 


3z4 

«260 

3a£4 

2?/4 

3x6 

C.+ 

0  1495 

2  03 

0  304 

0  244 

0  2465 

i'-C4 

0.408 

1  18 

0  481 

0  386 

0.427 

7^C4 

1.331 

1.0 

1  331 

1.070 

1.281 

C6 

0  480 

0  55 

0  264 

0.212 

0  434 

C6 

0  235 

0.30 

0  071 

0  057 

0  222 

C7 

0  157 

0  172 

0.027 

0  022 

0  154 

C8 

0  136 

0  090 

0  012 

0  010 

0  135 

360°F. 

0  104 

0.022 

0.002 

0  002 

0.104 

2.492 

1.2 


T  -  256°F. 


/ 

3zB 

<*260 

3aZ8 

2|/6 

3z6 

C3  + 

0.2465 

2  03 

0  500 

0  380 

0  382 

i-C4 

0  427 

1.18 

0.504 

0.383 

0  424 

n-C4 

1.281 

1.0 

1.281 

0.975 

1  186 

C6 

0.434 

0.55 

0.238 

0  181 

0  403 

C6 

0.222 

0.30 

0  067 

0  051 

0  216 

C7 

0.154 

0  172 

0.027 

0  021 

0.153 

C8 

0.135 

0.090 

0  012 

0  009 

0  134 

360°F. 

0  104 

0.022 

0  002 

0  002 

0.104 

2.631 

1.14   T  -  250°F. 


268 


FRACTIONAL  DISTILLATION 
TABLE  9-3  (Continued) 


3*« 

<*250 

3«z6 

2ye 

3x7 

c,+ 

0.382 

2.03 

0  776 

0.556 

0.558 

i-d 

0  424 

1  18 

0  501 

0.359 

0  400 

n-Ct 

1  186 

1  0 

1.186 

0.850 

1  061 

C6 

0  403 

0  55 

0  222 

0  159 

0  381 

C6 

0.216 

0.30 

0.065 

0.047 

0  212 

C7 

0.153 

0.172 

0.026 

0.019 

0.151 

Cs 

0  134 

0.090 

0  012 

0.009 

0  134 

360°F. 

0.104 

0  022 

0  002 

0.001 

0  103 

2.790 

d  -  1-075 


241  °F. 


3x7 

«260 

3a#7 

2?/7 

3x8 

C3  + 

0  558 

2  03 

1  135 

0  762 

0  764 

f-C4 

0.400 

1.18 

0  472 

0  317 

0  358 

n-C4 

1  061 

1  0 

1  061 

0  713 

0  924 

C6 

0  381 

0  55 

0  210 

0  141 

0  363 

C6 

0  212 

0  30 

0  064 

0.043 

0  208 

C7 

0  151 

0  172 

0  026 

0  018 

0  150 

C8 

0.134 

0  090 

0  012 

0  008 

0  133 

360°F. 

0  103 

0  022 

0.002 

0  001 

0  103 

2  982 

t-CU 


1.01 


232°F. 


3x8 

«260 

3aZ8 

2j/8 

C3  + 

0.764 

2  03 

1.550 

0  970 

i-C« 

0  358 

1  18 

0  422 

0  264 

n-C4 

0  924 

1  0 

0.924 

0  578 

C6 

0.363 

0  55 

0.200 

0  125 

Ce 

0.208 

0.30 

0  062 

0  039 

C7 

0.150 

0  172 

0  026 

0.016 

C8 

0  133 

0  090 

0  012 

0  008 

360°F. 

0.103 

0.022 

0  002 

0  001 

3  198 

-  0.94        T  =  220°F, 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES         269 
TABLE  9-4.     REMATCHING  FEED  PLATE  FROM  BELOW 


3z7 

«225 

3cu7 

2t/7 

3#8 

Ci 

— 

19  3 





0  0045 

C2 

0.036 

5.1 

0.181 

0  100 

0  100 

C3- 

0.183 

2.37 

0.432 

0  240 

0  240 

C3  + 

0  559 

2  07  , 

1  155 

0  642 

0  644 

i-C* 

0  400 

1  18 

0  472 

0  262 

0  303 

n-C4 

1  061 

1  0 

1  061 

0.590 

0.801 

C6 

0  381 

0  53 

0  202 

0.112 

0  334 

C6 

0  212 

0.27 

0.057 

0  032 

0  197 

C7 

0  151 

0  15 

0.023 

0  013 

0.145 

C8 

0  134 

0  072 

0  010 

0  006 

0  131 

360°F. 

0  103 

0  016 

0  002 

0  001 

0  103 

3  595 

c4  -  O.S33         T  =  205°F. 


3z8 

«200 

Sotfs 

2/8 

Ci 

0  0045 

21 

0  094 

0  024 

C2 

0  100 

5.3 

0  530 

0  134 

C3- 

0  240 

2.4 

0.576 

0  146 

a+ 

0  644 

2.1 

1.350 

0  341 

i-C* 

0  303 

1.2 

0  364 

0  092 

n-C4 

0  801 

1.0 

0  801 

0  203 

C6 

0.334 

0.5 

0  167 

0  042 

C6 

0.197 

0  24 

0  047 

0  012 

C7 

0.145 

0  125 

0  018 

0  004 

C8 

0  131 

0  057 

0  007 

0  002 

360°F. 

0.103 

0  012 

0  001 

0  0003 

3  955 

Kn 


0.76 


188°F. 


The  rematching  calculations  for  the  gasoline-stabilizer  problem  for 
above  the  feed  plate  are  given  in  Table  9-5. 

By  such  rematching  procedure,  it  is  possible  to  obtain  exact  agree- 
ment at  the  feed  plate  for  all  the  light  components  and  all  the  heavy 
components  because  they  are  arbitrarily  chosen  in  one  portion  of  the 
tower.  However,  it  may  be  impossible  to  obtain  an  exact  match  of 
the  key  components  because  an  even  number  of  theoretical  plates  will 
not  be  consistent  with  the  design  chosen.  In  this  case,  it  is  possible 
to  bracket  the  required  number  of  plates  within  a  difference  of  one, 
plate.  While  this  rematching  operation  gives  a  more  consistent  set  of 


270 


FRACTIONAL  DISTILLATION 


compositions,  in  most  cases  it  does  not  alter  the  conclusion  as  to  the 
number  of  theoretical  plates  required,  and  the  first  calculations  of  the 
type  illustrated  by  Tables  9-2  and  9-3  are  usually  sufficient  for  deter- 
mining the  number  of  theoretical  plates. 

After  such  adjustments,  it  is  noted  that  the  vapor  ys  of  Table  9-4 
and  the  vapor  i/r-3  of  Table  9-5  give  a  very  satisfactory  match.  The 
i-C4  from  Table  9-5  is  a  little  higher  than  in  Table  9-4,  indicating  that 
a  little  less  than  20  per  cent  of  the  i-d  would  go  overhead,  but  the  dif- 
ference is  so  small  that  it  does  not  justify  readjusting. 

These  concentrations  given  in  the  tables  are  plotted  in  Fig.  9-14. 

TABLE  9-5.     REMATCHING  FEED  PLATE  FROM  ABOVE 


2x>r-i 

3?/r_2 

AUO 

3?/r_2/a 

2xr~2 

2/T-3 

Ci 

0  0029 

0.066 

26 

0.0025 

0.003 

0  022 

C2 

0.103 

0.419 

6 

0  070 

0.084 

0.133 

CB- 

0  316 

0  506 

2.6 

0  195 

0.233 

0  141 

C8  + 

0.810 

1.200 

2.3 

0  522 

0  624 

0  338 

t-C4 

0  277 

0  299 

1  23 

0  243 

0  290 

0  104 

n-C4 

0  489 

0  508 

1  0 

0  508 

0.606 

0.208 

C6 

0.045 

0  045 

0  43 

0  105 

0.125 

0  042 

Co 

0.005 

0  005 

0.18 

0.028 

0.033 

0.011 

C7 

— 

— 

— 

— 

0  012 

0  004 

c* 

— 

— 

— 

— 

0.006 

0  002 

360°F. 

— 

— 

— 

— 

0  0009 

0  0003 

1  674 

Kn^<  -  0.56        T  -  163°F. 

A  heat  balance  around  the  feed  plate  indicates  that  the  feed  should 
enter  as  a  liquid  at  about  130°F.  to  give  the  vapor  and  liquid  flows 
assumed. 

Optimum  Feed-plate  Location.  The  feed  plate  was  chosen  so  that  the 
ratio  of  the  key  components  was  approximately  the  same  as  in  the 
feed.  Equation  (9-12)  indicates  the  optimum  feed-plate  ratio  for 
Cs  +  /n-Ct  should  be  1.14  as  compared  to  0.83  for  the  ratio  in  the  feed. 
After  rematching,  the  ratio  of  the  key  components  for  plate  8  in  Table 
9-3  is  0.803,  and  for  plate  9  (plate  T-2)  in  Table  9-4  the  ratio  is  1.03. 
The  ratio  for  these  two  plates  should  bracket  the  value  of  1.14.  These 
values  indicate  that  adding  the  feed  to  plate  9  would  give  more  effec- 
tive rectification  than  the  plate  that  was  employed. 

Theoretical  Plates  at  Total  Reflux.  The  relative  volatility  of  the  key 
components  does  not  vary  too  widely  from  the  still  to  the  condenser, 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        271 


272 


FRACTIONAL  DISTILLATION 


and  a  satisfactory  answer  for  the  number  of  theoretical  plates  at  total 
reflux  can  be  obtained  by  the  use  of  Eq.  (7-54).  Using  an  arithmetic 
average  of  the  relative  volatility  at  the  still  and  condenser, 

2.0  +  2.7      OQP, 

OJav   =  o =   2.35 


ln 


N  +  2  = 


0.211 


=  8.7 


In  2.35 
N  =  6.7  (or  7)  theoretical  plates 

Minimum  Reflux  Ratio.     The  minimum  reflux  ratio  will  be  calcu- 
lated for  this  separation  by  Eq.  (9-22). 
Basis:  100  mols  of  feed.    For  actual  case, 


v'  -1-6 

Om  =d.5Fm  =  Vm  +  W 
Vm  =  136.8 

-  =  2 
Fn  =  3D  =  94.8 


Therefore,  assume  Vm  —  Vn  =  42  for  minimum  reflux  conditions. 
Values  of  the  relative  volatility  corresponding  to  the  feed  plate  will 
be  used. 


a 

a 

Ci 

24  0 

c* 

0  46 

C2 

5  6 

C6 

0  21 

C8- 

2.5 

C7 

0.11 

C8+ 

2.2 

C8 

0.05 

*-C4 

1.21 

360°F. 

0.01 

n^C4 

1  0 

In  calculating  <t>  by  Eq.  (9-12),  it  is  necessary  to  have  values  of 
and  Vm.    As  a  first  trial,  assume  (0/D)miB  =  0.9. 


Vn  =  1.9(31.6)  =  60 
Vm  =  102 


On  =  28.4 
Om  =  170.4 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES 
in.  fQ-10Y 


By  Eq.  (9-10), 


273 


:  60 


2.0 


170.4 


-      170-4 
102(15.8) 

_      28.4 
60(15.8) 


+  10 


1  170.4 

102(3.68) 
_      28.4 
60(3.68) 


6.0 


-      170.4 
102(1.66) 

9ft  /t 


__      28.4 
60(1.66) 

K  values  at  a  temperature  of  175°F.  were  employed. 
ByEq.  (9-11), 

r  i  -  0-30(60) 

I30(15-2>  ,       0.308(1402) 
170.4 


=  0.148 


1 

I7O4 


.  0-14(60) 

OO    A 


*         U.JL^^WJ 

•*•  00  A  x 

+  0-14(11.3)  — 0-^  +  0.07(9.0)- 
170.4  l 


_  0.07(60) 
28.4 


J-    I 

1      0-03(60) 

-  °-°3(8-5)  ,       0.023(102l 
170.4 

T?^.       fC\    10\ 


1 


170.4 
0.007(60) 
28.4 


i 


a».4 
0.007(102) 

17O  A 


170.4 


By  Eq.  (9-12)3 

160(0.852)'^  I 

A  =  1102(0.976)       1J  _  ^  j  g 

iKn^r60(0-8S2)2        .T... 

15'0  +  [  102(0.976)  -T4-4 

Checking  distribution  of  t-C4  for  assumed  reflux  ratio. 
By  Eq.  (9-13), 

1  O^* 
.       _  L°28l 


=  0.024 


1.2 
By  Eq.  (9-15), 


OJ04Dxi> 


2.2  -  1.21  +  1, 


o2  ^IF 

2-2170 

91  /  68.4  \  /O.OQ25\ 


274  FRACTIONAL  DISTILLATION 

Equating  these  two  values  of  art.c4  and  using  Wxw  +  DxD  =  3.5, 

DxD  =  0.39 
Wxw  «  3.11 

About  11  per  cent  of  the  isobutane  would  go  overhead.  The  calcu- 
lations of  D  and  W  were  based  on  the  original  assumption  of  20  per 
cent  i-C4  overhead,  but  the  difference  is  not  large  and  no  correction 
will  be  made.  The  values  just  calculated  will  be  used  in  the  minimum 
reflux  ratio  calculations. 

By  Eq.  (9-22) 


w. 


-  0.019  j[l  +2.2(1.8)] 
+  1 


/0.39V 

L21  V3L6; 


24.0(0.063)       5.6(0.316)       2.5(0.19)       

24.0  -  1    "^    5.6-1    "*"  2.5  -  1   "*"    1.21  -  1 

60(68.4)  [0.46(0.222)       0.21(0.165)       0.11(0.132) 

102(31.6)  [2.2  -  0.46  "*"  2.2  -  0.21   +  2.2  -  0.11 

,  0.05(0.124)    ,   0.01(0.102) 


n. 


1   2.2  -  0.05    '   2.2  -  0.01 
~)      =  0.75  for  Case  I 

=  0.75  -  0.11  =  0.64  for  Case  II 

The  assumed  value  for  0/D  of  0.9  is  close  to  the  calculated  value  for 
Case  I.  A  recalculation  for  Case  II,  assuming  0/D  =  0.7,  gave  a 
calculated  value  of  0.63. 

Equation  (9-22)  was  employed  because  the  summation  term  for  the 
light  components  was  greater  than  for  heavy  components.  The  term 
for  the  isobutane  was  included  with  the  light  components,  but  it  could 
equally  well  have  been  included  with  the  heavy  components,  and  the 
total  would  have  been  essentially  the  same. 

It  is  interesting  to  note  that  a  binary  mixture  of  propane  and  butane 
of  the  same  ratio  as  in  the  feed  would  have  required  only  one-half  as 
much  vapor  per  mol  of  Cs  +  C*  separated  as  is  necessary  for  this 
multicomponent  mixture. 

For  a  comparison,  the  minimum  reflux  ratio  will  also  be  calculated 
by  Eq.  (947). 

By  Eq.  (9-17),  assume 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        275 

By  Eq.  (9-15), 

2.2(14.4/170.4)  A1KK 


By  Eq.  (9-13), 

Xlk  =  -  1.0(12.33/28.4)  _     _ 

"-"+*&)  Css) 

For  Eq.  (9-17), 

C  =  ~      (2'2  ~  1'°  +  °-30)  =  3'84 


E  =  gj  (2.2  -  1.0)  =  5.68 
O  =  M  =  0.387 

=  0-26 

°'165 


/  =  +  0.387  +  2.2  + 


68.4     '   v-""§    '      '    [2.2  -  0.46    '   2.2  -  0.21 

0.132 0.124  0.102     1 

"*"  2.2  -  0.11  +  2.2  -  0.05  "*"  2.2  -  0.01  J 

31.6/170.4\r  0.063 

^        68.4  V  28.4  /  [24  -  1  +  0.12(24) 
, 0.316  0.19 

I     K.  a  i      i     f\  -i  ci  t e  /2\      i 


5.6  -  1  +  0.12(5.6)    '   2.5  -  1  +  0.12(2.5) 


+  1 


0.39/31.6  1 

.21  -  1  +0.12(1.21)J 


=  -2.08  +  0.387  4-  0.825  +  0.617  =  -0.251 


J  =  0.26  +  1^1  (^^  )  (0.825)  +  0.205  =  0.762 

If  =  -0.868 
J'  -  0.465 

Case  I: 

W  68  4 

IL  /  +  j  =  5^Z  (-0.251)  +  0.762 

«  0.218 


0.218  +  ^/(0.218)2  -  ^fP  [(~0.251)(0.762)  -  (0.387) (0.26)1 
=  0.91 


276  FRACTIONAL  DISTILLATION 

Case  II: 


-1.415 


1    4.1  K   _L_        //'_ 

1    J.1  KN2    _ 

4(68.4) 

r    o  sfiR/'n  ifi^ 

/"n  ^ft7^^^  9  AM 

1.4J.O   -f  'V/V 

31.6    ' 

[        U.oOOViU.4:OO^ 

(\}.oot  )  \\},Zi\y)\ 

0.55 

2 

These  values  were  not  recalculated  for  a  new  assumed  value  of  0/D 
since  both  are  reasonably  close  to  the  first  value  assumed. 

Constancy  of  Molal  Flow  Rates.  The  design  calculations  pre- 
sented have  been  based  on  the  use  of  constant  molal  overflow  and 
vapor  rates.  Enthalpy  balances  similar  to  Eqs.  (7-13)  and  (7-27)  can 
be  written  for  each  of  the  components  of  a  multicomponent  mixture 
and,  if  the  data  are  available,  they  can  be  applied  plate  by  plate  in  the 
Stepwise  calculations.  This  procedure  usually  requires  trial  and  error 
because,  in  calculating  up  or  down  the  tower,  the  temperature  on  the 
next  plate  is  needed  to  complete  the  enthalpy  balance,  and  it  must  be 
assumed  and  checked  in  a  later  calculation.  The  most  serious  diffi- 
culty is  the  lack  of  the  necessary  enthalpy  data,  but  in  most  cases 
these  can  be  approximated  by  the  methods  on  pages  139  to  142. 

The  same  general  considerations  relative  to  the  constancy  of  flow 
rates  apply  to  multicomponent  and  binary  mixtures.  Thus  the  latent 
heat  of  vaporization  at  various  positions  in  the  column  should  serve 
as  the  principal  criterion,  although  in  multicomponent  systems  there  is 
more  possibility  of  large  sensible  heat  effects  changing  the  overflow  and 
vapor  rates. 

The  modified  latent  heat  method  (M.L.H.V.)  given  on  page  158  is 
applicable  to  multicomponent  mixtures,  and  it  probably  is  the  most 
satisfactory  procedure  for  handling  such  calculations.  It  should  give 
good  results  for  the  examples  considered  in  this  chapter,  because  the 
heats  of  mixing  for  the  mixtures  involved  would  be  small,  even  though 
the  latent  heats  of  vaporization  of  individual  components  for  the 
gasoline  stabilization  problem  differ  several  fold.  In  this  latter  exam- 
ple, there  is  undoubtedly  considerable  variation  in  the  molal  flow 
rates,  and  the  M.L.H.V.  method  will  give  more  satisfactory  results 
than  the  method  employed.  The  calculations  were  repeated  by  the 
M,L,H.V,  method  using  arithmetic  average  values  of  latent  heats 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES         277 

of  vaporization  at  the  still  and  the  condenser  to  calculate  the  val- 
ues of  ft  for  each  component.  Table  9-6  gives  the  latent  heat  values 
and  the  ft  terms  based  on  the  latent  heat  of  n-butane.  The  values 
of  ft  are  used  to  calculate  ftzF,  ftyo.n.,  ftxw,  and  z'F  =  ftzF/2(ftzF), 
2/o.H.  **  02/o.H./2(/ty0.H.),  and  x'w  =  ftxw/2(ftxw).  The  values  of  the 
terminal  flow  quantities  are  given  below: 

W'  =  68.4(1.467)  =  100 
D'  =  31.6(0.650)  =  20.5 
F'  =  100(1.205)  =  120.5 

Table  9-7  presents  the  plate-to-plate  calculations;  the  values  of 
2/o  H  are  taken  from  Table  9-6,  and  the  values  of  t/o.n  and  2xR  are  taken 
from  Table  9-2.  The  calculations  will  be  made  for  0/D  =  2.0  and 
for  Vm  —  Vn  =  42  which  are  the  same  as  the  values  used  in  Tables 
9-2  and  9-3.  For  F  =  100,  D  =  31.6,  0K  =  63.2,  and  the  fourth  col- 
umn of  Table  9-7  gives  ftxR  =  0.751  making  0'R  =  0.751(63.2)  =  47.5, 
and  O'JD'  =  47.5/20.5  =  2.32.  The  remaining  columns  of  Table 
9-7  are  based  on  D'  »  1.0,  0'  =  2.32,  V  =  3.32.  The  values  of  the 

fifth  column  are  2.82^  =  2.32     ,*R    •  the  values  of  332y'T  are  equal 

•^  (pxx) 

to  2.32x'R  +  2/o.H.-  In  order  to  go  to  yT,  the  values  of  3.32y'T  are 
divided  by  the  corresponding  ft  terms  and  yT  =  (y'T/ft)/2(y'T/ft).  The 
yT  values  are  used  to  calculate  XT,  using  the  relative  volatilities  of 
Table  9-2,  and  these  are  converted  to  x'T. 

The  general  flow  quantities  are  shown  in  Fig.  9-15  which  gives  both 
the  normal  and  modified  values.  For  F  =  100,  the  value  of  V'n  is  68 
throughout  the  top  section,  but  the  value  of  Vn  will  vary  from  plate  to 
3late.  Thus  VT  =  94.8,  but  7r-i  =  V^^yr-i/ft)  =  91,  and  this 
/alue  decreases  further  for  the  next  plate.  This  decrease  in  vapor 
'ate  is  due  to  the  fact  that  the  average  latent  heat  of  vaporization  is 
ncreasing  from  plate  to  plate  down  the  column,  resulting  in  lower 
fapor  and  liquid  rates  and  making  the  separation  more  difficult. 

The  calculations  are  carried  in  the  same  manner  down  to  yT~x-  The 
calculations  were  given  in  detail  to  show  the  values  of  y  and  x  as 
veil  as  those  for  y*  and  re',  but  this  is  not  necessary  because  it  is  possi- 
)le  to  go  from  y'  to  x'  directly  by  xf  »  (yf / a)  /  I,(y' / a)  in  which  the 
lormal  a  values  are  employed. 

The  calculations  for  the  lower  section  of  the  tower  are  given  in 
Table  9-8  on  the  basis  W  «  1.0,  O'm  =  2.186,  and  V'm  =  1.186.  The 
'alue  of  Vw  for  F  -  100  is  calculated  from  VjV(fow)  -  101.  This 


278 


FRACTIONAL  DISTILLATION 


represents  a  large  decrease  in  vapor  from  the  top  of  the  tower  where 
VT  •»  94.8.  With  Vm  -  Vn  «  42,  the  vapor  from  the  still  would  have 
been  136.8  on  the  usual  basis.  The  decrease  of  136.8  -  101  =  35.8  is 
due  to  variation  in  the  latent  heats  of  vaporization.  The  calculations 
up  to  x(  are  given  in  Table  9-8.  The  calculations  up  to  x*  were  made 
in  detail,  but  from  x%  to  x'6  the  values  of  l.lSQy'  were  calculated  from 


f'100 


(-  / 
Om 


W'68.4 


W'*tOO 
FIG.  9-15.     Comparison  of  flow  quantities  for  gasoline-stabilization  example. 

1.186(2.186az')/:S(2.186az').  This  method  requires  only  three  col- 
umns per  plate,  and  the  calculations  are  as  simple  as  the  normal  type 
of  calculations  given  in  Table  9-3.  The  calculations  were  continued, 
and  yj  gave  a  good  match  with  y'T_$  of  Table  9-7,  indicating  that  12 
theoretical  plates  would  be  required  as  compared  to  11  plates  for  con- 
stant flow  rate  conditions.  The  difference  would  be  greater  for  reflux 
ratios  nearer  the  minimum. 

The  minimum  reflux  ratio  can  be  calculated  on  the  modified  latent 
heat  basis  using  any  of  the  equations  derived  for  constant  molal  flow 
rates  with  the  x*  and  y'  values. 

The  calculations  for  (OfR/Df)min  by  Eq.  9-22  were  made  in  an  analo- 
gous manner  to  calculations  on  page  274.  The  values  are  summarized 


RECTIFICATION  OF  MULT1COMPONENT  MIXTURES        279 

below: 

A{  =  0.069;        Ai  =  0.026;        <#>'  =  1.055 
For  isobutane, 

D'x'D  =  0.3 
W'x'w  =  2.96 
By  Eq.  (9-22), 


=  1.0,  Case  I 
=  0.93,  Case  II 

TABLE  9-6 


Latent  heat  of  vaporization, 

B.t.u./lb.  mol 

f 

ft 

ZF 

ftZF 

ZF 

98°F. 

333°F. 

A.-V 

Ci 

1,100 

200 

650 

0.113 

0.02 

0.002 

0.0016 

C2 

3,500 

2,200 

2,850 

0.496 

0.10 

0.0494 

0  0413 

Ca  — 

4,700 

3,500 

4,100 

0  713 

0.06 

0  0425 

0  0354 

Cs-f- 

5,200 

4,000 

4,600 

0  80 

0.125 

0.10 

0  0826 

i-C4 

5,800 

4,900 

5,350 

0.93 

0  035 

0.0321 

0  027 

n-C4 

6,200 

5,300 

5,750 

1.0 

0.150 

0.150 

0  124 

C6 

7,300 

6,600 

6,950 

1.21 

0.152 

0.182 

0  152 

C6 

8,500 

7,700 

8,100 

1.41 

0.113 

0.159 

0.131 

C7 

10,000 

9,100 

9,550 

1.66 

0.090 

0.150 

0.124 

C8 

11,900 

10,700 

11,300 

1.97 

0  085 

0.166 

0.139 

360°F. 

14,900 

13,500 

14,200 

2.47 

0.070 

0  172 

0  143 

1 

1.205 

, 

/ 

XD 

fix* 

XD 

*       x* 

fixw 

xw 

Ci 

0.0633 

0.007 

0.011 

_*_ 

_,  

— 

C2 

0  3160 

0.157 

0.242 

— 

— 

— 

c,- 

0.1900 

0.135 

0  208 

— 

— 

— 

v/3  ~| 

0.3900 

0.312 

0.480 

0.0025 

0.002 

0  0014 

$~O4 

0  022 

0.020 

0.031 

0.0410 

0.038 

0.026 

n-C* 

0.019 

0.019 

0,029 

0.2110 

0.211 

0  144 

C* 

— 

— 

— 

0.2220 

0.268 

0.183 

C* 

— 

— 

— 

0.1650 

0.233 

0.152 

C7 

— 

— 

— 

0.1320 

0.219 

0  149 

C8 

— 

— 

— 

0.1240 

0.244 

0  166 

360°F. 

— 

— 

_ 

0.1020 

0.252 

0.172 

0.650 

1.467 

280 


FRACTIONAL  DISTILLATION 
TABLE  9-7 


2/O.H. 

J/O.H. 

2xR 

20afc 

2.324 

3.322/y 

3.32^ 

2/r 

ft 

Ci 

0  Oil 

0  0633 

0  012 

0.0014 

0.0022 

0  0132 

0  117 

0  025 

C2 

0.242 

0.316 

0.296 

0  147 

0.227 

0.469 

0.945 

0  204 

C3- 

0.208 

0.190 

0.440 

0.314 

0  484 

0.692 

0  970 

0.210 

tJ,+ 

0.480 

0  390 

1.0 

0  80 

1.232 

1.712 

2  14 

0  463 

i-C4 

0.031 

0  022 

0.117 

0.109 

0.168 

0.199 

0  214 

'0  046 

n-C4 

0.029 

0.019 

0.132 

0  132 

0  2035 

0  2325 

0  2325 

0  050 

1  503 

4.618 

VT  -  94.8 


yr 

XT 

ftXT 

2.324 

3.322/;_1 

3.32^ 

2/r-i 

2/r-i 

OflQQ 

ft 

«150 

Ci 

0.00068 

0  00191 

0  00022 

0  0006 

0  0116 

0  102 

0  023 

0  0008S 

C2 

0.0276 

0  0777 

0  0385 

0  112 

0  354 

0.714 

0  161 

0  027 

C8- 

0.070 

0  197 

0.141 

0  409 

0  617 

0.866 

0  195 

0  075 

C3  + 

0  172 

0  484 

0  387 

1  12 

1.60 

2  0 

0  45 

0  196 

i-C4 

0  035 

0.0985 

0.0915 

0  266 

0.297 

0.319 

0  072 

0  059 

n-C4 

0  050 

0.141 

0  141 

0  409 

0  438 

0.438 

0.099 

0  099 

0  3553 

0  799 

4  439 

0  4569 

VT-I  -  91 


XT-I 

ftXT-l 

2.324-! 

3.32^_2 

3.32^_2 

J/r-a 

2/r-2 

ft 

«160 

Ci 

0  00195 

0  00022 

0  00062 

0  0116 

0  102 

•  0  023 

0  00088 

C2 

0  059 

0.0293 

0.082 

0.324 

0  653 

0  150 

0  025 

C3- 

0  164 

0.117 

0  328 

0  536 

0  753 

0  173 

0  0665 

C3  + 

0  430 

0.344 

0  965 

1  445 

1  81 

0  416 

0  181 

*-C4 

0  129 

0  120 

0.336 

0  367 

0  395 

0  091 

0  074 

w-C4 

0.217 

0  217 

0  608 

0  637 

0  637 

0  147 

0  147 

0  8275 

4.350 

0  .  4944 

89 


XT-Z 

PXT~2 

2.324-a 

3.32^-8 

Ci 

0  0018 

0.0002 

0  00054 

0  0115 

C2 

0  051 

0.025 

0.068 

0.31 

C3- 

0.135 

0.096 

0  260 

0.468 

C8  + 

0  366 

0.296 

0  803 

1.283 

i~C4 

0  150 

0  140 

0  380 

0  411 

w-C4 

0  298 

0  298 

0  808 

0  837 

0  855 

RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        281 


TABLE  9-8 


*; 

xw 

1.1862/17 

l.lMpyw 

Vw 

1.186^ 

2.186^ 

C3-h 

0.0014 

0.0025 

0  0118 

0  0095 

0  0068 

0  0081 

0.0095 

;-c4 

0.026 

0  041 

0.113 

0  105 

0.0755 

0.090 

0.116 

rc-C4 

0.144 

0.211 

0.497 

0.497 

0.357 

0.424 

0.568 

C6 

0.183 

0.222 

0.304^ 

0  368 

0.265 

0.315 

0  498 

C6 

0  152 

0  165 

0  148 

0  209 

0  150 

0.178 

0  330 

Cr 

0.149 

0.132 

0  067 

0.111 

0  080 

0  095 

0.244 

C8 

0  166 

0.125 

0.035 

0  069 

0  050 

0.059 

0  225 

360°F. 

0  172 

0.102 

0  009 

0  022 

0.016 

0  019 

0.191 

1.390 

Vw  =  101 


2.186*; 

Xi 

«30C#1 

2/i 

ftyi 

1.186^ 

2.186a4 

ft 

Ca-h 

0.0119 

0  0071 

0  0142 

0.0215 

0  0172 

0  0186 

0.02 

<-C4 

0  1245 

0.074 

0.087 

0.132 

0  123 

0  133 

0  159 

^-C4 

0.568 

0.338 

0  338 

0.512 

0  512 

0  552 

0  696 

n 

^5 

0  411 

0  244 

0  141 

0.214 

0  259 

0.280 

0  463 

n 
^6 

0  234 

0  139 

0  053 

0.080 

0  113 

0  122 

0  274 

'"i 

^7 

0.141 

0.084 

0  018 

0  027 

0.045 

0.049 

0  198 

^ 
^8 

0  114 

0.068 

0.008 

0  012 

0  024 

0.026 

0  192 

560°F. 

0  077 

0.046 

0  002 

0.003 

0.007 

0.007 

0.179 

1.681 

0.661 

1.101 

108 


2.186^ 

*2 

<*300#2 

2/2 

fty* 

1.186^; 

2,186/3 

ft 

?•+ 

0  025 

0  014 

0.028 

0  0384 

0.0307 

0  034 

0.035 

-C4 

0.171 

0.097 

0.114 

0.156 

0.145 

0.161 

0  187 

^C4 

0.696 

0.396 

0.396 

0.543 

0.543 

0.602 

0  746 

-1 
-"6 

0.383 

0.218 

0.126 

0.173 

0,210 

0  232 

0  415 

"1 
^6 

0.194 

0.110 

0.042 

0  058 

0.082 

0  091 

0  243 

"i 
•^7 

0.119 

0.068 

0.015 

0  021 

0.035 

0  039 

0  188 

"^ 
-'S 

0  097 

0.055 

0.007 

0.010 

0  019 

0  021 

0  187 

*60°F. 

0.072 

0.041 

0.002 

0.002 

0  005 

0.006 

0  178 

1.757 

0.730 

1.070 

v  -  in 


282 


FRACTIONAL  DISTILLATION 
TABLE  9-8     (Continued) 


2.186a80ozJ 

1.186t/a 

2.186x4 

2.186^260^4 

1.186/4 

2.186^ 

c,+ 

0.070 

0.058 

0.059 

0.120 

0.098 

0.099 

i-C4 

0.221 

0.182 

0.208 

0.246 

0.200 

0.226 

n-C* 

0  746 

0.616 

0  760 

0.760 

0.619 

0.763 

C6 

0.241 

0.199 

0.381 

0  210 

0.171 

0.354 

C6 

0.092 

0.076 

0.228 

0  068 

0.056 

0.208 

C7 

0  040 

0  033 

0.182 

0  031 

0.025 

0.174 

C8 

0.022 

0.018 

0.084 

0.017 

0.014 

0.180 

360°F. 

0.007 

0.006 

0.178 

0.004 

0.003 

0.175 

1.439 

1.456 

In  calculating  these  values,  the  same  relations  used  on  pages  277  to 
278  were  employed  with  the  same  relative  volatilities.  Converting  the 
values  of  (OfK/Df}^  to  (0«/Z>)»in  by  the  factors  given  in  Table  9-7  gives 


( *?£  )      =  0.86,  case  I 

\  ///rain 

=  0.8,  case  II 


These  compare  with  0.75  and  0.64,  respectively,  obtained  on  page 
274,  indicating  that  the  variation  of  the  flow  rates  results  in  a  value  of 
(0«/Z>)min  about  20  per  cent  greater  than  would  be  estimated  on  the 
basis  of  constant  molal  flow  rates.  However,  the  mols  of  vapor  that 
would  need  to  be  generated  in  the  still  are  approximately  the  same  by 
both  methods. 

In  this  example,  the  use  of  the  modified  values  did  not  make  a  large 
difference  in  the  results,  but  if  the  design  conditions  had  been  nearer 
the  optimum,  e.g.,  (0R/D)  =  1.26(0^0)^  =  0.94,  then  calculations 
based  on  the  constant  molal  rates  would  be  seriously  in  error.  When 
the  average  latent  heats  of  vaporization  at  the  bottom  and  top  of  the 
column  differ  appreciably,  calculations  based  on  the  modified  basis  are 
to  be  preferred. 

Another  method  of  handling  unequal  molal  overflow  rates  for  multi- 
component  mixtures  is  to  make  the  calculations  on  the  constant  molal 
flow  rates  and  then  to  calculate  the  heat  load  required  to  give  these 
rates  for  a  few  positions  in  the  column.  If  the  heat  input  to  the  still 
is  made  equal  to  the  largest  requirements,  and  the  condenser  and 
tower  cross  section  correspondingly  corrected,  a  conservative  design 
will  be  obtained.  The  check  points  most  commonly  employed  are  (1) 
around  the  still,  (2)  the  entire  section  below  the  feed  plate,  (3)  the  feed 


RECTIFICATION  OF  MULTICOMPONENT  MIXTURES        283 

plate  and  section  below,  and  (4)  around  the  entire  column.  In  general, 
the  modified  latent  heat  of  vaporization  method  is  a  more  satisfactory 
procedure. 

Nomenclature 

C  «  number  of  components 
C*  **  factor  in  Eq.  (9-17) 

D  —  distillate  rate,  mols  per  unit  time 

E  **  factor  in  Eq.  (9-17) 

F  «  feed  rate,  mols  per  unit  time 

G  -  factor  in  Eq.  (9-17) 

H  =  factor  in  Eq.  (9-17) 
HL  «  molal  enthalpy  of  liquid 
Hv  **  molal  enthalpy  of  vapor 
/  =  factor  in  Eq.  (9-17) 

/'  -  factor  in  Eq.  (9-17) 

/  *•  factor  in  Eq.  (9-17) 

/'  «  factor  in  Eq.  (9-17) 

K  »  equilibrium  constant  «  y/x 

m  »  number  of  plates  below  feed  plate  (including  feed  plate) 

N  «  number  of  plates  in  column 

n  =  number  of  plates  above  feed  plate 

0  —  overflow  rate,  mols  per  unit  time 
P  «  vapor  pressure 

p  -  (0/44  -  0/)/F 
Q  «  heat  input 
T  =  temperature 

V  «  vapor  rate,  mols  per  unit  time 
W  =  bottom  product  rate,  mols  per  unit  time 
x  •»  mol  fraction  in  liquid 
y  «=  mol  fraction  in  vapor 
z  «  mol  fraction 
a  **  relative  volatility 

<t>  «  optimum  ratio  of  key  components  on  feed  plates 
TT  «  total  pressure 

Subscripts: 

B  refers  to  benzene 
D  refers  to  distillate 

F  refers  to  feed 

/  refers  to  feed  plate 

h  refers  to  heavy  component 
hk  refers  to  heavy  key  component 

1  refers  to  light  component 

Ik  refers  to  light  key  component 
m  refers  to  conditions  below  feed  plate 
n  refers  to  conditions  above  feed  plate 
R  refers  to  reflux 


284  FRACTIONAL  DISTILLATION 

T  refers  to  toluene 
W  refers  to  bottom  product 
X  refers  to  xylene 

References 

1.  GILLILAND,  Ind.  Eng.  Chem.,  32,  918  (1940). 

2.  GILLILAND,  Ind.  Eng.  Chem.,  32,  1101  (1940). 

3.  GILLILAND  and  REED,  Ind.  Eng.  Chem.,  34,  551  (1942), 

4.  LEWIS  and  COPE,  Ind.  Eng.  Chem.,  24,  498  (1932). 

5.  LEWIS  and  MATHESON,  Ind.  Eng.  Chem.,  24,  494  (1932). 

6  LEWIS  et  al,  Ind.  Eng.  Chem.,  26,  725  (1933);  Oil  Gas  J.,  32,  No.  45,  pp.  40, 
114  (1934). 

7.  RHODES,  WELLS,  and  MURRAY,  Ind.  Eng.  Chem.,  17,  1200  (1925). 

8.  ROBINSON  and   GILLILAND,   "Elements  of  Fractional  Distillation,"  3d  ed., 
McGraw-Hill  Book  Company,  Inc.,  New  York,  1939. 


CHAPTER  10 
EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 

The  design  engineer  frequently  must  separate  mixtures  for  which 
normal  distillation  methods  are  not  practical  due  either  to  the  forma- 
tion of  azeotropes  or  to  a  very  low  relative  volatility  over  a  wide  con- 
centration region.  In  the  first  caserne  separation  is  impossible  unless 
some  special  method  of  by-passing  the  azeotrope  is  employed  (page 
196),  an3  in  the  second  case  excessive  heat  consumption  and  equipment 
size  are  involved.  For  a  large  number  of  such  mixtures,  it  has  been 
found  possible  to  modify  the  relative  volatility  of  the  original  com- 
ponents by  the  addition  of  another  component  (or  components).  This 
technique  has  been  classified  into  two  categories:  extractive  distillation 
and  azeotropic  distillation. 

In  Essentially  all  separations  carried  out  to  date  by  these  techniques 
the  effect  of  the  added  component  is  in  the  liquid  phase,  although  it  is 
possible  to  modify  the  vapor-phase  properties  for  systems  operating  at 
high  pressure. 

The  added  component  by  being  present  in  the  liquid^ghase  can  alter 
the  activity  coefficient  of  the  various  components,  and  unless  the  com- 
ponents already  present  are  identical  in  physical  and  chemical  proper- 
ties, {Kep^entsige^^gem^e  activity  coefficients  will  be  different 
for  each'  component,  thereby  altering  their  relative  volatility.  This 
technique  is  effective  only  when  the  components  in  the  original  mix- 
ture do  not  obey  Raoult's  law,  and  in  general  the  greater  the  deviation 
from  Raoult's  law,  the  easier  it  becomes  to  alter  the  relative  volatility 
significantly  by  the  addition  of  another  component.  Thus  a  considera- 
tion of  the  deviations  from  Raoult's  law  is  essential  for  an  understand- 
ing^oTextractive  and  azeotropic  distillation.  While  no  exact  relation- 
ship has  been  developed  for  the  predictions  of  such  deviations,  the 
Van  Laar  equation  does  give  a  qualitative  picture  that  is  useful. 

As  the  simplest  case,  consider  a  binary  mixture  of  components  1  and 
2  which  is  to  be  modified  by  the  addition  of  component  3. 

The  relative  volatility  of  component  1  to  2  is 


_ 

(X    =    -    —    — 

xi  y* 

285 


286  FRACTIONAL  DISTILLATION 

By  Eq.  (3-58),  page  73, 

r  £3.4.32' 


Tin  72 


+  X*  + 


Over  moderate  temperature  ranges  the  ratio  of  the  vapor  pressures 
of  the  pure  components  does  not  change  appreciably.  The  main 
possibility  of  modifying  the  relative  volatility  lies  in  altering  the  ratio 
of  the  aci^itj_QQefl5cient.  By  subtracting  the  two  activity  coefficient 
equations,  it  is  possible  to  obtain  the  ratio, 

Tin  2 

72 


Consider  the  case  in  which  compounds  1  and  2  are  similar,  i.e., 
almost  obey  Raoult's  law,  e.g.,  ethanol  and  isopropanol.  For  this  case, 
J5i2  will  be  small  and  An  will  be  approximately  equal  to  unity.  Equa- 
tion (10-1)  can  be  simplified  for  these  conditions  to  the  approximate 
relationship, 


(10-3) 


where  V  is  the  effective  volume  fraction  of  the  component  in  the 
mixture. 

The  significance  of  this  equation  can  be  better  shown  by  comparing 
it  with  the  activity  coefficient  ratio,  (71/72)0,  for  the  binary  mixture 
without  the  added  agent.  By  Eqs.  (3-44)  and  (3-45) 

Tlnf— 

where  Fio,  Fao  are  the  volume  fractions  of  components  1  and  2,  respec- 
tively, without  agent  present,  and 

T  In  T^V  -  Bit  [2  ^|jj  (F8)  -  Flo  ~  ^uHo]      (10-4) 


EXTRACTIVE  AND  AZEOTROPIG  DISTILLATION  287 

For  any  effective  addition  agent  the  first  term  of  the  right-hand 
side  of  the  equation  will  be  large  in  comparison  to  the  last  term  and, 
approximately, 

T  In  T^V  =  2  VB^  VW*  V,  (10-5) 

(71/72)0 

For  a  given  binary  mixture,  the^ffectiveness  of  the  added  component 
is  indicated  by  how  much  the  activity  coefficient  group  of  Eq.  (10-5) 
can  be  made  to  differ  from  unity,  and  the  effectiveness  is  increased  by 
large  absolute  values  of  F3,  and  \/#i3.  The  derivation  of  Eq.  (10-5) 
was  for  A 12  approximately  equal  to  unity  and  for  small  values  of  #12. 
A  more  detailed  analysis  of  Eq.  (10-1)  for  the  general  case  leads  to 
essentially  the  same  conclusions;  i.e.,  the  ratio  of  the  activity  coeffi- 
cients is  changed  the  most  when  (1)  \^B\2  is  large,  (2)  \/#i3  is  large, 
and  (3)  Fs  is  large.  However,  in  this  case  it  will  make  a  greater  dif- 
ference whether  \/5i3  and  \/5i2  are  of  the  same  or  opposite  sign. 

A  high  value  of  Fs  is  obtained  by  using  a  large  quantity  of  the  added 
agent,  but  it  is  usually  not  economically  feasible  to  use  a  value  greater 
than  0.8  to  0.9.  To  obtain  these  values  for  F3  requires  a  4  to  9  volume 
of  the  addition  agent  per  volume  of  the  components  to  be  separated, 
and  to  obtain  a  value  of  Fs  equal  to  0.95  would  require  a  ratio  of  19. 
For  most  cases,  the  small  increase  in  the  relative  volatility  obtained  by 
increasing  Fs  from  0.9  to  0.95  does  not  justify  utilizing  twice  as  much 
of  the  added  agent.  In  some  cases,  values  of  Fs  as  low  as  0.4  to  0.5 
may  be  sufficient,  but  higher  values  are  usually  required. 

It  should  be  noted  that,  for  low  values  of  \/Bi2,  the  effective  change 
in  the  ratio  of  the  activity  coefficient  will  be  small.  Thus  this  tech- 
nique would  not  be  effective  for  ideal  solutions  (B^  =  0,  Ai2  =  1), 
and  the  larger  the  value  of  \/rBn,  the  easier  it  would  be  to  obtain  large 
effects. 

The  value  of  the  \/2?i3  is  probably  the  most  important  variable 
under  the  control  of  the  designer.  To  obtain  a  large  modification  of 
the  activity  coefficient  ratio,  it  is  desirable  that  the  absolute  value  of 
the  term  be  as  large  as  possible.  However,  there  is  an  upper  limit 
because  a  value  of  B  from  1,200  to  1,800  corresponds  to  immiscibility. 
While  it  is  necessary  for  the  agent  to  dissolve  only  10  to  20  volume  per 
cent  of  the  mixture,  it  is  doubtful  whether  the  value  of  BU  can  exceed 
2,000  to  3,000.  Qualitatively,  the  value  of  B  is  a  function  of  the  dif- 
ference in  polarity  of  the  two  compounds  under  consideration.  Thus 
a  large  value  of  B  is  associated  with  a  large  difference  in  polarity,  and 
for  the  case  in  question  it  would  be  desirable  to  add  an  agent  which 


288  FRACTIONAL  DISTILLATION 

differed  as  much  as  possible  in  polarity  from  component  1.  This  dif- 
ference could  be  either  a  compound  with  greater  polarity  or  one  with  a 
lower  polarity.  In  the  first  case,  the  A/Bis  would  be  negative  and  in 
the  second  case  positive.  Assuming  that  the  \/rBn  is  positive,  i.e., 
compound  1  more  polar  than  2,  a  positive  value  of  V^is  would  increase 
the  relative  volatility  while  a  negative  value  would  decrease  it.  It  is 
therefore  possible  either  to  decrease  or  to  increase  the  relative  vola- 
tility by  adding  another  component.  The  general  rule  is  (1)  if  the 
added  material  is  more  polar  than  the  components  of  the  original  mix- 
ture, it  will  increase  the  relative  volatility  of  the  less  polar  component 
relative  to  the  more  polar,  and  (2)  if  the  added  material  is  less  polar,  the 
reverse  will  be  true.  For  example,  consider  a  mixture  of  acetone  and 
methanol,  which  forms  an  azeotrope  at  atmospheric  pressure.  In  this 
mixture  methanol  is  more  polar  than  acetone,  and  by  adding  a  more 
polar  component  such  as  water,  it  is  possible  to  increase  the  relative 
volatility  of  acetone  to  methanol  to  such  a  degree  that  an  azeotrope  no 
longer  exists.  However,  if  a  material  of  low  polarity,  such  as  a  hydro- 
carbon, is  added  to  the  mixture,  the  volatility  of  methanol  will  be 
increased  relative  to  that  of  acetone.  Water  would  probably  be  pre- 
ferred to  a  hydrocarbon,  not  only  because  of  cheapness  and  ease  of 
separation  from  the  components,  but  also  because  the  vapor  pressure 
of  acetone  at  a  given  temperature  is  greater  than  that  of  methanol. 
Thus  water  will  act  to  aid  the  natural  difference  in  vapor  pressure  while 
a  hydrocarbon  will  work  against  it. 

For  large  values  of  Fs,  components  1  and  2  behave  as  if  they  were 
each  in  a  binary  mixture  with  component  3,  and  there  is  essentially  no 
interaction  between  1  and  2.  For  these  two  "binary"  systems,  Eq. 
(3-44)  can  be  applied;  for  Fa  approximately  equal  to  unity,  they  reduce 
to 

T  In  ^  =  J513  -  B23  (10-6) 

72 


This  equation  can  be  modified  to  give 


T  In    1  -  VBT*  (VET*  +  V5^)  (10-6a) 

72 

Equation  (10-6)  is  equivalent  to  Eq.  (10-5)  for  small  values  of  Bu 
and  A  12  approaching  unity,  but  it  is  not  limited  to  these  conditions. 
The  use  of  the  binary  principle  for  each  component  with  the  added 
agent  is  a  useful  guide,  both  for  selecting  effective  agents  and  for  esti- 
mating the  quantitative  vapor-liquid  relationships.  For  example, 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION  289 

data  are  available  on  the  vapor-liquid  equilibria  for  the  systems  ace- 
tone-water and  methanol-water.  The  activity  coefficients  for  acetone 
and  methanol  can  be  calculated  for  a  low  concentration,  and  these 
values  will  be  close  to  the  values  for  a  dilute  mixture  of  acetone  and 
methanol  in  water,  thus  allowing  their  relative  volatility  to  be  calcu- 
lated for  these  conditions.  Frequently,  it  is  easier  to  determine  the 
vapor-liquid  equilibria  for  the  tw.o  binaries  with  the  added  agent  than 
it  is  to  investigate  the  three-component  system. 

While  theoretically  it  is  possible  to  add  a  material  of  either  higher 
or  lower  polarity,  both  cases  are  not  usually  equally  attractive  or 
practical.  For  example,  in  separating  a  binary  mixture  of  a  paraffin 
and  an  olefin  of  the  same  number  of  carbon  atoms,  it  is  difficult  to  find 
practical  agents  of  lower  polarity,  and  the  only  real  possibility  is  to  use 
a  material  of  higher  polarity.  Such  an  agent  will  increase  the  vola- 
tility of  the  paraffin  relative  to  that  of  the  olefin,  and  it  is  unfortunate 
that  the  natural  volatility  of  the  olefin  is  often  greater  than  that  of  the 
paraffin.  Thus,  adding  small  quantities  of  the  agent  actually  makes 
the  separation  more  difficult,  and  large  enough  quantities  must  be  used 
to  reverse  the  normal  volatility  completely.  Obviously,  the  reversal 
must  be  sufficient  to  make  the  separation  by  distillation  appreciably 
easier  than  without  the  agent.  For  example,  aqueous  acetone  has 
been  used  for  the  separation  of  butadiene  1,3  from  butylenes.  In  this 
case,  the  solvent  added  is  of  high  polarity  and  increases  the  volatility 
of  the  olefines  relative  to  the  diolefine.  Normally  the  relative  vola- 
tility of  cis-butene-2  (one  of  the  constituents  of  the  mixture)  to 
butadiene  is  0.78,  and  the  value  increases  to  1.3  with  80  volume  per 
cent  of  aqueous  acetone.  A  relative  volatility  of  1.3  gives  an  ease  of 
separation  about  equal  to  a  value  of  0.78,  and  the  use  of  less  than  80 
volume  per  cent  of  aqueous  acetone  would  make  the  separation  of 
these  components  more  difficult  than  for  the  hydrocarbons  alone. 
/  Extractive  and  azeotropic  distillations  both  employ  this  technique 
of  adding  a  component  to  modify  the  volatilities.  They  differ  chiefly 
in  the  fact  that  the  agent  added  in  the  case  of  extractiy^distillatipn  is 
relatively  nonvolatile  as  compared  to  the  other  components  in  the  mix- 
ture, ~lvTmejii^^  distillation  the  volatility  is  essentially  the 
same  asjbhato^  therefore,  forms  an  azeotrope 
with  one  or  moze^ifjbhem  due  to  the  differences  in  polarity.  In  extrac- 
tive distillation  the  agent  is  usua3ly'tfcdded:iiesir^eTop  of  the  column, 
and  most  of  it  is  removed  with  the  liquid  at  the  bottom.  In  azeo- 
tropic distillation  the  agent  is  also  added  near  the  top  of  the  column, 


290  FRACTIONAL  DISTILLATION 

but  in  this  case  most  of  it  is  removed  with  the  overhead  vapor.  The 
distinction  is  not  sharp.  In  the  separation  of  pentane  and  amylenes 
using  acetone  as  the  added  agent,  most  of  the  acetone  is  removed  from 
the  still,  but  it  forms  an  azeotrope  with  the  pentane  and  a  small  portion 
is  carried  overhead.  For  the  purposes  of  this  text,  the  term  extractive 
distillation  will  be  applied  to  those  cases  in  which  the  agent  is  appre- 
ciably less  volatile  than  the  components  to  be  separated  and  in  which 
the  concentration  of  the  agent  is  relatively  constant  from  plate  to  plate 
except  as  affected  by  additions  or  withdrawals  from  the  column. 
Azeotropic  distillation  will  be  used  to  define  those  operations  in  which 
the  added  material  has  a  high  concentration  in  the  upper  portion  of  the 
column  and  then  decreases  to  a  relatively  low  value  in  the  lower  portion 
of  the  unit. 

In  any  given  case  there  are  usually  a  number  of  compounds  that  are 
effective  as  azeotropic-  or  extractive-distillation  agents,  and  the  choice 
depends  on  a  number  of  factors: 

L  Effectiveness  for  modifying  normal  volatility. 

2.  Solubility  relationships  with  system  in  question. 

3.  Cost  of  agent. 

4.  Stability  of  agent. 

5.  Volatility  of  agent. 

6.  Corrosiveness  of  agent. 

7.  Ease  of  separating  agent  from  original  components. 
Extractive  distillation  will  be  considered  first  because  the  mathe- 
matical analysis  for  it  is  simpler  than  for  azeotropic  distillation. 

EXTRACTIVE  DISTILLATION 

This  type  of  operation  has  been  used  for  several  important  separa- 
tions, and  it  is  one  of  the  most  valuable  techniques  in  fractional  dis- 
tillation. Its  earliest  use  probably  was  in  the  distillation  of  nitric  and 
hydrochloric  acids  using  sulfuric  acid  as  the  added  agent  to  aid  in  the 
separation.  It  is  extensively  used,  and  some  of  the  separations  that 
have  been  carried  out  commercially  are  given  in  Table  10-1. 

The  general  arrangement  normally  used  is  shown  in  Fig.  10-1.  The 
feed  is  introduced  into  the  main  extractive  distillation  tower,  and  the 
extractive  agent  is  introduced  a  few  plates  below  the  top  of  the  column. 
These  top  plates  serve  to  remove  the  agent  from  the  overhead  product. 
In  certain  cases,  such  as  the  isoprene-amylene  separation  with  acetone, 
the  agent  cannot  be  completely  eliminated  from  the  overhead  product 
by  this  method  owing  to  azeotrope  formation.  In  such  cases,  other 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


291 


TABLE  10-1.     COMMERCIAL  EXTRACTIVE  DISTILLATION  OPERATIONS 
System  Extractive  Agent 

HC1~H20 ....  H2S04 

HN03-H2O H2S04 

Ethanol-HaO ,        . .   Glycerine 

Butene-butane . .          Acetone,  furfural 

Butadiene-butene Acetone,  furfural 

Isoprene-pentene , Acetone 

Toltiene-paraffinic  hydrocarbons Phenol 

Aeetone-methanol Water 

means,  such  as  extraction,  must  be  employed  to  separate  the  agent. 
The  bottoms  from  the  tower  are  treated  to  separate  the  agent  and  the 
bottoms  product.  Occasionally,  a  portion  of  the  agent  is  added  with 
the  feed  in  order  to  maintain  its  concentration  essentially  the  same 
above  and  below  the  feed. 


Feed 


Top  product 
Extractive  agent 


I 


Bottom 
product 


FIG.  10-1.     Schematic  diagram  of  extractive  distillation  system. 

The  design  calculations  for  such  systems  are  straightforward,  assum- 
ing that  the  physical-chemical  data  on  the  systems  are  available.  In 
the  limiting  case  of  an  essentially  nonvolatile  extractive  agent,  the 
problem  reduces  to  a  standard  binary  or  multicomponent  problem 
depending  on  the  feed  to  the  unit,  except  for  the  modification  of  the 


292 


FRACTIONAL  DISTILLATION 


volatilities.  In  case  the  agent  is  volatile,  the  problem  is  more  com- 
plicated, but  it  can  be  handled  by  methods  analogous  to  those  used  fox 
regular  distillations. 

Total  Reflux.  This  limiting  condition  is  not  equivalent  to  that  for 
regular  rectification  because  to  maintain  a  given  concentration  of  the 
extractive  solvent  in  the  liquid  phase  would  require  the  addition  of  an 
infinite  amount  per  unit  of  feed.  Thus  the  concentration  of  the  com- 
ponents being  separated  would  approach  zero  in  the  bottoms  from  the 


Feed 


I 

•* — EJ *~  Top  product 


Bottom 
product 


Fio.  10-2.     Extractive  distillation  diagram  for  total  reflux. 

extraction  tower.  It  would  require  an  infinite  number  of  plates  to 
reduce  the  concentrations  from  the  finite  values  in  the  upper  part  of 
the  column  to  zero  at  the  bottom,  except  for  the  case  where  the  solvent 
was  nonvolatile.  This  limit  of  an  infinite  number  of  plates  and  infinite 
heat  consumption  does  not  aid  in  orienting  the  design  calculations. 
A  useful  limit  for  orientation  purposes  can  be  based  on  the  desired 
separation  of  the  key  components  from  the  overhead  to  bottoms. 
Thus  the  calculations  should  be  carried  down  the  column  until  the 
desired  bottoms  ratio  of  the  key  components  is  obtained.  Actually 
the  system  can  be  considered  as  shown  in  Fig.  10-2.  Column  A  is  the 
usual  extractive  distillation  unit,  which  obtains  the  desired  degree  of 
separation  of  the  key  components,  and  produces  a  bottoms  containing 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


293 


a  finite  concentration  of  the  key  components  in  the  desired  ratio.     B 

is  some  type  of  unit  that  produces  the  vapor  for  column  A  and  reduces 

the  concentration  of  the  key  components  without  changing  their  ratio. 

Column  C  is  the  tower  that  separates  the  solvent  from  the  bottom 

product.     In  case  the  solvent  is  nonvolatile,  unit  B  is  a  still.    The 

number  of  theoretical  plates  for  tower  A  is  the  desired  answer.    The 

minimum  number  of  plates  at  total  reflux 

will  therefore  be  calculated  on  the  basis  of 

the  desired  ratio  of  the  key  components  in 

the  overhead  and  bottoms.    If  the  relative 

volatility  is  reasonably  constant  over  the 

concentration  range  involved,  Eq.  (7-52)     f6ej 

should    be    satisfactory.     In    any    case, 

the  plate-to-plate  calculation  methods  of 

Chap.  9  can  be  applied. 

Minimum  Reflux  Ratio.     The  calcula- 
tion for  this  case  is  similar  to  that  for  the 

T  W 

usual  multicomponent  mixture,  but  with  I 1^. 

the   extractive   agent   included   at   both     FIG.  10-3.  Extractive  distiila- 
the  pinched-in  regions.     The  asymptotic      lon   iagram- 
value  is  calculated  in  the  same  manner  as  Eqs.  (9-13)  and  (9-15). 
Thus  for  Fig.  10-3,  the  asymptotic  concentration  above  the  feed  is 


Oihk 


XSn    = 


Oihk    —    OiS    —    OiS 


and,  below  the  feed, 


XSm   = 


as  +  oiS(Wxwik/OmXik) 


(10-7) 


(10-8) 


where  x8n,  xsm  =  asymptotic  values  of  solvent  above  and  below  the 

feed  plate  respectively 

Sxs  =  solvent  added  to  system  at  top  of  column 
In  many  cases,  the  mixture  to  be  separated  is  a  binary,  since  it  is 
usually  desirable  to  separate  all  but  two  components  by  regular  dis- 
tillation and  then  subject  them  to  the  extractive  operation.  Generally, 
the  amount  of  solvent  added  at  the  top  of  the  tower  is  varied  with  the 
reflux  ratio  in  order  to  maintain  a  constant  mol  fraction  of  solvent  in 
the  total  liquid  returned  to  the  top  region  of  the  column.  An  alterna- 
tive method  of  operation  is  to  employ  a  given  solvent  rate  independent 


294  FRACTIONAL  DISTILLATION 

of  the  reflux  ratio,  which  would  give  a  solvent  concentration  in  the 
tower  that  varied  widely  with  the  reflux  ratio.  The  first  method  of 
operation  appears  to  be  more  desirable  for  most  cases,  in  order  to 
obtain  a  relative  constant  concentration  of  solvent  in  the  tower  that 
approximates  the  desired  value.  The  equations  for  the  minimum 
reflux  ratio  have  therefore  been  derived  to  be  most  convenient  for  the 
first  method  of  operation.  For  the  general  case  in  which  the  feed  con- 
tains light  and  heavy  components  in  addition  to  the  key  components, 
it  is  recommended  that  the  minimum  reflux  ratio  be  calculated  by 


0«  = 

where  A  = 
H  = 


E  = 
G  <= 

JL             J^DS                                            \^K8             ^DSJ 

Oihk  —  oiS 

1  —  XDS  —  ~ 

t 

1   ~  XRS 
aik       (x 

XDS) 

Fs-  D 

XDS~\^      , 

I     _J—    Tit 

\XRS 

f     I      WXwM 

1          XRS 

,     ^(Wx* 

& 

a|*Lai*  ~"  <*> 

hk        Lj  \a'ik  — 

f 

Of-ik           ' 

*'*    J  +  P 

T  — 

(  DXIM 

G 

•            DXDI 

Dx 

DS      \ 

Oihk  I 
\<Xlk    ~    0 

•1" 

ahk  - 

-  as/ 

OR  =  reflux  from  condenser  before  solvent  is  added  at  top 
XDS,  XRS  =  mol  fraction  of  solvent  in  distillate  and  total  liquid  added 

to  the  top  of  tower,  including  reflux  and  solvent 
Fs  =  solvent  added  with  feed 
a,  af  =  relative  volatilities  for  solvent  concentrations  above  and 

below  feed,  respectively 

If  the  feed  is  binary  mixture,  the  terms  involving  XDI  and  XWH 
are  dropped.  For  certain  special  cases,  simpler  equations  can  be 
employed.  For  example,  if  solvent  is  added  with  the  feed  such  that 
its'  concentration  is  the  same  above  and  below  the  feed  plate,  the  mini- 
mum vapor  required  can  be  calculated  on  a  solvent-free  basis  by  the 
regular  minimum  reflux  equations,  and  then  this  vapor  requirement  is 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION  295 

increased  to  allow  for  the  solvent  in  the  vapor.  In  case  the  solvent 
is  so  high  boiling  that  its  concentration  in  the  vapor  is  negligible,  no 
correction  to  the  solvent-free  calculation  is  necessary.  In  making  the 
calculation  for  the  solvent-free  conditions,  the  relative  volatilities 
employed  are  those  for  the  pinched-in  region  with  the  solvent  present. 
By  equating  Eqs.  (10-7)  and  (10-8)  it  is  possible  to  calculate  the 
amount  of  solvent  that  must  be  added  with  the  feed  to  obtain  equal 
concentrations  above  and  below  the  feed  plate.  For  simplification, 
the  last  terms  in  the  numerator  can  be  neglected  because  they  are  usu- 
ally small. 


sr  -  —  -  —  (1(MO) 


where  SF  =  mols  of  solvent  added  with  feed 

A  =  Om  —  On  —  SF  =  increase  in  overflow  due  to  feed  to  tower 

The  value  of  xan  can  be  obtained  from  Eq.  (10-7),  and  On  can  be 

obtained  from  a  balance  on  the  upper  portion  of  the  tower.     For  the 

case  in  which  the  solvent  rate  is  maintained  proportional  to  the  reflux 

rate, 


I   —  XRS 

Feed-plate  Location.  In  case  solvent  is  added  with  the  feed  to 
maintain  the  same  concentration  in  the  upper  and  lower  sections,  the 
optimum  key  ratio  can  be  calculated  by  the  same  method  as  for  a 
regular  distillation.  If  the  solvent  concentration  is  not  the  same  in  the 
two  sections,  there  are  usually  two  main  factors  tending  to  modify  the 
optimum  ratio.  If  the  solvent  concentration  is  lower  below  the  feed 
plate  than  above,  then  (1)  the  relative  volatility  for  the  key  compo- 
nents may  be  less  favorable  in  the  lower  section  and  (2)  the  mol  fraction 
of  the  solvent  in  the  vapor  is  greater  in  the  upper  section.  The  first 
factor  would  make  it  desirable  to  use  a  low  key  component  ratio  for  the 
feed  plate  to  take  advantage  of  the  better  relative  volatility.  Owing  tc 
the  higher  solvent  vapor  concentration  above  the  feed  plate,  it  would 
be  desirable  to  use  a  higher  key  ratio  than  normal.  The  safest  method 
is  to  calculate  the  feed-plate  ratio  for  a  regular  distillation,  and  in  the 
plate-to-plate  analysis  test  several  plates  in  the  region  to  determine  the 
optimum  condition. 


296 


FRACTIONAL  DISTILLATION 


Concentration  of  Nitric  Acid  by  Extractive  Distillation.  The  con- 
centration of  nitric  acid  by  the  use  of  sulfuric  acid  will  be  used  as  an 
example  of  extractive  distillation  employing  a  nonvolatile  agent.  A 
typical  flow  sheet  of  a  commercial  unit  is  shown  in  Fig.  10-4.  By  the 
oxidation  of  ammonia  and  the  absorption  of  the  nitrogen  oxides  a  62 
weight  per  cent  nitric  acid  is  made.  A  portion  of  this  feed  acid  is 
mixed  with  92  weight  per  cent  sulfuric  acid  and  added  to  the  top  of 
the  tower.  The  remainder  of  the  feed  is  vaporized  and  introduced 
into  the  middle  region  of  the  column.  Direct  steam  is  added  at  the 


nnnn 
Vaporizer 


FIG.  10-4. 


Si  earn 
Concentration  of  nitric  acid 


bottom.  The  overhead  vapors  are  99  weight  per  cent  HN03,  and  an 
over-all  recovery  of  nitric  of  99  per  cent  is  obtained.  The  sulfuric  acid 
removed  from  the  bottom  is  65.0  weight  per  cent.  The  addition  of 
the  feed  to  the  top  of  the  tower  is  unusual  and,  in  general,  is  not  good 
distillation  practice,  but  in  this  case  the  strong  sulfuric  acid  would  cause 
decomposition  of  concentrated  nitric  acid  and  is  therefore  diluted  by 
the  feed.  The  feed  to  the  middle  region  is  vapor  so  that  there  will  be 
less  dilution  of  the  sulfuric  acid.  1.2  Ib  of  92  per  cent  sulfuric  is  used 
per  pound  of  62  per  cent  nitric  concentrated.  The  mixed  acid  added 
at  the  top  is  colder  than  the  tower  temperature,  and  the  condensate 
produced  in  the  column  to  heat  it  serves  as  reflux. 

The  operation  of  this  nitric  acid  system  will  be  analyzed  to  deter- 
mine the  number  of  theoretical  plates  involved.  An  analysis  of  the 
enthalpy  values  indicates  that  the  usual  simplifying  assumptions  will 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


297 


not  be  satisfied,  but  the  deviations  are  not  large  and  the  assumptions 
will  be  used  to  simplify  the  calculations.  The  problem  can  be  solved 
more  exactly  on  an  enthalpy-composition  diagram.  In  general  such 
a  diagram  is  not  suitable  for  a  three-component  mixture,  but  due  to  the 
negligible  volatility  of  the  sulfuric  acid,  a  modified  form  of  the  diagram 
can  be  used.  Equilibrium  data  (Refs.  2,  4)  are  given  in  Fig.  10-5. 


FIG.  10-5. 


02          03         04         05          0.6         07         0.8         0.9 
Mot  fraction  nitric  acid  in  licjuid,su!furic  acid-free  basis 

Vapor-liquid  equilibrium  system,  nitric  acid-water-sulfuric  acid. 


1.0 


These  equilibrium  data  illustrate  the  effect  of  the  added  agent  on  the 
relative  volatility.  With  no  sulfuric  acid  present,  nitric  acid  and 
water  form  a  maximum  boiling  azeotrope  containing  38  mol  per  cent 
acid.  The  62  weight  per  cent  feed  available  is  31.8  mol  per  cent  nitric 
acid.  It  is  therefore  impossible  to  make  99  weight  per  cent  nitric  acid 
from  this  feed  unless  some  method  of  passing  the  azeotrope  is  available. 
The  addition  of  sulfuric  increases  the  volatility  of  nitric  acid  relative  to 


298  FRACTIONAL  DISTILLATION 

water,  and  the  equilibrium  curves  for  liquid  phases  containing  various 
mol  fractions  of  sulfuric  acid  are  given  in  Fig.  10-5.  The  units  of  the 
ordinates  are  expressed  on  a  sulfuric  acid-free  basis.  When  the  liquid 
phase  contains  10  mol  per  cent  sulfuric  acid,  the  volatility  of  nitric 
acid  is  increased  and  the  azeotrope  composition  becomes  12  mol  per 
cent  nitric  acid.  This  sulfuric  acid  strength  could  be  employed  in  a 
two-tower  system  to  give  the  desired  separation.  The  31.8  mol  per 
cent  acid  could  be  treated  to  extractive  distillation  with  10  mol  per 
cent  H2S04  in  the  liquid  to  give  the  desired  concentrated  product  and 
a  bottoms  containing  about  15  mol  per  cent  nitric  acid.  These  bot- 
toms would  be  fractionated  without  sulfuric  acid  being  present  to  give 
water  overhead  and  31.8  per  cent  HN03  as  bottoms  which  would  be 
recycled.  Instead  of  this  two-tower  arrangement,  it  is  found  more 
practical  to  use  more  sulfuric  acid  and  make  the  complete  separation  in 
one  step.  In  order  to  obtain  a  satisfactory  relative  volatility  of  nitric 
acid  to  water  at  the  low  end  of  the  curve  requires  20  to  25  mol  per  cent 
sulfuric  acid  in  the  liquid  phase.  The  actual  acid  consumption  for  25 
mol  per  cent  sulfuric  acid  is  approximately  the  same  as  for  20  mol  per 
cent,  because  the  lower  relative  volatility  for  the  latter  requires  more 
stripping  steam  which  increases  the  acid  requirement. 

Solution.     Basis:  100  mols  of  62  weight  per  cent  nitric  acid. 


xD  -  0.965 
n  .  31.8(0.99) 
D          0.965      - 

Pounds  of  62  per  cent  HN08  -  31.8(63)  +  68.2(18)  -  3,225 
Pounds  of  92  per  cent  H2SO4  -  1.2(3,225)  -  3,870 


Mols  H20  in  with  H2S04  -  3>87°1(8°'08)  .  17.2 
Mols  H20  in  bottoms  - 


Mols  H80  in  feed  -  68.2 
Mols  H2O  in  overhead  -  32.6(0.035)  -  1.14 
Calculating  the  steam  rate,  8,  by  over-all  water  balance, 

8  *  106.5  +  1.14  -  17.2  -  68.2  -  22.2 

It  is  assumed  that  sufficient  nitric  acid  will  be  mixed  with  the  sulfuric  acid  at  the 
top  of  the  tower  so  that  the  combined  stream  contains  60  weight  per  cent  sulfuric 
acid. 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION  299 

Let  xN  »  weight  fraction  of  HNO8  in  top  mixture  and  0.4  —  XN  •"  weight  frac- 
tion of  H2O  in  top  mixture. 
H2O  balance: 


XN  -  0.215 
0.4  -  XN  -  0.185 

TTXT/A      ^    i    ^  3,870(0.92)  («y) 

62  per  cent  IINO3  added  at  top  -       0  62(0  60) 

=  9,570zAr  Ib.  -  2,060 
~  297zjv  Ib.  mols 

-  64(20.35  mols  100  per  cent  HN08) 

Vapor  and  liquid  rates  on  H^SO^free  basis: 

Vn  *  22.2  +  36  -  58.2 

On  »  58.2  -  32.6  -f  64  +  17.2  -  106.8 

Vm  -  22.2 

Om  «  106.8 

xD  -  0.965 

31.8(0.01)          mq 
**  -      106.8      "  °-003 

-  It  is  interesting  to  note  that  for  £  =  22.2  an  enthalpy  balance  gives  Vn  to  the 
top  plate  of  53.2  as  compared  to  the  58.2  obtained  on  the  basis  of  constant  0  and  V 
rates. 

The  mol  fraction  of  sulf  uric  acid  in  the  liquid  phase  is 

36'4         =0.254 


36.4  -f  106.8 

For  vapor-liquid  equilibria  the  curve  for  a  mol  fraction  of  sulfuric  acid  equal  to 
0.25  given  in  Fig.  10-5  will  be  used.     This  curve  is  replotted  in  Fig.  10-6,  and  the 
operating  lines  for  the  calculated  flow  rates  are  given. 
The  upper  operating  line  for  nitric  acid  is 

58.2?/n  -  106.8ztt+i  4-  31.5  -  20.35 
-  106.8*n+i  +  11.15 

The  lower  operating  line  is 


-  0.32 

ym  -  4.8zm+i  -  0.0144 

The  plates  are  stepped*  off  starting  at  yT  «"  0.99  and  continuing  down  to 
x  «  0.003  at  y  =•  0.  Approximately  eight  theoretical  plates  are  required. 

A  more  conventional  extractive  distillation  would  be  to  add  all  of  the  feed  in  the 
middle  region  of  the  tower  and  to  return  a  portion  of  the  overhead  product  as  reflux 
with  the  sulfuric  acid.  Figure  10-7  shows  such  a  system. 

An  exact  comparison  with  the  system  of  Fig.  10-4  can  not  be  made,  but  several 
cases  will  be  evaluated  using  25*4  mol  per  cent  E^SO*  in  the  liquid  phase,  and  the 
same  (0/V)  ratio  below  the  feed  plate, 


300 


FRACTIONAL  DISTILLATION 


Original  system 

—  Case  I 

—  CaseH 
x—  CoseM 


02          03          0.4          0.5          06          07          0.8         0.9          1.0 
Mol  fraction  nitric  acid  in  liquid,  sulfunc  acid-free  basis 

FIG.  10-6.     Design  diagram  for  nitric  acid  concentration. 

Case  I: 

Feed  all  vapor  and  using  92  per  cent  HaSCh  at  top.     Basis :  100  mols  of  feed. 

Vn  -  122.2         On  -  106.8 
Vm  =    22.2        Om  -  106.8 

Composition  of  acid  mixture  refluxed  to  tower,  3,870  Ib.  92  weight  per  cen 
H2SO4  -f  89.6  Ib.  mols  96.5  mol  per  cent  HNO8 


Lb 

Weight 

*Mol 

per  cent 

per  cent 

H2SO* 

3,560 

38.0 

25.4 

H20 

366 

3.9 

14.2 

HN08 

5,430 

58.1 

60.4 

9,356 

EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


301 


Lower  operating  line,  same  as  before. 
Upper  operating  line, 

122.22/n  -  106.8zn+i  +  31.5 
yn  **  0.873.Cn+i  -f  0.2575 

The  upper  operating  line  is  shown  on  Fig.  10-6.  it  requires  about  one  less 
theoretical  plate,  but  the  heat  requirements  would  be  greater  because  of  the 
necessity  of  vaporizing  all  of  the  feed. 


Feed 


rl— 6 

I        Sulfur ic  acid 


Steam 
FIG.  10-7. 


Dilute  suffuric  acid 
Concentration  of  nitric  acid. 


Case  II: 

Feed  36  per  cent  vapor  and  using  92  per  cent  H2SOi  at  top.  Basis:  100  mols  of 
feed.  In  order  to  maintain  the  sulfuric  acid  concentration  constant,  72  percent 
of  it  will  be  added  at  the  feed  plate. 


Vn  -  58.2 
Vm  -  22.2 


On 
Om 


30.4 
106.8 


Lower  operating  line,  same  as  before  but  extends  up  the  diagram  farther  because 
of  the  part  liquid  feed. 
Upper  operating  line, 


i  +  31.5 
y.  -  0.522ajn+i  +  0.541 

This  case  requires  the  same  heat  and  acid  consumption  as  the  original  example 
but  needs  one  additional  theoretical  plate.  With  the  feed  partly  vaporized  it 
would  have  been  better  to  separate  the  liquid  and  vapor  and  introduce  each  at  its 


302  FRACTIONAL  DISTILLATION 

optimum  location.  If  this  change  is  made,  the  system  reduces  to  the  original 
system  of  Fig.  10-4,  because  the  optimum  feed-plate  composition  for  the  liquid 
portion  of  the  feed  is  approximately  the  same  as  the  top  plate  composition. 

Case  III: 

Feed  36  per  cent  vapor  and  using  85  weight  per  cent  H2S04  at  top.    Basis  :    100 
mols  of  feed.    The  sulfuric  acid  will  be  split  as  for  Case  II. 

For  85  per  cent  acid, 

Mol  fraction  H2S04  -  0.51 

For  bottom  acid  strength, 

H2SQ4 


_ 
H2S04  +  (0.49/0.51)H2S04  +  (100  -  32.5)  +  8 

and,  for  same  slope  of  lower  operating  line, 

106.8       (0.49/0.51)H2S04  +  (100  -  32.6)  -f  8 
22.2   ""  8 

S  «•  30.2  mols 
H2S04  -  49.4  mols  '17  mols  at  top,  32.4  mols  at  feed) 

HaSOrfree  basis, 

Vn  *  30.2  +  36  -  66.2         On  -  49.95 
Vm  -  30.2  Om  -  145.1 

Lower  operating  line,  same  as  before  but  the  intersection  of  the  two  lines  will 
be  at  a  different  position. 
Upper  operating  line, 

66.22/n  -  49.950ft+i  4-  31.5 
yn  -  0.755*n+i  +  0.476 

These  operating  lines  are  shown  in  Fig.  10-6.  This  case  requires  fewer  plates 
than  Case  II  but  requires  more  steam.  It  requires  both  more  plates  and  steam 
than  the  original  example.  The  use  of  additional  steam  is  objectionable  both 
because  of  the  increased  steam  consumption  and  because  it  must  be  removed  from 
the  sulfuric  acid  in  the  concentrator. 

The  original  system  is  more  desirable  than  any  of  the  three  cases.  It  is  instruc- 
tive to  analyze  the  possibilities  of  improving  the  original  distillation  system. 
Below  the  feed  plate  it  would  be  desirable  to  reduce  the  steam  consumption  as 
much  as  possible,  but  for  the  25  mol  per  cent  sulfuric  acid,  Fig.  10-6  indicates  that 
the  ratio  of  0/7  for  this  section  cannot  be  increased  significantly  without  increas- 
ing the  difficulty  of  fractionation  excessively.  Using  this  same  slope  (0/V  «•  4.8), 
the  steam  consumption  is 

a      H20  (with  H2S04)  -f  67.4  (from  feed) 

8  --  0  - 

For  the  same  strength  nitric  acid  feed  and  the  given  acid  recovery  and  overhead 
concentration,  the  only  way  to  reduce  8  is  by  reducing  the  water  brought  in  by  the 
sulfuric  acid.  Higher  strength  sulfuric  acid  would  reduce  8  but  increase  the 
difficulty  of  reconcentrating  the  acid.  A  higher  mol  per  cent  sulfuric  acid  in  the 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


303 


liquid  phase  would  make  it  possible  to  reduce  S  but  would  increase  the  acid  recir- 
culation.  None  of  these  alternatives  for  the  lower  section  appears  to  offer  any 
great  advantage  over  the  original  system.  The  concentration  change  per  plate  is 
less  below  the  feed  plate  than  above,  and  it  is  desirable  to  shift  to  the  upper  operat- 
ing line  at  a  low  concentration.  The  original  system  accomplishes  this  result  by 
using  an  all-vapor  feed.  If  all  the  feed  is  added  as  a  vapor  (see  Case  I),  the  frac- 
tionation  in  the  enriching  section  is  very  easy  but  requires  vaporizing  all  of  the  feed. 
The  original  system  reduces  the  heat  consumption  by  vaporizing  only  a  portion  of 
the  feed  and  adding  the  remaining  feed  as  liquid  at  the  top  which  is  approximately 
the  optimum  feed-plate  location  for  the  liquid  feed.  This  still  gives  the  favorable 
intersection  of  the  operating  lines  and  reduces  the  heat  required  for  vaporizing  the 


FIG.  10-8.     Relative  volatility  of  isopropanol  to  ethanol  in  presence  of  water. 


feed.  It  makes  the  separation  a  little  more  difficult  than  using  all  of  the  feed  as 
vapor,  but  it  is  obvious  that  the  extra  theoretical  plate  required  is  well  justified 
by  the  savings  of  heat.  On  the  basis  of  Fig.  10-6,  it  would  appear  that  it  might  be 
advantageous  to  add  more  of  the  feed  as  liquid  at  the  top  and  thereby  reduce  the 
amount  to  be  vaporized.  This  would  necessitate  preheating  the  mixed  acid  added 
at  the  top  by  an  amount  equal  to  the  reduced  heat  input  with  the  vapor  feed,  but, 
low-pressure  waste  steam  might  be  used  for  this  purpose.  Such  a  change  would 
need  to  be  carefully  analyzed  on  an  enthalpy  basis  in  order  to  determine  whether  a 
pinched-in  condition  was  being  encountered  at  the  top  of  the  column.  In  the  case 
of  the  original  system,  one  portion  of  the  feed  was  vaporized  and  the  other  added 
as  liquid  at  the  top.  Some  improvement  would  be  obtained  by  vaporizing  under 


304 


FRACTIONAL  DISTILLATION 


equilibrium  conditions  such  that  the  vapor  feed  would  be  more  dilute  in  nitric' 
acid  and  the  liquid  stronger.  These  two  fractions  could  then  be  added  as  in  Fig. 
10-4,  and  the  operating  lines  would  be  more  favorable. 

In  most  extractive  distillation  cases  the  added  component  is  volatile. 
The  following  example  will  illustrate  the  application  of  general  equa- 
tions developed  in  this  chapter  when  using  an  extractive  agent  of 
appreciable  volatility. 

Separation  of  Ethanol  and  Isopropanol  by  Extractive  Distillation.  A  mixture 
containing  20,  4,  and  76  mol  per  cent  of  water,  isopropyl  alcohol,  and  ethyl  alcohol, 
respectively,  is  to  be  separated  into  an  ethyl  alcohol  product  containing  not  over 


0.30 


0.26 


I  022 


0.18 


0.14 


0.10 


0.06 


1.0 


0.2 


0.4 


0.6 
mols  C2H5OH 


0.8 


1.0 


E     moIsC2H5OH  +  mo!siC3H7OH 
Fio.  10-9.     Relative  volatility  of  water  to  ethanol. 

3.2  mol  per  cent  isopropyl  alcohol  on  a  water-free  basis  with  an  ethanol  recovery  of 
)8  per  cent.  It  has  been  decided  that  water  will  be  used  as  an  extractive  distilla- 
,ion  agent,  and  enough  water  will  be  added  to  the  reflux  to  make  the  liquid  added 
,o  the  tower  85  mol  per  cent  water.  The  feed  will  be  diluted  to  85  mol  per  cent 
vater  before  it  is  added  to  the  tower,  and  it  will  be  heated  such  that  Vn  «=*  Vm. 
Making  the  usual  simplifying  assumptions,  calculate: 

1.  The  minimum  number  of  plates  at  total  reflux. 

2.  The  minimum  reflux.ratio,  0/D. 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


305 


Wafer 


'Dilute  isopropanol 
Mo  I  fraction  of  water -0,85 


20%  wafer 

4%  isopropano/  /' 

76%  ethanol 


Mo/ fraction  of( 
water  '085 


FIG.  10-10. 


Dilute  ethanof 
Extractive  distillation  system  for  isopropanol-ethanol. 


3.  The  number  of  theoretical  plates  required  for  0/D  equal  to  1.5  minimum 
0/D. 

The  equilibrium  data  (Refs.  5,  6)  for  this  system  are  given  in  Figs.  10-8  and  10-9. 
A  schematic  diagram  is  shown  in  Fig.  10-10. 

Solution: 

Let  /  »  isopropyl  alcohol 
E  «"  ethyl  alcohol 
H  =  water 

Basis:  100  mols  of  undiluted  feed. 

Since,  in  the  presence  of  water,  isopopanol  is  more  volatile  than  ethanol,  the 
latter  will  be  largely  in  the  bottoms.  By  an  ethanol  balance 


76(0.98)  =  WXE  -  74.48 


For  design  conditions, 


Wxi 


Wxi  +  WXE 


=  0.002;        Wxi~  0.149 


and,  by  difference, 


DxE  -  1.52;        Dxi  -  3.85 


The  water  in  the  distillate  cannot  be  determined  until  the  composition  of  the 
top  plate  is  known,  and  this  calculation  is  complicated  by  the  fact  that  the  water 
concentration  on  the  top  plate  is  not  known.  The  water  content  of  the  reflux  to 
the  tower  is  0.85,  but  there  will  be  some  change  of  this  on  the  top  plate;  however, 
as  a  first  approximation,  a  mol  fraction  of  water  equal  to  0.85  will  be  assumed  for 


306 


FRACTIONAL  DISTILLATION 


this  plate.    Then  using  the  relative  volatilities  from  Figs,  10-8  and  10-9  a  balance 
is  applied  to  the  top  plate. 


Component 

ytop 

a 

u. 

a 

w 

1.52 

1  n 

1.52 

T 

D 
3.85 

i    >tQ 

D 
2.58 

D 
D  -5.37 

099 

D 
4.54D  -  24  4 

D 

D 

Y2/       4.  547)  -20.  3 

A«                  D 

XH  «  0.85  = 


4.54P  -  24.4 
4.54D  -  20.3 ' 


D  -  10.5 


Water  in  distillate  equals  10.5  —  5.37  **  6.13,  and  for  the  distillate  XE  —  0.145, 

xi  -  0.367        xH  -  0.488 

The  mols  of  water  added  to  dilute  the  feed  «  (80/0.15)  -  100  «  433.     Mols  of 
water  added  at  top  of  tower  «*  OR  f   '    ,    —  l.Oj  »  2.410# 

Solution  of  Part  1.  For  the  condition  of  total  reflux  the  operating  lines  for  the 
alcohols  are  y  »*  3.41$,  and  for  water  y  =  3.41#  —  2.41.  The  water  concentra- 
tion at  the  top  of  the  tower  will  be  approximately  0.85  and  will  increase  slightly 
down  the  tower  because  the  relative  volatility  of  water  to  the  alcohols  increases  as 
the  ethanol  concentration  increases.  The  concentration  at  the  bottom  of  the 
tower  is  assumed  as  0.86. 

For  these  water  concentrations  the  relative  volatility  of  isopropanol  jx>  ethanol  is 
1.5  at  the  top  and  1.64  at  the  bottom.  Thus, 

w   ,   !       log  (0.367/0.145)  (0.998/0.002) 

A\    -f-   I    =»  ..  ,     as    ^Q- 

log  1.57 

Therefore,  approximately  15  theoretical  stages  are  required. 

It  has  been  pointed  out  in  the  discussion  that  this  case  is  inconsistent  in  actual 
bottoms  concentration,  but  the  answer  just  calculated  gives  a  valuable  limit  for 
orientation,  and  the  actual  theoretical  plates  for  0/D  «•  1.5(0/D)min  will  be  about 
50  per  cent  greater. 

Solution  of  Part  2.  The  minimum  reflux  ratio  will  be  calculated  by  three 
methods: 

Method  1.  Since  the  amount  of  water  in  both  the  liquid  and  vapor  phases  is 
relatively  constant  from  plate  to  plate,  the  system  may  be  treated  as  a  binary  to 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION  307 

find  (0/D)min.  Using  the  mol  fractions  of  the  alcohols  on  a  water-free  basis,  which 
will  be  indicated  by  primes,  with  Fn  *  Vm,  at  the  feed  plate,  the  concentration  of 
isopropyl  alcohol  in  the  feed  is  Xj  «•  4/80  »  0.05  and  a/jj.  —1.6  (mol  fraction 
water  -  0.86). 

,;  .      *•*«>•<») 
Vl       1+0.6(0.05) 

on  this  basis, 


0.0776  « 


_  0.716  -  0.0776 
0.716  -  0.050 
0.959         ,9 
x  "  0^41  "  23'2 
D'  -  5.37 
and 

(O^)mm  =*  23.2(5.37)  =  125  mols  of  alcohol  in  overflow  to  feed  plate 
125  +  5.37  =  130.4  mols  of  alcohol  in  vapor 

___  =  833  (assume  XH  at  feed  plate  =»  0.85) 
0.15 


By  Eq.  (10-7)  the  asymptotic  concentration  of  the  solvent  is 

»] 


L0p.41(2«)-5, 


0-875 


*«»  -  i.o  -  0.2 

as  compared  to  0.85  assumed. 

This  value  is  essentially  independent  of  the  value  of  On,  since  the  numerator  of 
the  bracket  is  2.410^  —  5.1,  the  denominator  is  3.410#,  and  the  5.1  is  only  a  small 
correction. 

Thus,  with  xsn  -  0.875 

KM   •    -    125    • 
^»Mm  ^0.125  ' 


-  293 


10.5 


The  value  of  a;^  is  a  little  lower  than  xSn,  and  Eqs.  (10-7)  and  (10-8)  indicate 
that,  for  this  reflux  ratio,  approximately  530  mols  of  water  should  be  added  with 
the  feed  to  make  them  equal  instead  of  the  453  mols  used  to  make  the  mol  fraction 
of  water  in  the  feed  equal  to  0.85. 

Method  2.  By  Eq.  (10-9)  using  same  values  for  relative  volatilities  above  and 
below  the  feed  plate, 

XBS  «  0.85;        XDS  «  0.49 

(1  -  o.49)  -        *'°n3  (0.85  -  0.49) 

0.4 


308  FRACTIONAL  DISTILLATION 

(1  -  0.49)  -  1fiL6fto  (0.85  -  0.49) 

»- SnnS a667 

>i  -       1.0(3.86)        _  lfi 
A  ~  (1.6  -  1.0X0.4) 
u  1.6(74.48)  ^ 

F  "  (1.6  -  1.0)(0.667) 


0667 °-268 


0.4 

OR  .  268  +  V(268)24-4(16)(298)  _  2g5 

OR  __   285  _  07  t 
T3T  ~  10.5 

Using  this  value  of  the  reflux  ratio  it  is  possible  to  calculate  the  ratio  of  the  key 
components  at  the  feed  plate.  In  this  case,  the  value  of  the  ratio  is  0.055  or  only 
slightly  higher  than  the  ratio  in  the  feed. 

Method  3.  A  third  method  is  equating  the  pinched-in  ratio  for  the  key 
components  to  the  intersection  ratio,  <j>.  This  is  not  a  general  method  because  the 
intersection  ratio  may  be  considerably  different  from  the  optimum  matching  ratio. 


A  f     74.48          453  -  5.1\  _ 

U.6  -  1.0  +  1.6  -  0.2; 

_  _- 

ln/     3.85  5.1      \ 

V1'6  "  l'Q      L0  ""  °'2/  = 

" 


<f> 


0.04 

076  " I  -xik-  xsn 


calculating  xtk  by  Eq.  (9-13)  and  x8n  by  Eq.  (10-7). 

1.0  (8.85/Qn) 
0.0526 


1  ~ 


1.6  -  1.0  1.0  -  0.2 


6.41 
°'0526  " 


_ 
3.410u  -  6.41  -  S.OlOa  +  6.41 


The  value  for  OJB  by  Method  3  is  higher  because  the  optimum  key  component 
ratio  is  greater  than  that  used. 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


309 


A  value  of  $  equal  to  0.055  instead  of  0.0526  would  make  the  reflux  ratio  calcu- 
lated by  Method  3  equal  to  those  of  the  other  two  methods. 
Solution  of  Part  3.     Using  a  value  of  27.5  for  (Ou/Z))mm  gives 


and 


1.5(27.5)  -  41.3 


OR  -  41.3(10.5)  -  433  mols 


Water  added  at  top  -  2.41(433)  -  1,042  mols 

Water  out  with  bottom  =  1,042  -f  453  -  5  =  1,490  mols 

On  -  433  +  1,042  =  1,475  mols 

Vn  «  Vm  -  433  4-  10.5  -  443  mols 

Om  -  1,475  +  533  -  2,008  mols 

TABLE  10-2 


Bottoms 

Wxw 

Mols 

Mol  fraction 

2,008 

E 

74.5 

0  0476 

1  0 

0  0476 

0  067 

0  037 

I 

0  149 

0  0000955 

1  96 

0.000187 

0  000263 

0  00007 

W 

1,490.0 

0.952 

0.114 

0.1085 

0.1525 

0  742 

1,564  6 

0  1563 

.  00       ,   Wxw 
Xl  -  0.220,  +  2^8 

a. 

OiXi 

0.22yi 

Xz 

a 

0.222/2 

E 

0  104 

1  0 

0.104 

0  0966 

0  1335 

1  0 

0  101 

I 

0  000333 

1  62 

0  00054 

0  0005 

0  00057 

1.65 

0  000715 

W 

0  895 

0.148 

0.1325 

0  123 

0  865 

0  18 

0  118 

0  237 

Xs 

<x 

0.227/3 

#4 

x& 

x* 

X7 

E 

0.138 

1.0 

0.101 

0.138 

0.138 

0.138 

0  137 

I 

0  000785 

1.635 

0.000935 

0.0010 

0.00127 

0  0016 

0  00196 

W 

0  86 

0.188 

0.1185 

0.86 

0.860 

0.860 

0.860 

x* 

Xg 

#10 

Xn 

Xlt 

Xl* 

Xu 

E 

0.137 

0.137 

0  1365 

0.136 

0  1355 

0.135 

0.134 

I 

0.0024 

0.00293 

0.00355 

0.0042 

0.00505 

0.006 

0  0071 

W 

0.860 

0.860 

0.859 

0.869 

0.859 

0.859 

0.859 

310 


FRACTIONAL  DISTILLATION 


Operating  lines: 
Above  feed  plate  for  alcohols, 

4432/n  •"  l,475a;«+i  +  DXD 
DxD 


1,475'        1,475 
Above  feed  plate  for  water, 


0.00261  for  isopropanol 


-  Xn+i 


«  0.00103  for  ethanol 
5.13  -  1042 


1,475 
-  xn+i  -  0.703 


Below  the  feed  plate  for  alcohols, 

2,008zm+i  -  Wxw 

- 
«  -  xm+i  - 

where  W zir/2,008  equals  0.037  for  ethanol,  0.00007  for  isopropanol,  and  0.742  for 
water. 

The  calculations  are  carried  up  from  the  still  in  the  usual  manner,  with  relative 
volatility  values  from  Figs.  10-8  and  10-9.  The  results  are  summarized  in  Table 
10-2. 

The  intersection  ratio  of  the  key  components  is  0.0526,  and  the  actual  ratio  on 
the  thirteenth  plate  is  0.0444  and  0.053  on  the  fourteenth  plate.  In  this  case,  the 
optimum  ratio  is  slightly  higher  than  the  intersection  ratio,  and  the  fourteenth  plate 
will  be  used  as  the  feed  plate.  The  calculations  are  then  continued  using  the 
enriching  line  equations.  The  results  are  presented  in  Table  10-3. 

TABLE  10-3 


Xl6 

a?™ 

Xn 

#18 

X\9 

£20 

£21 

E 

0.130 

0.126 

0.123 

0.118 

0.1105 

0.101 

0.0886 

I 

0  0087 

0  0113 

0.0154 

0  0215 

0.0301 

0.042 

0  0563 

W 

0.861 

0.862 

0.862 

0.860 

0  859 

0  857 

0  855 

XM 

323 

2/28 

#24 

J/24 

XD 

E 

I 
W 

0.0744 
0.0723 
0.853 

0.060 
0  0889 
0.851 

0  1575 
0.354 

0.488 

0.0463 
0.104 
0.849 

0  119 
0.401 
0  480 

0.145 
0  367 

0  488 

The  x  and  y  values  are  given  for  plates  23  and  24.  It  will  be  noted  that  yn  is 
not  quite  up  to  #D,  but  that  yu  exceeds  it.  Thus  between  23  and  24  theoretical 
plates  in  addition  to  the  still  are  required.  The  water  concentration  on  the  top 
plate  is  close  to  the  assumed  value  of  0.85.  If  this  assumption  had  not  checked 
with  the  calculated  value,  it  would  be  necessary  to  estimate  whether  the  error 
would  materially  affect  the  result.  If  the  correction  was  large,  the  calculations 
might  need  to  be  repeated  to  obtain  a  satisfactory  result. 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


311 


t 


CM       o       oo       to 


\ 


3 


J-d 


O 


312  FRACTIONAL  DISTILLATION 

The  calculated  values  for  the  liquid  phase  are  plotted  in  Fig.  10-11,  There  is  a 
rapid  decrease  in  the  water  concentration  from  the  still  to  the  first  few  plates, 
and  then  the  value  is  relatively  constant  in  the  remainder  of  the  tower.  The 
decrease  through  the  tower  is  small  because  (1)  the  relative  volatility  of  water 
relative  to  the  alcohols  decreases  as  the  ratio  of  isopropanol  to  ethanol  increases 
and  (2)  the  water  concentration  in  the  still  is  high.  The  concentration  of  water  in 
the  bottoms  is  higher  because  the  vapor  removed  from  the  still  contains  a  much 
higher  ratio  of  alcohol  to  water  than  the  liquid  to  the  still. 

Because  the  water  concentration  is  relatively  constant,  approximate  calculations 
could  be  made  on  a  water-free  basis.  The  total  available  vapor  and  liquid  should 
be  decreased  by  an  amount  equal  to  that  required  for  the  water,  and  the  remaining 
vapor  and  liquid  used  to  separate  the  alcohols  as  a  binary  mixture  using  relative 
volatility  from  Fig.  10-8  at  the  assumed  water  concentration.  For  this  case,  the 
result  should  be  reasonably  close  to  the  more  rigorous  method  employed  in  this 
section,  but  it  is  doubtful  whether  the  calculation  is  much  simpler  or  less  time- 
consuming. 

AZEOTROPIC  DISTILLATION 

The  first  commercial  application  of  azeotropic  distillation  was  the 
use  of  benzene  by  Young  (Ref.  7)  for  the  azeotropic  dehydration  of 
aqueous  alcohol,  which  is  still  one  of  the  most  important  applications 
of  this  type  of  operation. 

It  has  been  pointed  out  that  this  system  differs  from  extractive  dis- 
tillation chiefly  in  the  behavior  of  the  agent.  For  example,  consider 
the  continuous  dehydration  of  ethyl  alcohol  by  the  use  of  benzene  as 
the  azeotropic  agent,  as  shown  in  Fig.  10-12.  Tower  1  serves  to 
remove  the  water  from  the  alcohol,  and  tower  2  serves  to  recover  the 
alcohol  and  benzene.  Essentially  anhydrous  alcohol  is  produced  as 
bottoms  in  tower  1,  and  sufficient  plates  are  used  above  the  feed  plate 
to  produce  an  overhead  vapor  that  will  give  two  liquid  layers  on  con- 
densation. The  benzene-alcohol  layer  is  used  as  reflux  for  tower  1, 
and  the  water  layer  containing  small  amounts  of  alcohol  and  benzene 
is  stripped  to  recover  these  constituents.  In  such  an  operation,  the 
agent,  benzene,  must  vary  from  essentially  zero  in  the  still  to  a  rela- 
tively high  concentration  in  the  tower.  Thus  there  is  a  wide  variation 
in  the  solvent  concentration  in  the  tower,  and  some  of  the  approxima- 
tions made  for  extractive  distillation  would  lead  to  serious  errors. 

The  benzene-alcohol-water  system  can  produce  an  overhead  vapor 
that  will  give  two  liquid  phases  on  condensation  which  makes  it  possi- 
ble to  by-pass  the  azeotrope  in  a  manner  analogous  to  that  which  was 
shown  for  partly  miscible  binary  distillations,  and  the  same  type  of 
two-tower  system  is  applicable.  It  should  be  noted  that  the  system 
does  not  produce  the  ternary  azeotrope  as  the  overhead  composition, 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


313 


but  it  is  essential  that  the  condensate  is  two  layers.  With  some  azeo- 
tropic  systems  the  overhead  will  not  give  two  liquid  layers,  and  some 
other  type  of  operation  must  be  used  to  split  the  overhead,  such  as 
extraction  or  dilution. 

In  order  to  illustrate  the  phenomena  involved  in  azeotropic  distilla- 
tion, an  example  will  be  considered  first  and  then  the  various  limiting 


FIG.  10-12. 


1 

Anhydrous 
alcohol 
Azeotropic  system  for  the  production  of  absolute  ethanol  using  benzene. 


conditions  will  be  reviewed.  Consider  the  production  of  anhydrous 
ethanol  using  benzene  as  the  azeotroping  agent.  In  such  cases,  it  is 
found  most  economical  to  concentrate  the  alcohol  by  normal  distilla- 
tion to  almost  the  binary  azeotrope  concentration  before  it  is  intro- 
duced into  the  azeotropic  system. 

Production  of  Absolute  Alcohol  by  Azeotropic  Distillation  with  Benzene. 
For  the  purposes  of  this  example  it  is  assumed  that  the  feed  to  the  dehydration 
system  contains  89  mol  per  cent  alcohol  and  11  mol  per  cent  water  A  two-tower 
system  will  be  employed  similar  to  that  illustrated  in  Fig.  10-12,  and  only  a  single 
liquid  layer  will  be  refluxed  to  each  tower.  Both  towers  will  be  designed  for  an 
overflow  rate  below  the  feed  plate  of  125  mols  per  100  mols  of  vapor,  and  the  usual 
simplifying  assumptions  will  be  made.  The  feed  to  the  alcohol  tower  will  be  such 
that  Vn  =  Vm,  and  it  is  assumed  that  any  condensation  due  to  the  reflux  liquids 


314 


FRACTIONAL  DISTILLATION 

Alcohol 


,  tower No.t  ' 


Alcohol -water 
azeotrope 


-L  iquid  compositions 
below  feed  plate 


\  -  -  ~  Liquid  compositions 
~\     above  feed  plate 

/T  *  *  -  -Benzene  •  alcohol 
azeotrope 


xw,  tower 
No  <? 

Water 


'Benzene 


Fia.  10-13.     Diagram  for  system,  ethanol-benzene-water. 


XA+XH    mol  alcohol  +  mots  water 
Fio.   10-14.     EquUibrium  data  for  system,  ethanol'ibenaene-water. 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


315 


being  at  a  lower  temperature  than  their  boiling  point  is  negligible.     The 

from  the  water  tower  are  to  contain  not  over  0.01  mol  per  cent  alcohol, 

anhydrous  alcohol  is  to  contain  not  over  0.01  and  0.1  mol  per  cent 

and  water,  respectively.    Theoretically, 

it  is  not  possible  to  set  the  exact  bot- 

toms concentration  because  the  compo- 

sition of  the  refluxes  is  limited  by  the 

solubility  relationships;  however,  it  is 

found  that  rather  wide  latitude  is  pos- 

sible in  selecting  the  composition  of  the 

bottoms  product.     For  this  example  it 

is  assumed  that  the  bottoms  are  99.9, 

0.01,  and  0.09  mol  per  cent  alcohol, 

benzene,  and  water,  respectively. 

The  physical-chemical  data  for  this 
system  are  taken  from  Cook  (Ref.  3) 
and  Barbaudy  (Ref.  1).  The  solubility 
data  for  25°C.  are  given  in  Table  10-4 
and  plotted  in  Fig.  10-13.  The  vapor- 
liquid  equilibrium  data  for  the  system 
at  atmospheric  pressure  are  presented 
in  Figs.  10-14  and  10-15. 

It  should  be  noted  that  equilibrium 
data  available  were  not  so  complete  or 
so  consistent  as  would  be  desired,  and 
that  these  two  figures  represent  a 
smoothing,  extrapolation,  and  inter- 
polation of  the  data. 

Solution.     Basis:  100  mols  of  feed. 

Alcohol  balance, 


bottoms 
and  the 
benzene 


0.999Tfi  +  0.0001  W 
Over-all  balance, 

Wi  +  W2  -  100 
Wi  -  89.1 
Wz  -  10.9 


-  89 


FIG.  10-15.     Equilibrium  data  for  system, 
ethanol-benzene-water. 


TABLE  10-4.     PAIKS  OP  TIB-LINE  COORDINATES 


Alcohol 

0  3  and  0.315 

0.225  and  0  23 

0  18  and  0.13 

0.08and0.04 

Benzene  
Water  

0.068  and  0.465 
0.632and0.22 

0.025  and  0.655 
0.75and0.115 

0.015and0.82 
0  805  and  0.05 

0.007and0.94 
0.914  and  0  02 

Tower  1: 


"  IM 

o!**Vm+  89.1  -  1257* 
Vm  -  356        Om  ~  445 
Vn  -  356        On  «  345 


316 


FRACTIONAL  DISTILLATION 


The  calculations  will  be  started  from  the  still,  using  a  basis  of  Om  —  1.0, 
V*  «  0.8,  and  W  *  0.2.  The  results  are  given  in  Table  10-5. 

The  feed  ratio  of  the  key  components  is  l%$  «•  0.123,  and  the  change  from  the 
lower  to  the  upper  section  should  be  made  at  about  this  ratio.  The  ratios  for 
plates  20,  21,  and  22  are  0.075, 0.0935,  and  0.114,  respectively,  from  which  it  would 

TABLE  10-5 
(B  -  Benzene,  A  =  Alcohol,  H  -  Water) 


xw 

0.2z 

a 

axw 

0.82/TT 

Xi 

B 
A 
H 

0.0001 
0.999 
0.0009 

0.00002 
0.1998 
0.00018 

3.6 
0,89 
1.0 

0.00036 
0.89 
0  0009 

0  00032 
0.7989 
0.00081 

0.000344 
0.9987 
0.00099 

0.89126 

a 

OiXl 

0.82/1 

X2 

aX2 

0.82/2 

x& 

B 
A 
H 

3  6 
0  89 
1.0 

0  00124 
0.899 
0.00099 

0.00111 
0.798  * 
0.00089 

0.00113 
0.9978 
0.00107 

0.00406 
0.888 
0.00107 

0.00364 
0  7954 
0.00096 

0.00366 
0.9952 
0  00114 

0.89123 

0.89313 

txXz 

0.82/3 

Z4 

a. 

ax  4 

0.82/4 

x& 

a 

B 

0.0132 

0.0117 

0.01172 

3.4 

0.0398 

0.0354 

0.03542 

3  2 

A 

0.886 

0.7873 

0.9871 

0.87 

0.859 

0.764 

0  9638 

0  82 

H 

0.00114 

0.00101 

0.00119 

1.0 

0.00119 

0.00106 

0.00124 

1.0 

0.89934 

0.89999 

ax& 

0.82/5 

36 

a. 

<xXo 

0.82/5 

x^ 

B 
A 
H 

0.113 
0.79 
0.00124 

0.10 
0.70 
0.0011 

0.10 
0.8998 
0.00129 

2  5 
0.73 
1.0 

0  25 
0.657 
0.00129 

0.22 
0.579 
0.00114 

0.22 
0.7788 
0.00132 

0.90424 

0.90829 

OL 

oe.X^ 

0.8j/7 

38 

a 

aXs 

0.82/8 

B 
A 
H 

1.58 
0.62 
1.0 

0.348 
0.483 
0.00132 

0.335 
0.464 
0.00127 

0.335 
0.664 
0.00145 

0.98 
0.54 
1.0 

0.328 
0.358 
0.00145 

0.382 
0.417 
0.00169 

0.83232 

0.68745 

EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 
TABLE  10-5     (Continued) 


317 


X9 

a 

aX9 

0.82/9 

XlQ 

a. 

aXiQ 

0,8i/io 

B 

0.382 

0.82 

0.313 

0  396 

0.396 

0.76 

0  301 

0.398 

A 

0.617 

0.515 

0.317 

0.402 

0.602   « 

0.5 

0.301 

0.398 

H 

0.00187 

1.0 

0.00187 

0.00237 

0  00255 

1.0 

0.00255 

0.00337 

0  63187 

0.60455 

Xn 

aXn 

0.83/11 

Zl2 

aXiz 

0.82/12 

B 

0.398 

0  302 

0  400 

0  400 

0  304 

0.400 

A 

0  598 

0  299 

0  396 

0.596 

0  298 

0  393 

H 

0.00355 

0  00355 

0  0047 

0.0049 

0  0049 

0  00645 

0  60455 

0.6069 

2?1S 

«Si3 

0.87/13 

Si4 

as  H 

0.8yi4 

S16 

B 

0.400 

0.304 

0  400 

0  400 

0  304 

0.399 

0.399 

A 

0  593 

0.297 

0  392 

0.592 

0  296 

0  389 

0  589 

H 

0.0066 

0.0066 

0  0087 

0.0089 

0  0089 

0.0117 

0.0119 

0.6076 

0  6089 

aXu 

0.82/15 

Xu 

OtXu 

0.82/16 

Xn 

a 

aXn 

B        0  303 

0.398 

0.398 

0  302 

0.395 

0  395 

0.76 

0.301 

A        0  295 

0  387 

0.587 

0  294 

0.384 

0.584 

0  51 

0.298 

H       0.0119 

0.0156 

0.0158 

0.0158 

0.0206 

0.0208 

1.0 

0.0208 

0.6099 

0.6118 

0  6198 

0.82/ 

17 

Sl8               <*Sl8            0.82/18              0?19 

a 

aXig 

0.8yi» 

B           0  389 

0.389     0.296     0.382       0.382 

0  82 

0  314 

0  385 

A           0  385 

0.585     0.298     0.384       0.584 

0.52 

0.304 

0  373 

H           0.0268 

0.027     0.027     0.0348     0.035 

1  0 

0.035 

0.0429 

0.621 

0.653 

S20 

aXzQ 

0.82/20 

xn 

a 

aXzi 

0.82/21 

S22 

B         0.385 

0.316 

0.385 

0.385 

0.82 

0.316 

0.38 

0.38 

A         0.573 

0.298 

0.363 

0.563 

0.53 

0.298 

0  358 

0.558 

H         0.0431 

0  043 

0.0524 

0.0526 

1.0 

0.0526 

0.063 

0.0632 

0  657 

0  6666 

315 


appear  that  plate  22  would  be  the  best  feed  plate.  By  trial  it  is  found  that  plate 
21  is  more  desirable.  This  is  largely  because,  below  the  feed  plate,  the  benzene 
liquid-phase  concentration  is  asymptotic  at  a  value  less  than  0.4,  but  above  the  feed 
plate,  this  composition  rises  rapidly  to  above  0.5.  This  increased  concentration 
increases  the  relative  volatility  of  water  to  ethanol  making  the  fractionation  easier, 
and  it  is  advantageous  to  do  more  of  the  fractionation  above  the  feed  plate. 

The  plate-to-plate  calculations  (Table  10-6)  are  carried  above  plate  21  in  a 
similar  manner.    A  basis  of   1  mol  of  liquid  is  used,   making  D 
0.032  and  F  *  1.032. 

TABLE  10-6.    FEED  PLATE  21 


1.032^/21 

0.032xD 

£22 

a 

a#22 

1.0322/22 

#23 

CL 

axn 

B 

0.49 

— 

0.49 

0  52 

0.255 

0.507 

0.507 

0.475 

0.241 

A 

0.462 

— 

0,462 

0.465 

0  215 

0.428 

0  428 

0.46 

0.1965 

H 

0  0815 

0.032 

0.0495 

1  0 

0  0495 

0  0985 

0.066S 

1.0 

0  0665 

0.5195 

0  5040 

1.0322/23 

£24 

« 

a#24 

1.032^/24 

£25 

a 

«Z25 

1.032s/26 

#26 

B 

0.493 

0  493 

0  55 

0.271 

0.495 

0,495 

0  69 

0  343 

0.52 

0.52 

A 

0.402 

0.402 

0.475 

0.191 

0.348 

0.348 

0.52 

0  181 

0.273 

0.273 

H 

0.136 

0  104 

1  0 

0  104 

0  190 

0  158 

I  0 

0  158 

0  239 

0  207 

0  566 

0  682 

The  top  plate  of  the  tower  is  determined  by  the  fact  that  the  reflux  to  it  must 
correspond  to  the  benzene  layer  of  the  two-phase  region.  In  order  to  illustrate  this 
condition,  the  liquid  compositions  are  plotted  on  the  triangular  diagram  in  Fig. 
10-13.  The  liquid  composition  #25  is  in  the  single-phase  region  while  aj2e  is  in  the 
two-phase  region;  thus  neither  of  them  satisfies  the  condition.  This  means  that 
the  exact  design  conditions  are  not  fulfilled  by  an  even  number  of  theoretical  plates 
and  that  between  24  and  25  theoretical  plates  are  required  since  £25  would  be  the 
reflux  to  plate  24  and  a?26  the  reflux  to  plate  25.  Theoretically,  the  exact  condi- 
tions could  be  satisfied  by  using  a  different  reflux  ratio,  feed-plate  location,  or 
bottoms  composition,  but  the  trial-and-error  procedure  involved  does  not  justify 
the  effort.  It  is  sufficient  to  know  that  between  24  and  25  theoretical  plates  are 
required.  The  x*t  value  could  be  satisfied  exactly  if  a  mixture  of  two  liquid  layers 
was  refluxed,  but  this  is  not  advantageous. 

While  it  is  not  necessary  to  obtain  an  even  number  of  theoretical  plates  in  the 
calculations,  it  is  essential  to  have  a  reasonably  accurate  estimate  of  the  composi- 
tion of  the  overhead  vapor  and  the  reflux  in  order  to  make  the  balances  on  the 
condenser.  A  satisfactory  method  of  evaluating  these  compositions  is  to  plot  the 
compositions,  as  in  Fig.  10-13,  and  use  the  composition  where  the  curve  cuts  the 
two-phase  boundary  as  the  composition  of  the  reflux  to  the  alcohol  tower,  x^ 
and  the  reflux  (or  feed)  to  the  water  tower,  XR^  will  be  the  liquid  in  equilibrium 
with  XRV  From  Fig.  10-13, 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION 


319 


XRl 

XR> 

1.032yri 

B 

0  51 

0.053 

0.51 

A 

0  298 

0.282 

0.298 

H 

0.192 

0.665 

0.224 

As  would  be  expected,  the  values  of  xRl  are  intermediate  between  Xss  and  $«. 

Using  as  basis  an  overflow  rate  of  1  mol  of  liquid  per  unit  time  for  the  water 
tower,  W*  »  0.2  and  F  =  0.8.  The  mols  of  each  component  in  the  overhead 
vapor  from  the  water  tower  are  equal  to  the  difference  between  the  mols  in  the 
reflux  and  the  bottoms.  Thus, 


3*2 

0.2ZTF 

O.Syr 

B 

0.053 

— 

0.053 

A 

0  282 

0  00002 

0.282 

H 

0.665 

0.2 

0.465 

While  both   1. 0322/1^  —  xRl  and  #/?2  —  0.8g/r  have  been  made  equal  to  the 
bottoms  from  the  water  tower,  they  are  on  different  bases  and  it  is  interesting  to 
consider  the  relative  quantity  of  the  two  refluxes. 
For  tower  1 : 

VT  -  OB  +  W a 

IF. 


OBI 


0.032 


For  tower  2: 


Therefore, 


-  VT 


0.2 


0Rl  -  6. 


and        VTl  - 


or,  on  the  basis  of  100  mols  of  feed, 


0Rl  -  345        and 


-  55.2 


The  plate-to-plate  calculations  for  tower  2  are  given  in  Table  10-7. 

The  mol  fraction  of  alcohol  in  #r~3  is  1.5  times  the  maximum  value  specified  while 
in  a?r~4,  the  concentration  is  much  lower,  and  between  four  and  five  theoretical 
plates  are  required.  Due  to  the  high  volatility  of  benzene  in  water,  its  concen- 
tration in  the  bottoms  of  this  tower  would  be  quite  low. 

If  two  liquid  layers  had  been  refluxed  to  match  #26  for  tower  1,  the  composition 
of  the  two  layers  would  be  the  terminal  points  of  the  solubility  tie  line  through  the 
composition  of  £*«.  The  reflux  to  the  water  tower  would  have  the  composition 
of  the  water-layer  end  of  the  tie  line.  In  this  case  this  latter  layer  would  hav0 
been  lower  in  benzene  and  alcohol,  thereby  making  the  f  ractionation  in  the  water 


320 


FRACTIONAL  DISTILLATION 
TABLE  10-7 


Q.SyT 

ar 

O.Syr 

XT 

O.S^r-i 

a 

Q.8yT-.i 

XT-l 

0.8yr-2 

a 

a 

a 

B 

A 
H 

0  053 
0  282 
0.465 

150 
8.0 
1.0 

0.00035 
0.035 

0.465 

0.0007 
0  070 
0.93 

0  0007 
0.07 
0  73 

200 
9.7 
1.0 

0.0000035 
0.0072 
0.73 

5  X  10~« 
0  0097 
0.99 

6  X  10-« 
0.0097 
0.79 

200 
10 
1 

0  50035 

0.737 

> 

0.8yr-2 
a 

XT~Z 

0.8s/r_3 

a 

200 
10 
1 

Q.Syr-8 

XT-3 

0.8yr_« 

0.8yr-4 

XT~4 

a 

a 

B 
A 
H 

2  5  X  lO-s 
0  00097 
0  79 

3  X  10-8 
0  0012 
0  999 

3  X  10-8 
0.0012 
0.799 

1.5  X  10-JO 
0  00012 
0  799 

2  X  10-" 
0.00015 
0  99985 

2  X  10-io 
0  00013 
0  79984 

1Q-12 

0  000013 
0  79985 

10-12 

0  000016 
0.99998 

0.79 

0  799 

0.79986 

tower  easier.  The  reflux  to  tower  1  would  be  two  layers,  but  the  liquid  from  the 
top  plate  would  be  a  single  phase  of  composition  #25.  However,  it  is  possible  to 
operate  with  two  liquid  phases  on  a  plate,  provided  the  mechanical  design  is  satis- 
factory. Thus  in  the  present  case  the  calculations  can  be  carried  past  #28,  but 
allowance  must  be  made  for  the  two  liquid  phases.  For  example,  the  compositions 
of  the  two  phases  corresponding  to  ic26  are  as  follows:1 


fee)i 

(£20)  2 

Oil 

<*(#26)l 

1.032^26 

#27 

B 

0.56 

0  04 

0.54 

0.302 

0.527 

0.527 

A 

0.275 

0.26 

0.45 

0.124 

0.217 

0.217 

H 

0.165 

0.7 

1.0 

0.165 

0.288 

0.256 

0  591 

The  composition  of  the  vapor  was  calculated  on  the  basis  of  (#26)1  because  it  is 
believed  that  the  equilibrium  data  are  more  reliable  in  the  high  benzene  region 
than  in  the  high  water  region.  If  26  theoretical  plates  were  employed  and  two 
liquid  layers  were  refluxed  to  match  #27,  then  plate  26  would  have  two  liquid  layers 
present.  Usually  only  one  liquid  layer  is  refluxed. 

The  tables  of  data  and  the  liquid  compositions  plotted  on  Fig.  10-13  illustrate 
the  factors  involved.  Starting  at  the  bottom  of  tower  1,  the  system  behaves  like  a 
mixture  of  benzene  and  alcohol,  and  the  benzene  concentration  increases  rapidly. 
The  relative  volatility  of  water  is  low  and  does  not  increase  significantly  until  the 
benzene  concentration  has  built  up  enough  to  increase  the  volatility  of  the  water. 

1  These  calculations  are  not  exact,  because  the  solubility  data  given  in  Fig.  10-13 
are  for  25°C.,  while  the  temperature  on  plate  26  is  about  66°C.  The  solubilities 
are  somewhat  different  at  the  two  temperatures,  but  the  25°C.  data  are  used  to 
illustrate  the  principle. 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION  321 

The  water  concentration  then  increases  rapidly  until  the  feed  plate  is  reached. 
Above  the  feed  plate  the  heavy  key  component  (alcohol)  decreases  rapidly,  and  the 
benzene  and  water  attain  values  that  result  in  two-layer  formation.  It  will  be 
noted  that  the  liquid  compositions  are  heading  for  ternary  azeotrope  composition 
as  reported  by  Young  (Ref.  7).  The  composition  of  the  mixed  vapors  to  the 
condenser  is  a  point  on  the  tie  line  through  XR^  and  XR^  and  the  relative  distances 
from  this  composition  to  XR  and  XR%  are  inversely  as  0^  and  0#8. 

The  limiting  conditions  for  azeotropic  conditions  are  not  easily 
expressed  in  analytical  equation,  but  they  can  be  evaluated  for  each 
specific  case. 

Minimum  Number  of  Theoretical  Plates  at  Total  Reflux.  Owing  to 
the  wide  variation  of  the  relative  volatility,  equations  of  the  type  of 
(7-53)  are  not  applicable.  The  number  of  theoretical  plates  required  is 
calculated  best  by  the  plate-to-plate  method  using  y  =  x  as  the  operat- 
ing line  for  each  component.  For  the  benzene-alcohol-water  system 
considered  in  the  preceding  section,  this  plate-to-plate  method  indi- 
cates that  between  12  and  13  theoretical  plates  are  required  at  total 
reflux. 

Minimum  Reflux  Ratio.  This  limit  corresponds  to  a  pinched-in 
position,  or  positions,  in  the  tower.  Because  of  the  wide  variation  in 
relative  volatilities  with  composition,  this  limit  frequently  corre- 
sponds to  a  tangent  condition  of  the  operating  lines  and  equilibrium 
values  rather  than  an  intersection.  In  such  cases  it  is  difficult  to  cal- 
culate the  exact  tangent  condition,  and  each  system  is  essentially  a 
new  problem.  However,  in  a  number  of  cases,  the  minimum  reflux 
ratio  is  determined  by  intersections  of  the  operating  lines  and  the 
equilibrium  values,  and  these  often  occur  near  the  feed  plate,  because 
the  mixture  to  be  treated  is  usually  a  binary  and  the  azeotrope  agent  is 
approximately  constant  above  and  below  the  feed  plate.  For  these 
cases,  the  general  principles  employed  for  multicomponent  mixtures 
can  be  applied.  As  an  example,  consider  the  benzene-alcohol-water 
system  already  studied,  which  has  this  type  of  limiting  condition.  The 
asymptotic  concentrations  below  the  feed  plate  are  given  by  equations 
of  the  type  of  (9-15).  Solving  for  the  values  between  water  and 
benzene, 

r    _ 

Xa  ~ 


_ 

s  -  an  +  aff(W/Om)(xWB/xB) 

The  concentration  of  benzene  in  the  bottoms,  XWB,  is  much  smaller 
than  the  asymptotic  value,  xs,  and  the  last  term  of  the  denominator 


322 


FRACTIONAL  DISTILLATION 


will  be  neglected.  The  value  of  x«  is  much  larger  than  the  numerator 
of  the  right-hand  side  of  the  equation,  and  this  necessitates  <XB  being 
essentially  equal  to  a#.  Thus,  for  this  case  where  XWB  and  XWB  are 
very  small,  the  pinched-in  condition  corresponds  to  the  relative  vola- 
tility of  benzene  to  water  being  unity;  i.e.,  a^n  =  1.0.  A  study  of  Fig. 
10-15  indicates  that  OLBH  ==  1  for  only  a  limited  concentration  range  for 
benzene.  Above  the  feed  plate  the  net  alcohol  and  benzene  removals 
are  very  small,  and  the  same  type  of  analysis  leads  to  the  conclusion  that 
<XB  «*  &*•  The  conditions  as  a  =  1  below  the  feed  plate  and  GAB  =  1.0 


Alcohol 


Wafer 


*  Benzene-  alcohol 


FIG.  10-16. 


above  the  feed  plate  can  be  used  to  evaluate  the  minimum  reflux 
ratio.  One  approximation  for  this  limit  can  be  obtained  by  equating 
the  concentration  ratio  of  alcohol  to  water  for  otB  —  <XH  to  the  feed 
ratio.  The  composition  for  this  condition  can  be  obtained  by  drawing 
a  line  through  XA  =  0.89,  XH  =  0.11,  and  the  benzene  corner  of  the 
diagram.  Where  this  line  cuts  the  an  =  as  line  gives  the  desired 
values.  This  construction  has  been  carried  out  in  Fig.  10-16,  and  the 
intersection  gives  XA  «  0.55,  xa  «  0.07,  and  XB  -  0.38.  From  Figs. 
10-14  and  10-15,  aAa  «  0.54  and  etas  «  1.0.  By  Eq.  (9-15), 


EXTRACTIVE  AND  AZEOTROPIC  DISTILLATION  323 


n  ss  _  WOm)  (0.999) 
0.55  --  !  _  0.54 

5  =  0.253 

vflj 

Om  =  352        and        Vm  =  263 


t  =  26-3  =  L34 
or,  taking  the  net  distillate,  D',  as  11  mols, 

On  =  252        Vn  =  263 
Tfr  )      =  22.8 


The  value  of  Om/Vm  -  1.25  employed  in  the  plate-to-plate  calcula- 
tions corresponds  to 


-- 
D'  "  TT  ~     < 

A  similar  calculation  can  be  made  above  the  feed  plate  using  ou  =  afl, 
and  the  value  obtained  is  0Rl/D'  =  17.5.  This  difference  is  due  to 
the  fact  that  the  ratio  of  the  key  components  was  taken  the  same  as  in 
the  feed.  It  has  already  been  pointed  out  that  the  optimum  feed- 
plate  composition  corresponds  to  a  higher  ratio  of  alcohol  to  water  than 
in  the  feed. 

Another  method  of  calculating  the  minimum  reflux  ratio  is  to  equate 
the  ratio  of  the  key  components  for  the  two  pinched-in  regions.  This 
involves  a  trial-and-error  procedure  to  find  a  composition  on  the 
OLE  =  oiB  line  that  gives  the  same  0RJD'  value  as  a  composition  on  the 
<XB  =  OLA  line  when  the  ratio  of  XA/X  a  is  the  same  at  both.  This  calcu- 
lation gave  0Rl/Df  =  21  for  a  key  component  ratio  of  11  to  1  as  com^- 
pared  to  8  to  1  in  the  feed.  This  last  answer  for  the  minimum  reflux 
ratio  should  be  near  the  true  value. 

The  minimum  reflux  ratios  for  other  cases  can  be  handled  in  a  simi- 
lar manner. 

Nomenclature 

A,B  —  constants  in  Van  Laar  equation 
F  **  feed  rate 
0  «*  overflow  rate 
P  »  vapor  pressure 


324  FRACTIONAL  DISTILLATION 

p  -  (On  -  Om)/F 

T  «•  temperature 

V  «•  vapor  rate 

x  «•  mol  fraction  in  liquid 

y  **  niol  fraction  in  vapor 

a.  «"  relative  volatility 

y  =  activity  coefficient 

Subscripts: 

A  refers  to  alcohol 
B  refers  to  benzene 
D  refers  to  distillate 
E  refers  to  ethanol 
F  refers  to  feed 
H  refers  to  water 
h  refers  to  heavy  component 
hk  refers  to  heavy  key  component 
/  refers  to  isopropanol 
I  refers  to  light  component 
Ik  refers  to  light  key  component 
m  refers  to  below  feed  plate 
n  refers  to  above  feed  plate 
R  refers  to  reflux 
S  refers  to  extractive  agent 
T  refers  to  top  plate 
1,2,3,  refer  to  component  or  plate  number 

References 

1.  BARBAUDY,  Sc.D.  thesis  in  physical  sciences,  University  of  Paris,  1925. 

2.  CARPENTER  and  BABOR,  Trans.  Am.  lust.  Chem.  Engrs.,  16,  Part  1,  III  (1924). 

3.  COOK,  M.S.  thesis  in  chemical  engineering,  M.I.T.,  1940. 

4.  "International  Critical  Tables,"  Vol.  Ill,  306,  McGraw-Hill  Book  Company, 
Inc.,  New  York,  1928. 

5.  SHARPE  and  SIEGFRIED,  M.S.  thesis  in  chemical  engineering,  M.I.T.,  1947. 

6.  TOBIN,  M.S.  thesis  in  chemical  engineering,  M.I.T.,  1946. 

7.  YOUNG,  "Distillation  Principles  and  Process,"  Macmillan  &  Co.,  Ltd.,  London, 
1922. 


CHAPTER  11 
RECTIFICATION  OF  COMPLEX  MIXTURES 

The  design  methods  considered  for  multicomponent  mixtures  in 
Chap.  9  were  based  on  a  limited  number  of  definitely  known  compo- 
nents. In  some  cases,  the  mixtures  are  so  complex  that  the  composi- 
tion with  reference  to  the  pure  component  is  not  known.  This  is  par- 
ticularly true  of  the  petroleum  naphthas  and  oils  which  are  mixtures 
of  many  series  of  hydrocarbons,  many  of  the  substances  present  having 
boiling  points  so  close  together  that  it  is  practically  impossible  to  sepa- 
rate them  into  the  pure  components  by  fractional  distillation  or  any 
other  means.  Even  if  it  were  possible  to  determine  the  composition 
of  the  mixture  exactly,  there  are  so  many  components  present  that  the 
methods  of  Chap.  9  would  be  too  laborious.  It  has  become  customary 
to  characterize  such  mixtures  by  methods  other  than  the  amount  of  the 
individual  components  they  contain,  such  as  simple  distillation  or 
true-boiling-point  curves,  density,  aromaticity  (or  some  other  factor 
related  to  types  of  compounds),  refractive  index,  etc. 

The  simple  distillation  curve  is  the  temperature  as  a  function  of  the 
per  cent  distilled  in  a  simple  or  Rayleigh  type  of  distillation.  This 
type  of  distillation  is  approximated  by  the  laboratory  A.S.T.M.  dis- 
tillation which  is  widely  used  to  characterize  petroleum  fractions. 
The  A.S.T.M.  procedure  gives  some  reflux  and  rectification,  and  the 
results  are  not  exactly  equal  to  the  simple  batch  distillation,  although 
the  difference  is  not  large.  The  temperature  normally  measured  is 
the  condensation  temperature  of  the  vapor  flowing  from  the  still  to  the 
condenser.  Curve  A  of  Fig.  11-1  is  typical  for  the  simple  distillation 
of  a  complex  mixture.  The  temperature  at  any  point  is  the  averaged 
result  of  a  large  number  of  components  and  includes  all  the  effects  of 
nonideality  in  the  solutions.  Thus  in  most  cases  it  is  impossible  to 
relate  such  a  curve  to  the  volatility  of  the  individual  components 
involved.  As  a  result,  such  simple  distillation  curves  are  not  of  much 
direct  value  for  the  solution  of  rectification  problems. 

The  true-boiling-point  curve  is  an  attempt  to  separate  the  complex 
mixture  into  its  individual  components.  Actually  it  is  a  batch  dis- 

325 


326 


FRACTIONAL  DISTILLATION 


tillation  carried  out  under  rectification  conditions.  Usually  a  labora- 
tory distillation  column  equivalent  to  a  large  number  of  plates  is 
employed,  and  the  separation  is  made  at  a  high  reflux  ratio  to  obtain 
efficient  fractionation.  Because  it  is  a  batch  operation,  low  liquid 
holdup  in  the  column  is  important,  and  packed  columns  are  generally 
employed.  Ideally  a  true-boiling-point  curve  for  a  mixture  corre- 
sponding to  the  simple  distillation  curve  A  of  Fig.  11-1  might  be  repre- 
sented by  curve  B  of  this  figure,  showing  individual  horizontal  lines  for 
each  component  with  sharp  increases  in  temperature  in  going  from  one 
constituent  to  the  next.  In  most  petroleum  mixtures  the  curve 
obtained  is  similar  to  B  of  Fig.  11-2,  and  no  definite  steps  are  obtained. 


P«r  C«n*  DisKIIed 
FIG.  11-1. 


100 


15940 

Per Cen+ Drilled  Over 
FIG.  11-2. 


100 


This  result  is  obtained  because  (1)  the  number  of  components  is  very 
large  and  no  single  step  would  be  very  significant  and  (2)  the  degree  of 
fractionation  usually,  employed  is  not  sufficient  to  give  the  sharp  breaks 
in  the  curve.  The  sharpness  of  the  fractionation  between  the  different 
components  is  also  lowered  by  the  formation  of  azeotropes  and  by 
other  solution  abnormalities.  However,  the  true-boiling-point  curve 
probably  represents  a  fairly  high  degree  of  separation  in  most  cases. 
It  is  interesting  to  compare  curves  A  and  B  of  Fig.  11-2.  The  true- 
boiling-point  curve  begins  at  a  lower  and  ends  at  a  higher  temperature 
than  the  simple  distillation  curve  because  the  latter  gives  an  averaging 
effect.  Actually  a  simple  distillation  should  give  the  same  final  tem- 
perature as  the  true-boiling-point  distillation,  because  the  last  material 
to  be  vaporized  should  be  the  pure  highest  boiling  component  in  both 
cases,  but  in  general  the  distillations  cannot  be  carried  to  100  per  cent 
distilled.  A  number  of  methods  have  been  proposed  for  calculating 


RECTIFICATION  OF  COMPLEX  MIXTURES 


327 


the  true-boiling-point  and  simple  distillation  curves  from  each  other, 
and  they  are  useful  in  some  cases,  but  if  solution  abnormalities  are 
involved,  they  can  be  in  error. 

It  has  been  found  possible  to  use  true-boiling-point  curves  to  define 
the  compositions  for  distillation  calculation.  A  fraction  distilling  over 
a  narrow  range  is  taken  as  an  individual  component.  Thus  the  frac- 
tion coming  over  as  distillate  between  39  and  40  per  cent  in  curve  B  of 
Fig.  11-2  might  be  considered  as  a  component,  the  boiling  point  of 
which,  at  the  pressure  at  which  the  distillation  was  carried  out,  being 


Per  Cent  Distilled  Over 
FIG.  11-3. 


100 


the  average  of  the  two  temperatures  corresponding  to  39  and  40  per 
cent.  Ill  this  manner  the  curve  can  be  divided  into  any  desired  num- 
ber of  "components"  with  estimated  vapor-liquid  characteristics  cor- 
responding to  their  distillation  temperature.  These  components  can 
then  be  employed  in  the  distillation  calculations  using  the  various 
methods  given  in  Chaps.  9  and  12.  The  components  in  the  various 
fractions  can  be  recombined  to  give  the  true-boiling-point  curve  of  the 
products. 

If  the  distillate  during  a  true-boiling-point  distillation  were  to  be 
divided  into  two  fractions  at  some  convenient  point,  A,  corresponding 
to  the  temperature  t\,  and  simple  distillation  and  true-boiling-point 
curves  obtained  for  the  two  fractious,  the  results  would  resemble  the 


328 


FRACTIONAL  DISTILLATION 


curves  shown  in  Fig.  11-3  where  curvet  is  the  original  true-boiling- 
point  curve  and  Bf  the  A.S.T.M.  distillation  curve  for  the  same  mix- 
ture. The  true  and  the  A.S.T.M.  curve  of  the  two  fractions  are 
shown  a$  (7  and  t>  dttfves  for  the  more  volatile  and  the  less  volatile 
fractions,  respectively. 

It  will  be  ndted  that  the  initial  temperature  of  curve  D'  is  consider- 
ably highei*  than  the  final  boiling  point  of  curve  C".    This  difference, 


Bottoms  io  Succeeding  Mill 


FIG.  11-4.     Flow  sheet  of  still  and  tower  used  in  test. 

the  so-called  "gap,"  is  frequently  used  as  design  specification  for  the 
separation  desired.  The  averaging  effect  of  the  simple  distillation 
technique  tends  to  give  large  temperature  gaps  even  though  the  lower 
boiling  fraction  may  have  components  that  boil  higher  than  some  of 
those  in  the  less  volatile  fraction.  The  fractionation  in  actual  cases 
will  not  be  so  good  as  assumed  for  Fig.  11-3,  and  there  will  always  be 
some  of  each  component  in  each  fraction.  Thus,  theoretically,  the 
true-boiling-point  curve  of  all  fractions  in  a  given  system  would  begin 
and  end  at  the  same  temperature.  However,  with  reasonably  good 
rectification  it  is  possible  to  obtain  fractions  that  will  give  considerable 
temperature  gaps  by  an  A.S.T.M.  distillation.  In  the  case  of  a  low 
degree  of  separation,  the  A.S.T.M.  curves  for  two  successive  or  adjacent 
fractions  may  give  initial  and  final  temperatures  that  overlap. 


RECTIFICATION  OF  COMPLEX  MIXTURES 


329 


Lewis  and  Wilde  Method.  These  authors  (Ref.  2)  applied  the 
Sorel-Lewis  method  described  in  Chap.  9  to  complex  petroleum  frac- 
tions employing  the  true-boiling-point  curves  combined  with  Raoult's 
law.  Data  were  obtained  in  a  test  on  a  fractionating  column  used  in  a 
petroleum  refinery.  The  plates  in  the  column  Were  9  ft.  in  diameter 
and  fitted  with  the  usual  type  of  bubble  caps.  A  schematic  diagram 
of  the  unit  is  shown  in  Fig.  11-4,  and  some  of  their  data  are  summarized 
in  Table  11-1.  In  the  table,  the  column  called  " Average  boiling 
point'7  is  the  temperature  at  which  the  fraction  as  a  whole  boils  and 
not  the  average  that  would  be  obtained  during  an  A.S.T.M.  distillation. 

Weight  Per  Cent  Over 
2.4      6  .    6      10      l£ .    M.      16      (8     20     22     Z4 


.0      6 


96100 


24-     32     40     46     06     H     72     60    86 
Weight  Per  Cent  Over 

FIG.  11-5.     True-&oiling-point  curves  of  feed  residuum  distillate. 

The  true-boiling-point  curves  for  the  feed,  the  distillate,  and  the  resi- 
due are  given  in  Fig.  11*5.  The  curves  for  the  liquids  sampled  from 
the  plates  are  given  in  Fig.  11-6. 

Lewis  and  Wilde's  method  consists  of  breaking  the  true-boiling-point 
curve  of  the  feed  up  into  fractions  boiling  within  narrow  tempera- 
ture limits.  Thus  the  feed  is  divided  into  10  or  20°F.  fractions  and 
expressed  as  a  component  boiling  between  definite  temperature  limits, 
such  as  420  to  430°F.  fraction  which  is  present  to  the  extent  of  1.5 
weight  per  cent.  Such  cuts  are  then  used  as  individual  components 
by  the  methods  given  in  Chap.  9.  The  true-boiling-point  curve  on 
any  plate  in  the  tower  is  constructed  from  the  calculations  for  that 
plate,  by -simply  recombining  the  cuts  in  the  proportion  that  the  calcu- 
lations indicate. 

The  vapor  above  the  plate  of  an  actual  column  is  not  in  equilibrium 
with  the  liquid  leaving  the  plate  owing  to  inadequate  contact.  In 


330 


FRACTIONAL  DISTILLATION 


JS88 


I    8    §    I   I    I 


RECTIFICATION  OF  COMPLEX  MIXTURES  331 

TABLE  11-1.    SXTMMABY  OF  DATA.  OBSBBVBD  AT  BATTBBY  AND  IN  LABOBATOEY 


Item 

Column 
temp.,  °F. 

Gravity, 
°A.P.I. 

Average 
boiling 
point,  °F. 

Molecular 
weight 

Rate, 
gal.  per 
hr. 

Feed  to  battery  

38.6 

28,920 

Total  gasoline  produced  .  . 

58.4 





10,200 

Gasoline  from  still  4  

330 

50.6 

320 

112 

1,585 

Feed  to  still  4         

446 

30.8 

460 

230 

feesiduum  from  still  4  

490 

29.2 

515 

250 

Kerosene  from  still  5  
Liquid  on  plate  1  

447 

45.3 
32.7 

405 
463 

2,000 

Liquid  on  plate  2  

442 

32.6 

457 

Liquid  on  plate  3  

438 

34.0 

444 

212 

Liquid  on  plate  4  

391 

44.8 

402 

141 

Liquid  on  plate  5.     

373 

46.9 

388 

140 

Liquid  on  plate  6  

370 

47  5 

380 

Liquid  on  plate  7 

360 

47  8 

377 

Liquid  on  plate  8  

358    - 

48.1 

372 

Liquid  on  plate  9 

343 

48.8 

360 

Liquid  on  plate  10  
Reflux  to  top  plate 

340 

49.0 
49.4 

357 
340 

Vapor  from  still  to  bottom  of 
tower  .  .  „  .  .              ... 

485 

43.2 

419 

150 

Gravity  of  cold  oil  through  partial  condenser 38.6°  A.P.L 

Average  rate  of  cold  oil  through  partial  condenser 10,600  gal.  per  hr. 

Average  temperature  of  oil  into  partial  condenser .     .       .          77°F. 
Average  temperature  of  cold  oil  out  of  partial  condenser. . .      181  °F. 

Total  steam  in  vapor  from  tower 100  gal.  per  hr. 

Steam  used  in  heating  feed  to  tower 26  gal.  per  hr. 

Barometric  pressure 758  mm.  of  Hg 

Pressure  at  bottom  of  tower 16  mm.  Hg  above  barometer 

Pressure  at  top  of  tower 23  mm.  Hg  below  barometer 

analyzing  the  behatior  of  an  actual  column,  this  plate  efficiency  must 
always  be  included.  Using  a  plate  efficiency  of  65  per  cent,  Lewis  and 
Wilde  estimated  the  proportion  of  the  420  to  430°F.  component  on  the 
several  plates  above  the  bottom,  and  compared  it  with  the  actual 
amounts  found  in  the  test  as  a  measure  of  the  accuracy  of  their  calcu- 
lations. This  comparison  is  given  in  Fig.  11-7  where  the  curve  repre- 
sents the  calculated  concentration  and  the  points  the  actual  ones  as 
found.  It  will  be  noted  that  very  satisfactory  agreement  was  obtained, 
indicating  the  utility  of  this  method. 

Graphical  Method.    An  alternate  method  (Eefs.  1,  3,  4)  has  been 
proposed  by  which  the  complex  mixture  is  treated  as  a  binary  mixture 


332 


FRACTIONAL  DISTILLATION 


of  components,  consisting  of  the  fraction  above  and  below  the  tem- 
perature 8/t  which  the  cut  is  being  made.  The  vapor-liquid  equilibria 
are  constructed  from  the  characteristics  of  the  true-boiling-point  analy- 
sis or  A.S.T.M.  distillation  curves,  and  the  calculation  is  carried  out 
as  in  the  McCabe-Thiele  method. 

Laboratory  Studies  of  Complex  Mixtures.  Where  laboratory  space 
and  facilities  are  available,  it  is  very  wise  to  design  petroleum  equip- 
ment on  the  basis  of  laboratory  experiments,  using  the  data  thus 
obtained  as  a  starting  point  in  calculations  of  the  sort  just  indicated. 

0.10 


0.09 
0.08 
0.07 


.1  0.05 


\ 

\ 

\ 

420-430  1)eg.f> 
Stock 

\ 

\ 

N 

\  i  '  * 

' 

\ 

360-370  !)<&.< 

i       i 

p. 

s 

\ 

x 

S' 

f'' 

X 

k 

? 

' 

0,02 


01       23456       789      10 
Number  of  Plate  above  bottom 
FIG.  11-7. 

It  is  believed  that  a  rational  analysis  of  laboratory  data  collected  with 
a  thorough  understanding  of  the  requirements  for  subsequent  calcu- 
lations offers  the  safest  method  for  the  study  of  commercial  problems. 

An  illustration  of  such  laboratory  data,  taken  by  Smoley  (Ref.  5),  is 
given  in  the  following  pages. 

A  large-scale  laboratory  column  with  10  plates  was  operated  with 
total  reflux,  so  that  all  of  the  distillate  was  returned  to  the  top  of  the 
column.  Under  this  condition,  with  total  reflux  and  no  distillate,  the 
column  was  operating  with  maximum  separation  per  plate. 

A  mixture  of  benzene  and  toluene  was  distilled  in  this  apparatus, 
and  the  composition  of  the  liquid  on  the  several  plates  determined, 
with  the  results  shown  by  the  solid  line  in  Fig.  11*8.  The  effect  of  the 


RECTIFICATION  OF  COMPLEX  MIXTURES 


333 


efficiency  of  the  actual  plate  as  contrasted  with  the  theoretical  plate 
is  shown  in  the  same  figure.  The  dotted  line  was  obtained  by  plate- 
to-plate  calculations  at  total  reflux  using  theoretical  plates.  Thus,  to 
produce  a  90  mol  per  cent  distillate  requires  about  10  steps  in  the 
actual  column,  whereas  the  same  effect  is  obtained  in  6  steps  in  the 
perfec^  column,  indicating  a  plate  efficiency  of  around  60  per  cent. 

Th^s  sai&0  column  was  then  operated  in  the  same  way  but  using  a 
mixture  of  benzene,  toluene,  and  xylenes  so  as  to  produce  as  high  a 
concentration  of  benzene  as  possible  in  the  condenser  and  to  segregate 
the  xylenes  in  as  concentrated  a  form  as  possible  at  the  bottom.  The 


Condenser 
Top  PI  ate 
Plate  $ 

e 

7 
6 

6 
4 
3 
Z 

Plate  J 
Still 


1.0 


O.Z    03     0.4-     0.5     0.6     0.7     0.&     0.9 

Mol.  Fraction  of  Benzene 
FIG.  11-8.     Operation  of  benzene-toluene  column. 

results  are  shown  in  Fig.  11-9.  The  amount  of  xylenes  was  small  so 
that  a  large  proportion  of  toluene  was  present  on  the  bottom  plate. 
The  highest  concentration  of  toluene  occurred  on  the  fifth  plate,  this 
component  thus  tending  to  segregate  in  the  column,  Under  ordinary 
conditions,  a  column  would  be  operated  at  a  lower  temperature  level  so 
that  the  benzene  at  the  top  would  have  contained  less  toluene,  thus 
delivering  the  toluene  and  xylene  together  from  the  bottom  for  subse- 
quent separation  in  a  second  column.  This  experimental  column  had 
insufficient  plates  to  do  this. 

The  column  was  then  operated  with  total  reflux  on  a  cracked  petro- 
leum distillate  obtained  from  a  Winkler-West  Texas  crude  oil.  The 
liquid  samples  from  the  several  plates  in  the  column  were  then  analyzed 
in  a  true-boiling-point  still,  being  separated  into  components  of  5°C. 


334 


FRACTIONAL  DISTILLATION 


boiling-point  range.  Each  of  these  components  was  indicated  by  its 
mid-temperature.  Thus  a  component  boiling  on  the  true  boiling- 
point  apparatus  between  75  and  80°C.  was  called  the  77.5°C.  com- 
ponent. The  results  of  this  experiment  are  given  in  Fig.  11-10,  where 
each  component  is  indicated  by  a  concentration  curve. 

It  will  be  noted  that  each  component  tends  to  segregate  in  the  col- 
umn, the  segregation  point  depending  on  its  boiling  point.  This  segre- 
gation of  a  component  in  a  continuous  column  is  the  basis  for  the  type 
of  still  frequently  found  in  petroleum  refineries  where  streams  or  cuts 


JopPlcrh 
Phte  9 
8 

7 

e 

5 
4 
3 
2 
Mate  / 

X 

x 

X 

x 

^ 

x 

^ 

/* 

X 

ft* 

e^ 

' 

Px 

/ 

\ 

/ 

\ 

\ 

I 

ft 

y 

f 

\ 

% 

y 

^ 

, 

, 

/ 

FIG.  11-9. 


0     0.1     02    03    OA    0.5    0£     0.7     0.6    09     1.0 

Mot.  Fraction  of  Coroponenf 
Operation  of  column  on  benzene-toluene-xylenes  mixture. 


are  taken  from  the  central  portions  of  the  column  as  well  as  from  the 
top  and  bottom.  It  is  evident  from  Fig.  11-10  that  such  side  cuts 
cannot  be  all  pure  or  free  from  other  components;  in  the  commercial 
column,  where  total  reflux  is  n6t  employed,  the  segregation  is  much 
less  pronounced  than  is  indicated  in  Fig.  11-10. 

The  maxima  ift  concentrations  of  the  fractions  shown  in  Figs.  11-9 
and  11-10  are  definitely  related  to  the  volatility  of  the  component  in 
question  and  the  temperature  in  the  distillation  column.  These  curves 
were  obtained  at  total  reflux,  and  for  this  condition  the  composition 
of  the  vapor  entering  a  plate  is  equal  to  the  liquid  leaving  the  plate  for 
all  components,  i.e., 

yn  =»  #n-fl 

and 

y*  «  K  «a» 


giving 


RECTIFICATION  OF  COMPLEX  MIXTURES 


Kn 


For  the  position  in  the  tower  at  which  a  component  is  going  through 
a  maximum  concentration,  the  value  of  the  liquid  composition  on  sue- 


Curve  No.       Ave.  B.P  cfCo/nporterti 

/                           ?7.5*C. 
2                            67.S*C. 
J                            97.5'C. 
4                            f07.59C. 

6                         I27.$°C. 
8                         Mf>C 

Condenser 

TopP/at* 
Ha+e  9 

S 

\ 

3 
2 

f>/crht 
WJ/ 

N 

^ 

2\ 

^ 

V 

\ 

y 

^ 

s^ 

\ 

/ 

\ 

/ 

^ 

v 

\ 

/ 

x 

\ 

^ 

\^ 

s^ 

, 

(/ 

\ 

\-- 

^x 

^-^ 

/ 

' 

N 

f 

^ 

'\ 

\ 

^> 

s 

\ 

^ 

\ 

.0 

\ 

/ 

1 

^ 

/ 

N 

\ 

7 

I 

^ 

^ 

1  / 

s\ 

y 

J*"*"^ 

DU 

/^ 

f 

> 

^^ 

r 

X 

X 

aoa  ao4  0.06  aoa  aio  aiz  O.M-  o,i&  aie 

Mol  Fraction  <tf  Component 
FIG.  11-10.     Operation  of  column  on  petroleum  distillate. 

cessive  plates  is  approximately  the  same,  making  Kn  =  1.0.  Thus  the 
maximum  occurs  at  the  position  in  the  column  where  the  temperature 
is  such  that  the  equilibrium  constant  of  the  component  is  equal  to  1.0. 

References 

1.  BROWN,  Chem.  Eng.  Congress,  World  Power  Conf,,  2,  324  (1936). 

2.  LEWIS  and  WILDE,  Trans.  Am.  Inst.  Chem.  Engrs.,  21, 99  (1928). 

3.  PETERS  and  OBBYADCHIKOV,  Nestyanoe,  Klozyaistro,  24,  50  (1933), 

4.  SINGBB,  WILSON,  and  BROWN,  Ind.  Eng.  Chtm,,  28,  824  (1926), 

5.  SMOLEY,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1930. 


CHAPTER  12 

ALTERNATE  DESIGN  METHODS 
FOR  MULTICOMPONENT  MIXTURES 

In  Chap.  9  the  Lewis  and  Matheson  procedure  for  SorePs  plate-to- 
plate  method  was  presented.  Many  other  design  methods  have  been 
proposed  based  on  alternate  methods  of  analysis  or  approximations. 
None  of  them  illustrates  the  phenomena  involved  in  multicomponent 
rectification  so  well  as  the  Lewis  and  Matheson  ihethod.  A  number 
of  the  methods  require  less  effort  to  obtain  certain  design  factors  than 
the  stepwise  procedure  and  are  useful  in  cases  where  similar  systems  are 
to  be  analyzed  repeatedly.  When  a  new  type  of  problem  is  to  be  con- 
sidered, the  information  obtained  by  the  plate-to-plate  method  is  well 
worth  the  effort  involved.  Actually  a  detailed  analysis  by  methods  of 
Chap.  9  does  not  usually  require  over  a  few  hours,  and  the  confidence 
in  the  result  and  the  insight  obtained  of  the  operation  justify  the 
effort  involved. 

The  space  available  in  this  text  does  not  allow  a  detailed  analysis  of 
these  various  design  methods,  but  a  brief  review  will  be  given  of  some 
of  them.  A  number  of  the  methods  involve  assumptions  that  are  not 
justified  in  many  cases,  and  the  design  engineer  must  appreciate  these 
limitations  or  misleading  results  will  be  obtained. 

PLATE-TO-PLATE  METHODS 

In  addition  to  the  Lewis  and  Matheson,  and  Lewis  and  Cope  meth- 
ods given  in  Chap.  9,  plate-to-plate  procedures  have  been  given  by 
Thiele  and  Geddes  (Eef.  14)  and  Hummel  (Ref.  9). 

The  Thiele  and  Geddes  method  is  a  stepwise  procedure  based  on 
using  a  ratio  of  the  concentration  of  a  component  to  its  terminal  con- 
centration. Starting  at  the  top  of  the  column,  for  any  component, 

XT  —  %£-  =  «?•  (for  total  condenser) 
/Vr        /ir 

336 


ALTERNATE  DESIGN  METHODS  337 

and 


XD 
In  general, 


j/n    __    Qn+1  /Wl  i\     ,      1 

—  •  x    t    -j—    j. 

XD          Vn    \XD  / 


and 

?  =  7TT-  (12-3) 

XD          Jtv  nXo 

Thus,  if  the  values  of  the  equilibrium  constants  and  the  ratios  of 
0/V  are  known  for  each  plate,  it  is  possible  to  calculate  the  ratio 
XH/XD  for  any  plate  above  the  feed  plate  without  knowing  the  value 

ofXD. 

Below  the  feed  plate  a  similar  analysis  can  be  made. 

Xw 
and,  from  the  operating  line, 

—  =  ^  ( —  -  1 )  +  1  (12-4) 

xw       0i  \xw        /  v        ' 

in  general, 


Xw 


jfe  «  A 

Xw  / 


+  1 


The  calculations  can  be  carried  up  from  the  still  and  down  from  the 
condenser  to  the  feed  plate,  giving  values  of  ym/xw  and  yn/xo.  Assum- 
ing that  the  feed  plate  is  such  that  the  vapor  and  liquid  leaving  are  in 
equilibrium,  then  the  values  of  the  vapor  composition  in  the  two  ratios 
must  be  equal  and  the  value  of  XD/XW  can  be  calculated  for  each  com- 
ponent. This  ratio  can  be  used  to  calculate  D  and  W. 


338  FRACTIONAL  DISTILLATION 

Thus,  for  each  component, 

DXD  +  Wxw  =  FzF 

Wxw  * — ^ 


W 


1  4-  F~D  txw 

"*"      D 
Summing  the  Wxw  and  Da:i>  terms  for  all  components, 

i  — 


(12-7) 
(12-8) 


(12-9) 


where  S  indicates  the  sum  of  the  terms  for  all  components.  With  the 
value  of  XD/XW  for  each  component,  W/F  or  D/F  can  be  evaluated. 
As  compared  to  the  Lewis  and  Matheson  method,  this  method  has 
the  advantage  that  it  is  easier  to  calculate  the  separation  to  be  obtained 
for  a  given  number  of  theoretical  plates  at  a  specified  reflux  ratio.  In 
case  the  separation  and  reflux  ratio  are  specified,  the  Lewis  and 
Matheson  method  is  the  easier  to  apply. 

Thiele  and  Geddes  Calculation  for  Benzene-Toluene-Xylene  Separation.  To 
illustrate  the  application  of  this  method,  consider  the  separation  of  the  benzene- 
toluene-xylene  mixture  of  page  219  in  a  tower  having  five  theoretical  plates  with 
the  feed  entering  the  middle  plate.  The  reflux  ratio  0/D  will  be  2.0,  Vn  *  7Wj 
the  vapor  leaving  the  feed  plate  will  be  in  equilibrium  with  the  liquid  leaving,  and 
the  usual  simplifying  assumptions  will  be  made. 

Solution.    Basis:  100  mols  feed.    0/D  «  2,  (0/V)n  -  0.667. 


XP 

Mols 
feed 

Distil- 
late 
comp. 

Assume 
T  - 
85°C. 
KT 

Xr  =   * 
XD      KT 

3K£=1  .  0.667  (  ^  -A+l 
XD                 \XD        J 

Benzene.  .  .  . 

0,60 

60 

XDB 

1.15 

0.87 

0.913 

Toluene  .... 

0.30 

30 

XDT 

0.452 

2,21 

1.807 

Xylene  

0.10 

10 

XDX 

0.184 

5.44 

3.96 

ALTERNATE  DESIGN  METHODS 


339 


Assume 
T  -  85°C. 
Kr-i 

XT-I  ^     1     (yT-i\ 

XD  **  /Cr-i  \  #£>  / 

yr-a 
a?z> 

Assume 
T  -  90°C. 
•Kr-2 

sr-a 

«D 

gr-t 
JCD 

c. 

1.15 

0.793 

0.862 

1.38 

0.649 

0.166 

C7 

0.452 

3.99 

3.0 

0.533 

5.63 

4,09 

C8 

0.184 

21.5 

15.35 

0.221 

69.4 

46.6 

From  still  up,  assume  D  «  66;  then  Vn  »  198  -  Vm,0m  **  232,  (FW/0TO)  -  0.853. 


C6 
Cr 

Assume 
T  -  115°C. 

Xw        '        \Xw        ) 

Assume 
T  =  110°C. 

£ 

£ 

Assume 

T  m  110°C. 

2.62 

1.12       r 

2.38 

1.102 

2.29 
0.975 

5.45 
1.075 

4  8 
1  064 

2.29 
0.925 

Cg 

0.513 

0  587 

0.435 

0  255 

0.365 

0.435 

2/2 

a?8 

Assume  T  *  100°C. 

^3 

xw 

xw 

#3 

xw 

C6 

11.0 

9.53 

1.745 

16.65 

C7 

1  038 

1.032 

0.735 

0.76 

C8 

0.159 

0.282 

0  316 

0.104 

Equating  y9  =*  2/r-s  gives 


C7 

Cg 


By  Eq.  (12-9), 


1  « 


0,6 


21.8 
0.186 
0.0022 


0.3 


-f; 


0.1 


~20.8(ff  /F)  -f  21.8  ^  0.814(TT/F)  +  0.186   '  0.9978(TT/F)  +  0.0022 


-  0.334 


The  section  below  the  feed  was  calculated  on  the  basis  W/F  «»  0.34  compared 
to  the  calculated  value  of  0.334,  and  it  will  not  be  rechecked.  This  is  one  of  the 
difficulties  with  the  Thiele  and  Geddes  method,  i.e.,  specifying  the  reflux  ratio  and 
feed  condition  still  leaves  trial  and  error  for  both  the  plate  temperatures  and  the 
ratio  of  0/V  below  the  feed  plate. 


340  FRACTIONAL  DISTILLATION 

The  temperature  assumptions  will  now  be  reviewed. 


XD,  by  Eq.  (12-8) 

XT 

#r-i 

ZT-I 

sr-s 

t 

C6 
C7 

C8 

0.88 
0.12 
0.00065 

0.765 
0.265 
0.0035 

0.803 
0.217 
0.0025 

0.698 
0.479 
0  014 

0.571 
0.676 
0.045 

1.0335 

1.0225 

1.191 

1.292 

The  fact  that  the  sum  of  XT  is  greater  than  1.0  indicates  that  the  actual  tempera- 
ture is  slightly  higher  than  85°C.  The  assumed  temperatures  for  both  plates  T  —  1 
and  T— 2  are  considerably  low  as  indicated  by  the  sums  of  XT-I  and  XT-Z  being 
greater  than  1.0,  and  the  results  should  be  recalculated  for  a  satisfactory  design. 

Below  the  feed  plate, 


xw 

y« 

2/1 

2/2 

2/3 

C6 

Cr 
C8 

0.042 
0.658 
0.300 

0.11 
0  737 
0.153 

0.229 
0.707 
0.076 

0.462 
0.693 
0.048 

0.699 
0.50 
0.031 

1.000 

1.012 

1.203 

1.230 

The  assumed  temperatures  for  the  still  and  first  plate  are  satisfactory,  but  the 
assumed  temperature  of  plate  2  is  too  high.  The  sum  of  3/3  is  larger  than  1.0,  and 
this  would  appear  to  indicate  that  the  assumed  temperature  was  much  too  high, 
actually  most  of  the  excess  is  from  plate  2.  The  fact  that  Sy2  >  1.0  made 
Sa?a  >  1.0  and,  even  if  the  assumed  temperature  for  plate  3  were  correct,  Sg/s  will 
be  greater  than  1.0.  Thus  an  error  in  the  assumed  temperature  for  one  plate 
carries  through  succeeding  plates.  It  is  obvious  that  the  calculation  requires  con- 
siderable trial  and  error. 

A  method  similar  to  the  Thiele-Geddes  method  has  been  proposed 
by  Hummel  (Ref.  9).  In  this  method  plate-to-plate  calculations  are 
made  for  a  few  plates  at  each  end  and  around  the  feed  plate  to  estab- 
lish the  temperatures.  With  the  values  and  the  known  number  of 
theoretical  plates  the  temperature  gradient  in  the  tower  is  drawn. 
This  then  gives  the  temperatures  to  employ  in  a  Thiele-Geddes  type  of 
calculation.  For  a  given  number  of  plates  and  a  given  reflux  ratio, 
this  method  requires  an  estimation  of  the  distillate,  bottoms,  and  feed- 
plate  compositions.  Basically,  Hummel's  method  furnishes  a  sys- 
tematic method  of  successive  approximations  for  the  plate  tempera- 
tures to  be  used  in  evaluating  the  equilibrium  constants. 


ALTERNATE  DESIGN  METHODS  341 

REDUCED  RELATIVE  VOLATILITY  METHODS 

Underwood  (Ref.  15)  and  Gilliland  (Ref.  5)  have  proposed  design 
methods  for  applying  a  total  reflux  type  equation  with  a  reduced  rela- 
tive volatility.  For  the  enriching  section  of  the  tower, 


-a, 

T 


where  T  refers  to  top  plate.     For  a  total  condenser, 


XB/D  XB/T 

and,  by  the  operating  material  balances, 


= 
i        1  +  (D/0)(XBD/XST)  \XB 

=  fr-i(fO  (12-10) 

\XB/T 

and 


1 

where 


«/+A  (  «/+_i\  (  ^A  (%A\ 

/W  \  /»/W  V  /s//  W/ 


Mn   "    1   +    (D/On)(xBD/XBn+l) 

and  below  the  feed  plate, 


xj\         ctf-i(af-2\  (ai\(ai\(<x  w\  (x  A 

A"S^WV        WWWWr 


where 


To  simplify  the  calculation,  average  values  of  a  and  0  are  employed, 


where  ATn  ^  number  of  theoretical  plates  above  feed  plate 

JVTO  «  number  of  theoretical  plates  below  feed  plate  including 
feed  plate 


342  FRACTIONAL  DISTILLATION 

Arithmetic  averages  have  been  employed  for  a  and  /3 

a,  +  (q/0), 


2  (12-17) 


(12-18) 


The  values  of  £/  and  /9/_i  involve  the  composition  on  the  feed  plate 
and  the  plate  above.  The  feed-plate  composition  is  determined  by 
assuming  that  the  components  more  volatile  than  the  key  components 
are  negligible  in  the  bottoms,  that  those  less  volatile  are  negligible  in 
the  distillate,  and  that  the  concentration  of  both  the  light  and  heavy 
components  are  asymptotic  at  the  feed  plate. 

Thus,  for  the  more  volatile  components, 


Vy/  =  Oxf+i  +  DxD 
Using  y/  =  Kfxf  and  x/+i  =  x/, 

VKfXf  =  Ox/  +  DXD 

_ 

Xf  - 


(KfVn/On)  - 
Similarly,  for  the  less  volatile  components, 

(W/Om)xw 


Xf 

The  values  of  x/+i  needed  for  the  calculation  of  ft/  are  obtained  by 
stepwise  calculation  from  #/.  For  approximate  values,  £/  can  be  cal- 
culated with  Xf  instead  of  x/+i. 

Reduced  Relative  Volatility  Calculations  for  Benzene-Toluene-Xylene  Separa- 
tion. The  benzene-toluene-xylene  example  of  page  219  will  be  solved  by  this 
method. 

Solution.     Basis:  100  mols  of  feed.     (See  page  220  for  design  quantities.) 
Estimation  of  feed-plate  composition.     By  Eq.  (12-20)  the  mol  fraction  of  xylene 
in  the  feed  plate  is 

„ 
Xfx 


1  -  (Kf-iVJ/OJ 

10/220.2 
*  1  -  Kf-i  (180.3/220.2) 

Assuming  Kf^  ~  K/  -  0.22  (t  -  90°C.), 

x/x  •»  0.0555 


ALTERNATE  DESIGN  METHODS 


343 


This  compares  with  a  value  of  0.058  obtained  by  the  stepwise  calculations,  page 
223.    A  better  check  would  be  obtained  using  2C/-i. 
By  mol  fraction  balances, 

X/B  +  XfT  -  1  -  0,0555  -  0.9445 
y/js  +  y/T  -  1  ~  0.22(0.0555)  -  0.9^8 

Using  KfB  -  1.33,  KfT  *  0.533  (t  -  90°C.), 

1.332/s  -h  0.533s/r  -  0.988 

S/B  •"  0.609        and         X/T  =*  0.336 

By  plate-to-plate  calculation, 


xf 

*/ 

y/ 

3/+1 

C6 
CT 
C8 

0.609 
0.336 
0  0555 

1.33 
0.533 
0  22 

0  811 
0.179 
0  Oil 

0.718 
0.267 
0.016 

1  001 

For  benzene  relative  to  toluene 


0.995 


2(0.718)  _ 

" 


-. 
""  2(0.267) 

_  1  -  (39.9/180.3)  (0.005/0.609) 
/-I       i  _  (39.9/180.3)  (0.744/0.336) 


From  page  223, 


«  2.63 

=  2.36 
-  2.49 

~^ 
1.69 

2.49 


fa\          __  1. 

Wnav     " 


.47  -f  2.63 


-  2.05 


1.81  +  2.36 


1.84 


By  Eq.  (12-15), 
By  Eq.  (12-16) 


r    4-  l  -  lQg  (0.995/0.005)  (0.336/0.609)  ^  Q  6 
Ar        log  (0.609/0.336)  (0.744/0.005)       o  9 

J\  m    as r- — •    »   \J,6 


344  FRACTIONAL  DISTILLATION 

N  «•  Nn  +  Nm  =*  14.8  theoretical  plates  vs.  16  found  on  page  225  by  stepwise 
calculation.     Because  («/£)/  and  (a/£)/-i  are  near  to  1.0,  the  geometric  mean 
should  be  more  conservative. 
By  Eq.  (12-21), 


-  1  +  V0.47(1.63T  -  1.87 

\M/n  av 

and 

-  1  +  V0.31(1.36)  =  1.65 


Nn  +  I    -  7.5 

Nm  •  11.2 
Nn  +  Nm  =  17.7  theoretical  plates 

The  xylene  in  the  distillate  can  be  estimated  by  Eq.  (12-15). 
For  toluene  relative  to  xylene, 

j    ,       0.005 


2(0.267)    _1Q 


XDX 
).0555) 

O  /iQ      I     O  /4O 

-  2.45 

z  z 

i™  m  nn?i  x^^^/n  nKKK  /n  9 
7.5 


2  2 

log  (0.005 /gpy)  (0.0555/0.336) 

log  2.45 
»  0-005(0.0555)   ^  10^6 
^DX  "     830(0.336) 

An  approximate  check  on  the  assumed  feed-plate  temperature  can  be  obtained 
by  assuming  that  the  temperature  gradient  is  linear  from  the  still  to  the  condenser. 
From  pages  222  and  225,  tw  -  116°C.  and  to  =  80°C. 

7-5   (116  -  80)  +  80 


"       18.7 
-  94.4°C. 

as  compared  to  the  assumed  temperature  of  90°C. 

The  results  given  by  these  equations  are  only  approximate,  and 
their  accuracy  increases  as  the  reflux  ratio,  0/D,  increases.  For  reflux 
ratios  near  the  minimum  value,  /?/  becomes  equal  to  a,  and  the  equa- 
tions should  be  applied  with  caution  because  they  give  too  few  the- 
oretical plates  under  these  conditions.  In  fact,  owing  to  the  method 
of  obtaining  (a/£)av,  they  can  indicate  a  finite  number  of  plates  at 
values  of  the  reflux  ratio  less  than  the  true  minimum.  For  values  of 
(a/ ft)  near  to  1.0  a  better  average  is  obtained  by 


ALTERNATE  DESIGN  METHODS  345 


These  averages  force  the  equation  to  give  an  infinite  number  of 
plates  for  (a/  ft)  =  1.0. 

ABSORPTION  FACTOR  METHOD 

Brown  and  Souders  (Ref  .  3)  suggested  the  use  of  the  absorption  fac- 
tor method  of  Kremser  (Ref.  11)  as  a  design  procedure  for  multicom- 
ponent  mixtures. 

By  a  material  balance  on  one  component  starting  at  the  bottom  of 
the  column, 


V 


where  S  -  KV/0. 

V  (V  \ 

xs  =  -Q  2/2  -  r  -Q  yw  -  xil 


™  yw  -        2/TT  - 


and 


—  (  Q  2/TT  -  XlJ 


Assuming  S  is  a  constant  and  letting  Nm  equal  the  number  of  theo- 
retical plates  below  feed, 

\ 

{  +  s  4.  $2  +  .  .  .  +  ##,»)     (12-23) 

__  z 

+  (1  +  S  +  S*  +  •  •  •  +  S»*-i) 


346 


FRACTIONAL  DISTILLATION 


ir 

Al 


+ 


(1  +  8  + 


Xf  """""  X\ 

I        V* 


8(1  +  S  + 


1  +  S  +  S*  + 


(12-24) 


Multiplying  the  numerator  and  denominator  of  the  right-hand  side 
by  1  —  8  gives 

xf  -  xi       _  SNm+1  -  S 


xf  -   yw  »     - 

A  similar  analysis  above  the  feed  plate  gives 

y/-Vr    _  AN^  -  A 
Vf  - 


(12-25) 


(12-26) 


where  XR  *=  composition  of  reflux  to  top  of  tower 
A  =  0/KV 

Equation  (12-26)  is  applied  to  the  heavy  key  component  above  the 
feed  plate,  and  Eq.  (12-25)  to  the  light  key  component  below  the  feed 
plate. 

An  average  value  of  K  is  employed  and,  above  the  feed, 


K  av    = 


+ 


below  feed  plate, 


Absorption  Factor  Calculation  for  Benzene-Toluene-Xylene  Separation.  This 
method  will  be  applied  to  the  benzene-toluene-xylene  example  (data  from  pages 
220  and  343). 


»-«. 

Xr 

*,+, 

y/ 

3/TT 

^ 

Xf 

JTr 

K« 

c, 

—  . 



— 

— 

0.0131 

0.0116 

0.605 

2.63 

1.52 

c, 

0.005 

0.38 

0.48 

0.180 

—  — 

—  . 

— 

•—  " 

•  — 

ALTERNATE  DESIGN  METHODS  347 

Above  feed  plate  for  toluene, 

„         0.38  +  0.48 

/v.v  =  -  g  - 

A  °  2  1  „ 

A  *  TV  -  3(043)  =  IM* 
By  Eq.  (12-26), 

0.18  -  0.005  1.55***1  -  1.55 

0.18  -  0.38(0.005)  "     1.55"n+i  -  1 
Nn  -f  1  -  7.95 
Nn  -  6.95 

Below  feed  plate  for  benzene, 

_  2.63  +  1.52 

Aav    sss   -  S  -  -6.U/ 

5  -  2.07  ("%ao)  -  1.7 
By  Eq.  (12-25), 

0.605  -  0.0116  i.7*m+i  -  1.7 

0.605  -  (0.0131/2.63)         1.7^*»+l  -  1 
tf«  +  1  -  9.3 
#m  =  8.3 

Total  plates  «  7  +  8.3  +  1  -  16.3. 

Edminster  (Ref.  4)  has  presented  a  modified  absorption  factor 
method  that  determines  the  molal  quantities  for  each  component  as  a 
fraction  of  their  values  in  the  distillate  and  bottoms  in  a  manner  some- 
what similar  to  the  Thiele  and  Geddes  equations.  The  geometric 
mean  of  the  absorption  and  stripping  factors  at  the  ends  of  the  section 
under  consideration  is  employed,  and  empirical  correction  terms  are 
applied  to  these  averages. 

GRAPHICAL  CORRELATIONS 

When  the  number  of  theoretical  plates  is  plotted  as  a  function 
of  reflux  ratio,  the  curve  is  hyperbolic  in  type  with  asymptotes  at 
Nnn  and  (0/Z))min.  These  two  limiting  conditions  as  asymptotes  are 
useful  in  drawing  such  a  curve,  but  they  would  be  more  helpful  as  defi- 
nite points  on  the  diagram.  By  modifying  the  variables,  they  can  be 
made  definite  points;  in  fact  they  can  be  made  the  same  points  for  all 
cases.  There  are  many  ways  in  which  the  variables  can  be  modified, 
and  one  that  has  been  useful  (Ref.  6)  is  shown  in  Fig.  12-1.  The  ordi- 
nate  is  (S  —  Sm)/(S  +1),  where  S  is  the  total  theoretical  steps  includ- 
ing any  enrichment  in  the  still  and  condenser,  and  Sm  is  the  value  of  S 


for  total  reflux,  0/D  =00.     The  abscissa  is  |      -  + 


348 


FRACTIONAL  DISTILLATION 


At  total  reflux  the  ordinate  is  0.0  and  the  abscissa  is  1.0,  while  at  the 
minimum  reflux  ratio  the  ordinate  is  1.0  and  the  abscissa  is  0.0.  As 
the  reflux  ratio  is  increased  from  the  minimum  to  total  reflux,  a  given 
design  problem  will  give  a  curve  that  goes  from  1,0  to  0,1.  It  was 
expected  that  a  series  of  curves  between  these  two  points  would  be 
obtained,  depending  on  (1)  the  degree  of  separation,  (2)  the  relative 


!0r 


0.8 


0.6 


0.4 


02 


0.2 


^ 


0.4 


tO/ 


06 


08 


10 


f  • 

FIG.  12-1.     Graphical  correlation  for  design  calculations. 

volatilities,  and  (3)  the  components  lighter  and  heavier  than  the  key 
components.  The  results  of  plate-to-plate  calculations  were  plotted 
and  gave  a  narrow  band  which  could  be  reasonably  represented  by  a 
single  line.  It  can  be  shown  theoretically  that  a  single  line  cannot 
represent  all  cases  exactly,  and  the  correlation  can  be  improved  by 
using  more  than  one  line.  For  example,  the  position  of  the  line  is  a 
function  of  the  fraction  of  the  feed  that  is  vapor.  The  best  line  drawn 
through  the  all-vapor  feed  cases  on  a  plot  such  as  Fig.  12-1  is  lower 
than  the  corresponding  line  for  all-liquid  feeds.  It  is  also  possible  to 


ALTERNATE  DESIGN  METHODS 


349 


improve  the  correlation  by  changing  the  variable  groups,  but  it  is 
doubtful  whether  the  increased  accuracy  justifies  the  added  complica- 
tions. The  accuracy  of  such  a  correlation  will  always  be  limited  by 
the  errors  in  8m  and  (0/D)^  It  is  believed  that  it  is  of  real  value 
when  it  is  applied  as  (1)  a  rapid  but  approximate  method  for  pre- 
liminary design  calculation  or  (2)  a  guide  for  interpolating  and  extra- 


I.U 

08 
0.6 

04 
0.3 

02 

•»• 

C/5 

0.1 
008 

0.06 

0.04 
0.03 

0.02 
0 

1 

SSBMMM. 

—  «•• 

BBSS 

— 

-«— 

•—  1 

•«, 

—  « 

» 

—  • 

i= 

"^1 

""^1^^ 

"Si, 

X 

\ 

s 

. 

s 

s 

s 

\ 

V 

\ 

^ 

I 

\\ 

\\ 

V 

1 

1 

01             0,02     0.03  0.04      0.06  0.08010              02       03    0.4       0.6    OB  1. 

JB..IS 
0 


Fio.  12-2. 


WofO 

Graphical  correlation  for  design  calculation. 


polating  plate-to-plate  calculations.  In  this  latter  case,  if  only  one 
plate-to-plate  result  is  available  at  a  reflux  ratio  from  1.1  to  2.0  times 
(0/D)min,  this  point  can  be  plotted  on  the  diagram  and  a  curve  of  simi- 
lar shape  to  the  correlation  curve  fitted  to  it.  Such  a  method  should 
give  good  results  for  other  reflux  ratios,  assuming  the  values  of  Sm  and 
(0/D)min  are  reasonably  accurate. 


Use  of  Graphical  Correlation  for  Benzene-Toluene-Xylene  Separation. 

ing  this  correlation  to  the  benzene-toluene-xylene  example: 
From  page  259,  (0/D)mi»  -  1.0. 


Apply- 


350  FRACTIONAL  DISTILLATION 

Calculation  of  &»in.    From  Fig.  9-7,  page  233,  for  benzene  relative  to  toluene, 

as  «  2.6,        aw  *  2.3 

«av  -  V2^(2T)  -  2.45 


0.00fi/\0.005     _       . 
.4 


(0/D)  -  (0/D)m      2.0  -  1.0 

(0/D)  +  1         "  2.0  4-  1.0  -  °'333 

From  Fig.  12-1, 

^~Y  -  0.32 

8  -  17.2 

N  «  16.2  theoretical  plates 

This  compares  with  16  plates  as  determined  by  the  stepwise  procedure. 

An  interesting  fact  is  that  the  commonly  used  design  reflux  ratios  of 
1.2  to  1.5  times  (0/D)  mi*  usually  correspond  to  values  of  (S  —  Sm)/ 
(S  +  1)  from  0.4  to  0.6,  and  because  S  is  usually  large  in  comparison 
to  1.0,  a  rough  working  rule  is  that  the  number  of  theoretical  steps 
required  at  the  optimum  reflux  ratio  will  be  twice  the  number  needed 
at  total  reflux.  This  is  a  rule  that  can  be  applied  for  orientation  pur- 
poses and  requires  only  an  estimation  of  the  number  of  theoretical 
plates  at  total  reflux. 

The  correlation  also  illustrates  why  the  economic  reflux  ratio  is  usu- 
ally so  close  to  (0/Z))min.  A  small  initial  increase  in  the  reflux  ratio 
group  above  0.0  makes  a  large  decrease  in  the  theoretical  plate  group, 
but  further  increases  become  less  effective. 

A  modification  of  the  above  correlation  has  been  suggested  by 
Schiebel  (Ref.  12),  and  a  somewhat  different  graphical  correlation  has 
been  published  by  Brown  and  Martin  (Ref.  2). 

MODIFIED  EQUILIBRIUM  CURVES  AND  OPERATING  LINES 
Several  methods  have  been  given  which  treat  a  multicomponent 
mixture  as  a  modified  binary  mixture  of  the  key  components  that  can 
be  analyzed  graphically  on  a  y,x  type  diagram. 

Jenny  (Ref.  10)  published  a  graphical  method  for  multicomponent 
design  calculation.  A  few  plate-to-plate  calculations  are  made  at  the 
top  and  bottom  of  the  column  and  above  and  below  the  feed  plate. 
For  the  section  below  the  feed  plate,  a  yfx  diagram  is  made  for  the  light 
key  component  using  the  calculated  values  to  place  the  effective  equi- 
librium curve  for  this  component  on  the  diagram.  The  operating  line 
is  drawn  in  the  usual  manner  and  the  plates  determined  by  the  stepwise 


ALTERNATE  DESIGN  METHODS  351 

procedure.    Above  the  feed  plate  a  diagram  is  constructed  for  the 
heavy  key  component  in  the  same  manner. 

Hengstebeck  (Ref.  8)  developed  a  graphical  method  for  multicom- 
ponent mixtures  .which  employed  the  key  components  on  a  binary- 
type  diagram.  The  mol  fractions  for  the  distillate  and  bottoms  were 
calculated  on  a  key  component  basis 

r/     _        %wik 

xwlk  — 


XWlk  +  Xwhk 

and  the  equilibrium  curve  for  a  binary  mixture  of  the  key  components 
is  employed.  The  operating  lines  are  corrected  for  the  quantity  of  the 
light  and  heavy  components  present. 

Schiebel  and  Montross  (Ref.  13)  developed  a  method  for  making 
multicomponent  design  calculations  on  the  basis  of  a  pseudobinary 
mixture.  A  modification  of  this  method  by  Bailey  and  Coates  (Ref.  1) 
will  be  reviewed. 

The  analysis  is  made  on  a  key  component  basis,  and  plate-to-plate 
calculations  are  made  above  and  below  the  feed  plate  to  determine 
xf+i>  £/+i>  2//>  x'f,  if)  and  2//_!.  The  feed-plate  composition  is  obtained 
by  the  method  given  on  page  342.  The  operating  line  in  the  stripping 


section  is  drawn  as  a  straight  line  through  y'wlk  =  x'wik  = —^ and 

Xwlk  +  Xwhk 

i            x/ik         t                  y(f—i)ik  j    •    M  j 

Xflk  — 9  2/</_i)Zfc  = i  9  an(*  similar  procedures 

are  used  to  calculate  x'Dik,  y'ftk,  and  x[f+l)ik,  where  the  primed  mol  frac- 
tions are  on  a  key  component  basis.  The  relative  volatility  of  the  key 
components  is  plotted  vs.  xr  using  the  values  at,  x'D  and  a/+i,  x'f+i  to 
give  a  curve  for  volatility  above  the  feed  plate  and  aw,  x'w  and  a/,  x'f 
to  give  a  curve  below  the  feed  plate.  These  relative  volatility  curves 
are  used  to  calculate  the  yf  vs.  x'  equilibrium  curves  by 

v 


(a  -  IX  +  1 

In  general,  the  equilibrium  curves  calculated  from  the  two  curves 
do  not  match  exactly,  and  each  is  used  for  its  respective  portion  of  the 
column.  Plates  are  stepped  off  between  the  operating  lines  and  equi- 
librium curves  in  the  usual  manner. 

Bailey  and  Coates  (Ref.  1)  also  modified  the  Schiebel  and  Montross 
method  of  calculating  the  minimum  reflux  ratio  for  a  multicomponent 


352  FRACTIONAL  DISTILLATION 

mixture.     They  give 


+ 


<*"> 


|W«r+  «u)J 


(12-27) 


where  (0/D)'M  is  the  pseudo-minimum  reflux  ratio. 

This  equation  appears  somewhat  similar  to  Eq.  (9-21),  and  the 
nomenclature  is  the  same.     (0/D)fM  is  calculated  by 


0 


(1228} 

(       ] 


where  x'D, 


=  mol  fractions  on  key  component  basis  for  distillate  and 

at  intersection  of  pseudo  operating  lines 
=  average  relative  volatility 


m)[(0/D)'M 


-  x'D 


(0/D)'U 


(12-29) 


where  m  == 


ML  - 


Mv  - 


Mv 


=  mol  fraction  on  key  component  basis  for  feed 

=  mols  of  liquid  in  feed 

=  mols  of  vapor  in  feed 

=  mols  of  components  heavier  than  heavy  key  component  in 

feed 
=  mols  of  components  lighter  than  light  key  component  in 

feed 

Values  of  x'D  are  calculated  from  the  terminal  conditions,  and  x[  and 
(0/D)'M  obtained  by  simultaneous  solution  of  Eqs.  (12-28)  and  (12-29). 
The  calculation  of  (0/D)'M  is  frequently  laborious,  and  Eq.  (9-21)  is 
easier  to  use. 

Solution  of  Benzene-Toluene-Xylene  Example  by  Modified  Equilibrium  Curve 
Method.  The  use  of  this  method  will  be  illustrated  by  the  benzene-toluene-xylene 
example.  Benzene  and  toluene  are  the  key  components  and  the  necessary  feed- 
plate  calculations  were  made  fox  this  example  on  page  343. 

The  results  are  summarized  for  the  key  components: 


Xf 

Vf 

s/+i 

y/-i 

C6 
C7 

0  609 
0.336 

0.811 
0.179 

0.718 
0.267 

0.739 
0.246 

-  90°C. 


/-i  -  95°C. 


ALTERNATE  DESIGN  METHODS 


353 


f.o 


0,8 


8. 
5 

c 

i 

I  0.6 
.8 

c 
o 

u 

I 

I  04- 


0.2 


A 


02  04  06  08 

x'-pseuclo  mol  fraction  of  benzene  in  liquid 

FIG.  12-3. 


10 


Z.O 

i" 

Q> 
£ 

2.0, 

/ 

^*- 

^ 

fr      ii** 

.^  -"• 

^—  — 

.^  —  —  - 

^  —  — 

^^ 

)            02           0.4           0.6           0,8           I. 
x',  pseudo  mol  fraction 

FIG.  12-4. 

354  FRACTIONAL  DISTILLATION 

Coordinates  for  benzene  for  lower  operating  line, 

0-°°67 


,          0.739         _ 
'/-'  '  0985  =  °'75 

'       °'609      n  <u* 
*'  ~  0945  -  a645 


Coordinate  for  enriching  line, 

/          .        0.995 


0.995 


0.718         7 
0985  "  °'73 

Straight  lines  are  drawn  through  these  points  on  Fig.  12-3.  The  relative  volatil- 
ity values  corresponding  to  at,  «/+i,  «/,  and  aw  were  plotted  as  a  function  of  xf  in 
Fig.  12-4.  These  values  were  used  to  calculate  the  equilibrium  curve  of  Fig.  12-3. 
The  steps  are  made  on  this  diagram  in  the  usual  manner,  and  the  procedure  gives 
between  16  and  17  theoretical  plates.  This  method  is  particularly  good  in  this 
case  because  the  components  other  than  the  key  components  are  present  only  in 
small  quantities. 

The  minimum  reflux  ratio  by  this  method  is  obtained  as  follows.    By  Eq.  (12-28), 


M 

By  Eq.  (12-29), 


m~— o °° 


100  -  10 
m  =» 

,        0.6 

Xj,    » 

By  Eq.  (12-27), 


,    --0*7        and 


,1Q1    ,  3MF    0.4(01)    1 
n  ^  60.6  [2.34  -  0.4  J 

-  1.02 

This  compares  closely  with  the  values  given  on  page  259. 

ANALYTICAL  EQUATIONS 

Exact  mathematical  equations  for  the  case  of  constant  molal  over- 
flow rate  and  constant  relative  volatilities  have  been  presented  by 
Harbert  (Ref.  7)  and  Underwood  (Ref.  16).  Underwood's  equation 
for  a  three-component  mixture  can  be  arranged  as  follows: 


ALTERNATE  DESIGN  METHODS  355 

Above  feed  plate, 


XDI  _  [  01  ]  *     XFI 
Xpt        \02/          Xpt 


Xp9  _  I  03  \  "      ^Ty8 
Xpl        \0i/          XFl 


Similar  ratio  equations  can  be  written  for  any  two  components  of  a 
multicomponent  mixture  where 


i  B         i  cpC         . 

—  — 


OLA  —  0i 
XD  =  distillate  composition 
XD»  XDa  =  XDl  with  0i  replaced  by  02  and  03,  respectively 

XF  =  ^D  with  XDA,  XDBK  XDC  replaced  by  XFA,  XFB,  and  XFC, 

respectively 

XF  =  feed-plate  composition 
Nn  =  number  of  plates  above  feed  plate 
0i,  02,  03  =  roots  of  the  equation 


D  D  D 

-  XDA        Ois  -r-  XDB        Oic  --  XDC 


OLA    —    0  OJJ3    —    0  «C    —    0 


(12-31) 


The  roots  will  be  bracketed  by  the  relative  volatilities;  thus,  for  a 
three-component  mixture, 

CiA   >    01   >   OtB,  OiB   >    02   >   «C,  <*C   >    02   >   0 

Below  the  feed  plate  similar  equations  are  obtained. 


~w~  ~~  i  ~77  /      V'  (12-32) 

AFa          \02/        ^TF, 

Similar  relations  can  be  written  for  the  other  components  where 

X'F  —  XF  except  0'  used  instead  of  0 

X'w  =  Xp  with  XDA,  XPB,  XDC  replaced  by  XWA,  XWB,  Xwc,  respec- 
tively, and  using  0'  instead  of  0 

Nm  —  number  of  theoretical  plates,  feed  plate  and  below 
0i>  02>  03  =  roots  of  the  equation 

WWW 
otA  TV-  XWA       OLB  fr-  XWB       etc  TF~"  Xwc 


356  FRACTIONAL  DISTILLATION 

In  many  cases,  the  use  of  these  equations  is  complicated  by  the  fact 
that  a  trial-and-error  procedure  is  involved.  If  the  terminal  composi- 
tions are  known,  the  trial-and-error  operation  involves  matching  at 
the  feed  plate.  Usually  the  terminal  conditions  are  not  completely 
known,  and  additional  trial  and  error  may  be  required.  In  most  cases 
a  three-component  problem  can  be  solved  just  as  rapidly  by  the  usual 
stepwise  procedure,  and  variations  in  the  relative  volatility  can  be 
included. 

Underwood  has  used  these  equations  to  calculate  the  minimum 
reflux  ratio. 

t  +  lss  sfzrg  (12"34> 

where  0  is  the  root  of  the  equation. 

_j _| _|_  .  .  .   a-  1  -J-  p          (12-35) 

CtA   —    V  &B   —    0  OLc   —    0 

where  p  =  (0»  -  Om)/F. 

This  equation  has  several  roots.  The  one  employed  in  Eq.  (12-34) 
is  the  value  lying  between  aik  and  othk- 

Solution  of  Benzene-Toluene-Xylene  Example  by  Analytical  Equations.     These 
equations  will  be  applied  to  the  benzene-toluene-xylene  example.     Relative  vola- 
tilities for  the  three  components  will  be  used  as  as  -  2.45,  <XT  ~  1.0,  and  ax  =0.4. 
By  Eq.  (12-31), 

2.45(0.995/3)    ,  0.005/3  ^  . 

2.45  -  0      +  1  -  0 
2.45  >  0i  >  1.0,         1.0  >  02  >  0 

The  term  for  xylene  is  neglected  because  its  concentration  in  the  distillate  is 
unknown,  and  it  will  make  little  difference  in  the  values  of  0i  and  02.  Solution 
of  this  equation  gives 

0i  -  1.64,         02  =  0.9963 
By  Eq.  (12-33), 

2.45 (0.005)  (39.9/180.3)       1.0 (0.744)  (39.9/180.3)       0.4(0.251)  (39.9/180.3)  _      . 
2.45  -  0'  1-0'  0.4-0' 

0J  >  2.45,        2.45  >  02  >  1.0,         1.0  >  0J  >  0,4 
0'8  -  0.4173,        02  -  1.169,        0J  «  2.45308 

Using  these  values  for  the  key  components, 


2.45  -  2.45308  "  1  -  2.45308  ""  0.4  -  2.45308 
-  0.1952** 


ALTERNATE  DESIGN  METHODS  357 

It  is  obvious  that  the  first  term  will  be  the  important  factor. 

X'Wi  «  -795(0.005)  -  0.688(0.744)  -  0.195(0.251) 

1  =  -4.64 

v'    _       2.45xfa        ,         XFT        ,     ,  0.4^^ 
AF*  ~  2.45  -  1.169  "^  1  -  1.169  "*"  0.4  -  1.169 

«  I.QI&XPB  —  5.92a*r  -  QM9xFx 
X'w%  -  -4.53 

By  Eq.  (12-32), 

2.45  \y-  /-4.64\ 
l-169/      V-4.53/ 


-  Q.G88XFT  -  0. 


By  a  similar  procedure  for  XFl  and  XF8, 

/  1.169  V^  /  -4.53  \ 
\0.4173/      V-4.514/ 


-f  1. 

With  XFB  +  XFT  +  ZFX  =  1.0,  there  are  three  equations  and  four  unknowns. 
By  using  two  similar  equations  for  the  section  above  the  feed,  only  one  additional 
unknown,  Nn,  is  introduced,  and  the  equation  can  be  solved.  However,  with 
XDX  unknown,  only  one  equation  is  available.  XDX  is  small,  and  to  satisfy  Eq. 
(12-31)  it  can  be  shown  that  <fo  =  0.4  —  0.334#z>;x-,  and  the  correction  term 
—  0.334o?i>jr  can  be  neglected  in  all  cases  except  the  (<£3  —  0.4)  term.  This  intro- 
duces an  additional  unknown,  and  an  additional  specification  can  be  added  such 
as  the  ratio  of  the  key  components  on  the  feed  plate  or  that  Nn  +  Nm  is  to  be  a 
minimum.  The  trial-and-error  solution  for  the  latter  case  is  time-consuming; 
instead,  the  feed-plate  compositions  obtained  by  the  stepwise  procedure  on  page 
223  will  be  employed;  XPB  =•  0.605,  XPT  =  0.336,  and  XPX  *  0.058. 

For  ratio  of  1  to  2,  the  relation  reduces  to 


-4.53 


/   481    \ 
V-0.86/ 


1.169/  -4.64 

Nm  =  8.5  theoretical  plates 
For  ratio  of  2  to  3, 

/  1.169  YM  _  4.514  /  0.86  \ 
V0.4173/  4.53  V0.0447 

Nm  -  2.9 

The  last  value  for  Nm  is  very  sensitive  to  the  value  employed  for  XFX,  and  a 
value  of  0.0562  for  xylene  at  the  feed  plate  instead  of  0.058  would  make  the  last 
equation  give  Nm  —  8.5.  This  value  of  Nm  «  8.5  compares  with  nine  plates  by 
the  stepwise  procedure.  The  difference  is  due  to  the  fact  that  the  relative  vola- 
tility in  the  lower  section  of  the  tower  averaged  less  than  2.45. 

A  similar  calculation  for  the  upper  section  using  the  same  feed  composition  gave 
Nn  +  1  -  8.6,  Nn  -  7.6  or  total  plates, 

N  -  Nn  -f  Nm  -  8.5  4-  7.6  *  16.1 
as  compared  to  16  plates  by  the  Lewis  and  Matheson  method. 


358  FRACTIONAL  DISTILLATION 

For  the  minimum  reflux  ratio,  by  Eq.  (12-33), 

2.45(0.6)       1(0.3)       0.4(0.1) 
2.45  -  0  ~l"  1^1  "*"  0.4  -  0  *  U 


The  desired  $  is  between  1  and  2.45. 


e  -  1.25 

2.45 


a   '  2.45  -  1.25 

=  2,02 

(£)-.  -  ^ 

This  compares  closely  with  the  values  given  on  pages  259  and  354. 

SUMMARY 

The  Lewis  and  Matheson  method  appears  to  be  the  most  satisfac- 
tory method  of  handling  the  general  multicomponent  design  problem. 
In  most  cases,  it  requires  relatively  little  trial  and  error  and  will  handle 
cases  of  normal  and  abnormal  vapor-liquid  equilibria.  It  is  especially 
well  suited  to  the  cases  in  which  the  reflux  ratio  and  the  separation  of 
the  key  components  are  specified  and  the  problem  is  to  determine  the 
number  of  theoretical  plates  and  the  component  concentrations  in  the 
column.  In  some  specific  cases,  other  methods  may  have  advantages, 
but  unless  a  number  of  problems  of  the  same  type  are  to  be  handled,  it 
is  more  desirable  to  have  one  method  that  will  apply  to  essentially  all 
cases. 

The  Thiele  and  Geddes  method  is  advantageous  when  the  number 
of  theoretical  plates  and  the  reflux  ratio  are  specified  and  the  calcula- 
tion of  the  separation  is  desired.  Even  in  this  case,  the  trial  and  error 
involved  in  obtaining  the  proper  equilibrium  constants  for  each  plate  is 
formidable. 

The  reduced  relative  volatility  and  absorptipn  factor  methods  are 
rapid  but  can  be  appreciably  in  error  because  of  the  approximations 
involved.  If  the  design  engineer  understands  their  limitations,  they 
can  be  useful.  In  cases  involving  abnormal  vapor-liquid  equilibrium 
conditions  or  at  reflux  ratios  near  to  the  minimum,  these  methods  may 
be  so  in  error  that  the  results  are  of  little  value. 

The  methods  based  on  modified  equilibrium  curves  and  operating 
lines  are  simple  to  use  and  can  give  satisfactory  results  for  many  cases. 
They  do  not  give  information  on  the  light  and  heavy  components 
without  considerable  additional  effort.  In  general,  they  are  not  satis- 
factory for  mixtures  with  abnormal  vapor-liquid  equilibria. 


ALTERNATE  DESIGN  METHODS  359 

The  analytical  solutions  for  the  case  of  constant  molal  overflow  and 
constant  relative  volatility  are  a  mathematical  accomplishment,  but 
they  do  not  appear  to  be  well  suited  for  the  average  design  problem. 
The  greatest  contribution  of  these  methods  will  probably  be  as  an  aid 
in  studying  the  process  of  rectification,  particularly  the  minimum  reflux 
ratio  condition. 

The  graphical  correlations  are  useful  in  obtaining  approximate 
answers  rapidly.  They  are  not  applicable  to  all  cases  and  can  be 
seriously  in  error  under  some  conditions. 

The  various  approximate  methods  are  helpful  for  orientation  pur- 
poses, but  the  greater  confidence  that  the  design  engineer  can  place  in 
the  rigorous  plate-to-plate  calculation  justifies  any  greater  effort 
involved. 

Nomenclature 

A  =  absorption  factor  -  On/KVn 
D  —  molal  distillate  rate 
F  «•  molal  feed  rate 
K  =  equilibrium  constant  =»  y/x 
m  —  ratio,  see  page  352 
ML  =  mols  of  liquid  in  feed 
My  =  mols  of  vapor  in  feed 

^MHF  =  mols  of  components  heavier  than  heavy  key  component  in  feed 
XMiF  —  mols  of  components  lighter  than  light  key  component  in  feed 
Nn  =  number  of  theoretical  plates  above  feed  plate 
Nm  =  number  of  theoretical  plates  feed  plate  and  below 
0  =  molal  overflow  rate 
8  «*  stripping  factor  «  KVm/Om 
t  •»  temperature 
V  =  molal  vapor  rate 
W  =  molal  bottoms  rate 
X  =  mol  fraction  in  liquid 
X  «  composition  factor  in  Eq.  (12-30) 
a/  =  mol  fraction  in  liquid  based  on  key  components  only 
y  «=  mol  fraction  in  vapor 

yf  =  mol  fraction  in  vapor  based  on  key  components  only 
a  «  relative  volatility 
0  =»  relative  operability 
<M'  -  roots  of  Eqs.  (12-32)  and  (12-33) 
0  «  root  of  Eq.  (12-35) 

Subscripts: 

A  refers  to  component  A 
B  refers  to  benzene  or  component  B 
Df  refers  to  distillate 
F  refers  to  feed 


360  FRACTIONAL  DISTILLATION 

f  refers  to  feed  plate 

h  refers  to  heavy  component 
hk  refers  to  heavy  key  component 
I  refers  to  light  component 

Ik  refers  to  light  key  component 

n  refers  to  section  above  feed  plate 

m  refers  to  section  below  feed  plate 

T  refers  to  toluene  or  to  top  plate 
W  refers  to  bottoms 

X  refers  to  xylene 

References 

1.  BAILEY  and  COATES,  Petroleum  Refiner,  27,  30,  87  (1948). 

2.  BROWN  and  MARTIN,  Trans.  Am.  Inst.  Chem.  Engrs.,  35,  679  (1939). 

3.  BROWN  and  SOUDERS,  Trans.  Am.  Inst.  Chem.  Engrs.,  30,  438  (1933). 

4.  EDMINSTER,  Chem.  Eng.  Progress,  44,  615  (1948). 

5.  GILLILAND,  Ind.  Eng.  Chem.,  27,  260  (1935). 

6.  GILLILAND,  Ind.  Eng.  Chem.,  32,  1220  (1940). 

7.  HARBERT,  Ind.  Eng.  Chem.,  37,  1162  (1945). 

8.  HENGSTEBECK,  Trans.  Am.  Inst.  Chem.  Engrs.,  42,  309  (1946). 

9.  HUMMEL,  Trans.  Am.  Inst.  Chem.  Engrs.,  40,  445  (1944). 

10.  JENNY,  Trans.  Am.  Inst.  Chem.  Engrs.,  36,  635  (1939), 

11.  KREMSER,  Nat.  Petroleum  News,  22,  No.  21,  48  (1930). 

12.  SCHIEBEL,  Ind.  Eng.  Chem.,  38,  397  (1946). 

13.  SCHIEBEL  and  MONTROSS,  Ind.  Eng.  Chem.,  38,  268  (1946). 

14.  THIELE  and  GEDDES,  Ind.  Eng.  Chem.,  25,  289  (1933). 

15.  UNDERWOOD,  J.  Soc.  Chem.  Ind.,  52,  223  (1933). 

16.  UNDERWOOD,  J.  Inst.  Petroleum,  32,  614  (1946). 


CHAPTER  13 
SIMULTANEOUS  RECTIFICATION  AND  CHEMICAL  REACTION 

In  most  rectification  systems,  chemical  reactions  among  the  com- 
ponents to  form  new  stable  compounds  do  not  usually  occur,  but  in  a 
few  special  cases  such  a  condition  is  involved.  In  some  cases,  the 
combination  of  rectification  and  reaction  is  beneficial  as  in  the  prepara- 
tion of  esters;  while  in  other  cases,  it  may  be  detrimental  in  that  it 
decreases  the  effectiveness  of  the  separation  or  the  yield  of  the  desired 
component. 

In  the  preparation  of  esters  such  as  ethyl  acetate  from  acetic  acid  and 
ethanol,  the  equilibrium  is  such  that  only  a  moderate  conversion  is 
obtained.  If  the  reaction  mixture  is  brought  to  equilibrium  and  then 
fractionally  distilled  to  remove  the  ethyl  acetate,  leaving  water,  unre- 
acted  acetic  acid,  and  ethanol,  it  would  give  only  a  low  conversion  on 
further  reaction.  The  reaction  can  be  carried  out  in  a  fractionating 
column  with  the  esterification  occurring  in  the  liquid  on  the  plates. 
The  acetic  acid  is  added  to  the  upper  portion  of  the  column,  and  the 
alcohol  is  introduced  in  the  lower  portion.  The  plates  below  the 
alcohol  addition  are  used  to  strip  the  alcohol  out  of  water.  The  middle 
section  between  the  plates  where  the  alcohol  and  acetic  acid  are  added 
is  the  chief  esterification  section.  The  upper  portion  of  this  middle 
section  has  a  high  ratio  of  acetic  acid  to  alcohol  and  gives  good  cleanup 
of  the  ethanol.  The  lower  portion  is  high  in  alcohol  and  gives  a  rapid 
reaction  of  the  acetic  acid.  The  top  part  of  the  column  fractionates 
the  ethyl  acetate  out  of  the  acetic  acid.  A  small  amount  of  acid 
catalyst  is  added  at  the  top  of  the  column.  The  unit  produces  the 
ethyl  acetate  and  water  and,  by  separating  them,  carries  the  reaction 
almost  to  completion. 

In  the  esterification  system,  if  the  liquid  level  on  the  plates  is  low 
and  a  normal  vapor  rate  is  employed,  the  amount  of  reaction  per  unit 
time  will  be  small  relative  to  the  vapor  rate  and  a  high  reflux  ratio  will 
be  necessary.  This  will  result  in  a  high  heat  consumption  per  gallon 
of  ethyl  acetate  produced.  If  a  very  deep  liquid  level  is  employed  on 
the  plate  to  increase  the  amount  of  reaction  relative  to  the  vapor  rate, 
it  may  be  that  the  rate  at  which  the  ethyl  acetate  is  removed  is  so  low 
that  the  liquid  on  the  plates  is  near  equilibrium  and  the  reaction  rate 

361 


362  FRACTIONAL  DISTILLATION 

will  be  retarded.  Thus  the  large  volume  of  column  required  because 
of  the  deep  liquid  level  will  be  used  ineffectively.  The  volume  of  the 
liquid  should  be  proportioned  to  the  vapor  rate  such  that  there  is  an 
adequate  supply  of  ester  for  separation  but  not  to  the  extent  that  it 
seriously  retards  further  reaction. 

The  large  liquid  volume  required  can  be  obtained  by  using  a  deep 
liquid  level  on  the  plates  or  by  having  the  overflow  from  a  plate  pass 
through  a  holding  tank  for  the  chemical  reaction  before  it  is  added  to 
the  plate  below.  Deep  liquid  levels  give  undesirable  action  with  nor- 
mal-type bubble  plates,  but  tall  caps  and  risers  can  be  used  to  obtain 
satisfactory  operation. 

In  other  cases,  the  chemical  reaction  may  be  undesirable.  For 
example,  in  the  distillation  of  an  aqueous  solution  of  alcohols  and  alde- 
hydes, aldols  can  form  in  the  upper  portion  of  the  column  and,  being 
of  lower  volatility,  go  down  the  column  and  be  hydrolyzed  by  the 
water  at  the  bottom.  This  cycling  action  is  undesirable.  Polymeri- 
zation and  thermal  decomposition  occur  in  some  cases  and  are  usually 
objectionable.  In  most  of  these  cases,  the  undesirable  reactions  take 
place  in  the  liquid  phase  and  can  be  minimized  by  using  low  liquid 
depth  on  the  plates  and  by  keeping  the  temperatures  as  low  as  practi- 
cal. Frequently  inhibitors  can  be  added  to  the  fractionating  system 
to  reduce  the  amount  of  reaction.  For  example,  in  the  f ractionation  of 
styrene  from  ethyl  benzene  or  of  butadiene  from  butylene,  the  poly- 
merization of  the  styrene  and  butadiene  can  be  reduced  by  adding 
inhibitors  such  as  sulfur  or  tertiary  butyl  catechol  with  the  reflux. 
The  styrene  rectification  is  also  carried  out  under  vacuum  to  reduce 
the  temperature.  The  separation  of  styrene  from  ethyl  benzene 
requires  a  large  number  of  plates,  and  even  with  specially  designed 
bubble-cap  plates  the  pressure  drop  would  be  so  great  that  the  still 
temperature  would  be  excessive.  To  avoid  this  difficulty  the  rectify- 
ing column  is  made  in  two  sections.  The  vapor  from  the  section  that 
serves  as  the  bottom  portion  of  the  tower  is  liquefied  under  vacuum  in 
a  condenser  and  then  pumped,  vaporized,  and  added  at  the  bottom  of 
the  other  section.  The  liquid  from  the  bottom  of  the  top  section  is 
added  to  the  top  of  the  lower  section.  A  vacuum  condenser  is  also 
employed  for  the  top  section.  In  this  manner  the  average  pressure 
and  temperature  in  the  column  can  be  made  lower  than  for  the  single 
tower. 

The  design  calculations  for  these  systems  can  be  made  by  the  usual 
stepwise  procedure  making  allowance  for  the  chemical  reaction  on  each 
plate.  The  calculations  frequently  require  trial-and-error  methods, 


SIMULTANEOUS  RECTIFICATION  AND  CHEMICAL  REACTION    363 

because  the  terminal  conditions  and  flow  rates  are  a  function  of  the 
chemical  reaction.  The  method  of  calculation  will  be  illustrated  by 
the  fractional  distillation  of  cyclopentadiene. 

Example  on  Preparation  of  Cyclopentadiene.  Cyclopentadiene  is  obtained  in 
high-temperature  vapor-phase  petroleum  cracking  operations.  It  is  mixed  with 
other  hydrocarbons,  and  its  separation  is  complicated  by  the  fact  that  it  will 
dimerize  and  the  dimer  will  depolymerize  at  about  normal  distillation  conditions. 
One  method  of  operating  is  to  dimerize  the  pentadiene  and  then  remove  all  remain- 
ing constituents  below  Cr.  The  residue  is  then  given  a  thermal  treatment  which 
will  depolymerize  the  dimer,  and  the  mixture  is  distilled  to  obtain  the  cyclo- 
pentadiene. A  unit  is  to  be  designed  for  this  final  fractional  distillation,  and  an 
estimate  is  to  be  made  of  the  amount  of  polymerization  that  will  be  obtained. 

As  a  basis  for  the  estimate,  it  is  assumed  that  the  feed  is  a  binary  mixture  con- 
taining 20  mol  per  cent  cyclopentadiene  and  80  mol  per  cent  C7.  The  overhead 
product  is  to  contain  98  mol  per  cent  cyclopentadiene,  and  the  cyclopentadiene 
content  of  the  bottoms  is  to  be  0.5  mol  per  cent.  The  column  will  operate  with  a 
still  and  total  condenser  at  atmospheric  pressure.  A  reflux  ratio,  0/D,  three  times 
the  minimum  reflux  ratio  for  the  separation  of  the  binary  mixture  with  no  poly- 
merization will  be  used. 

Data  and  Notes 

1.  Holdup  in  the  condenser  is  negligible. 

2.  Holdup  per  plate  is  equivalent  to  a  liquid  depth  of  3  in.  on  the  superficial 
area. 

3.  Holdup  in  the  still  is  equivalent  to  one  plate. 

4.  Superficial  vapor  velocity  is  equivalent  to  1.0  f.p.s.  (S.T.P.) 

5.  The  liquid  on  each  plate  is  well  mixed,  and  the  plate  efficiency  is  100  per  cent. 

6.  The  feed  enters  such  that  Vn  »  Vm. 

7.  Neglect  polymerization  in  vapor. 

8.  For  simplification  assume  the  relative  volatilities  of  cyclopentadiene  and 
dicyclopentadiene  to  C7  are  6.0  and  0.1,  respectively. 

9.  The  rate  of  depolymerization  at  the  temperature  of  distillation  is  negligible. 

10.  The  rate  of  polymerization  in  the  liquid  phase  is  a  function  of  temperature, 
but  for  simplification  an  average  value  of  the  rate  constant  will  be  employed  in  the 
equation  —  (dr/d6)  =  2£V2,  where  r  —  g.  mol  of  cyclopentadiene  per  liter,  0  «  sec., 
K  -  6  X  10-fil./sec.-g.  mol. 

11.  Assume  that  liquid  volumes  are  additive.     Densities  of  CB,  CT,  and  Cio 
average  0.80,  0.68,  and  0.92,  respectively. 

12.  Assume  that  the  same  number  of  mols  of  vapor  enters  each  plate  per  unit  of 
time. 

Solution.  Minimum  reflux  ratio  for  binary  mixture:  Because  cyclopentadiene 
is  the  more  volatile  component,  the  mol  fractions  for  the  binary  mixture  will  be 
used  on  the  basis  of  this  component.  With  Vn  «•  Vm,  the  intersection  of  the 
operating  lines  will  be  at  x  ~  XF  «  0.2. 

With  a  constant  relative  volatility  of  6.0,  the  pinched-in  condition  will  occur  at 
the  intersection  of  the  operating  lines.  The  y  coordinate  of  the  intersection  is 

F  °*  ___  6(0.20) 


1  +  («  -  D*       1  +  (6  -  1)(0.20) 


364  FRACTIONAL  DISTILLATION 

From  these  coordinates  and  because  XD  **  0.98,  the  slope  of  the  enriching  line 
and  the  reflux  ratio  are  found. 

0\  0.98  -  0.60      0.38 

F/min        0.98  -  0.20  "  0.78 

°'38  0.95 


<D/min.       0.78  -  0.38 
Design  case: 

For  this  case,  O/D  »  3(O/D)mm. 

~  -  2.85        and        ~  =  3.85 

Because  the  amount  of  polymerization  is  unknown  and  because  the  complete 
composition  of  neither  the  distillate  nor  the  bottoms  is  known,  the  calculations  are 
a  series  of  trial-and-error  calculations. 

Basis:  1  sq.  ft.  of  plate  area  and  1  sec.  of  time. 

The  volume  of  liquid  on  a  plate  *  J£  cu.  ft.  =  Y± (28.32)  «  7.08  1. 

The  molal  volumes  of  each  of  the  three  components  may  be  found  by  dividing 
their  respective  molecular  weights  by  their  densities. 


C6: 
C7: 

Cio: 

v  -  66.1/0.80  -  0.0826  l./g.  mol' 
v  -  100.2/0.68  -  0.1475  l./g.  mol 
v  -  132.2/0.92  -  0.144  l./g.  mol 
1 

0.0826z8  +  0.1475*7  +  0.144zi0       0.0826 


or  the  number  of  mols  of  Cs  that  polymerize  on  any  plate  in  1  sec.  is  given  by 
p  -  6(7.08)10~B  r*  -  (4.25)10-4  r* 

From  the  superficial  vapor  rate  of  1  f  .p.s.  at  standard  temperature  and  pressure, 
the  conversion  factor  of  454  g./lb.,  and  the  fact  that  the  pound  molal  volume  of  a 
gas  is  359  cu.  ft.  (S.T.P.),  the  vapor  rate  V  may  be  found. 

V  =  |g  -  1.265  g.  mob/sec. 

V  1  9fi^ 

*          D-  775  -iS-  0-828  g-moh/aeo. 

0R  _  7  -  D  -  1.265  -  0.328  -  0.937  g.  mols/sec. 

Top  plate: 

Because  the  conditions  at  the  top  of  the  tower  are  much  better  known  than  are 
those  at  the  bottom  of  the  tower,  the  calculations  will  be  started  at  the  top.  The 
volatility  of  Cio  is  so  low  that  it  is  assumed  first  that  the  Cio  in  the  distillate  is 
negligible. 

To  calculate  the  amount  of  polymerization  on  the  top  plate,  the  ratios  XT/X&  and 
XM/XS  must  be  known. 


SIMULTANEOUS  RECTIFICATION  AND  CHEMICAL  REACTION    365 

The  value  of  Xi0/x*  is  not  known  until  the  amount  of  polymerization  is  known,  so 
it  is  assumed  as  0.0252. 


-  0.0826  -f  0.1475(0.1225)  +  0.144(0.0252)  «  0.104 


With  Vt~i  =  Vt  (subscript  tt  t  —  1  refer  to  top  plate  and  plate  below,  respec- 
tively), the  polymerization  will  make  Ot  less  than  OR  by  the  decrease  in  mols,  p/2. 

0<  .  0R  -  %  m  0.936  -  0.0195  »  0.916 
& 

Assuming  that  0.00082  more  mols  of  Cio  enter  the  top  plate  in  the  vapor  than 
leave  it  in  the  distillate, 


Ozio  -      +  0.00082  -  0.02027 

<u 

+  Ox,  =  0t-  Ozio  =  0.916  -  0.020  -  0.896 


U.OoO 


,*    t—r\n 

°'798 


&       LT225 
n  SQR 

0x7  =        (°-1225)  -  °- 


0.02027/0.798  —  0.0254,  which  checks  the  assumption  satisfactorily. 


#10 
2/io  =  — 


0.020 
:  0.0976 


(0.02)  (0.10)  -  0.00041 


which  checks  the  assumption  that  it  is  negligible. 

The  following  table  may  now  be  constructed  for  the  top  plate : 


Comp. 

yt 

Otxt 

DxD 

5 

0  98 

0.798 

0  3215 

7 

0.02 

0  0976 

0.00656 

10 

0  00041 

0.02027 

0.000134 

Plate*  -  1: 

Calculations  may  proceed  downward  to  the  next  plate  by  making  material  bal- 
ances on  components  Cg  and  €7  and  by  assuming,  and  later  checking,  vapor  values 
for  Cio. 


For  C8, 


DXD  +  Oxt  +  pt  -  0,3215  +  0.798  +  0.0389  -  1.159 


366 
For  C7, 


FRACTIONAL  DISTILLATION 


Vyt-i  -  DxD  4-  #*t  -  0.00656  +  0.0976  -  0.1042 
Knowing  the  ratio  of  the  y's,  the  ratio  of  the  a?'s  may  be  found,  using  a 

x7         0.1042 


1,159 


and  assuming  XIQ/X&  »  0.0495, 


-  0.0826  +  0.1475(0.540)  4-  0.144(0.0495)  -  0.1694 


Assuming  that  0.0006  more  mols  of  Cio  enter  plate  (t  —  1)  in  the  vapor  than  leave 
in  the  vapor, 

Oxio  -  0.02027  +  0.0074  -f  0.0006  -  0.0283 

0M  -  Ot  -  ^  -  0.916  -  0.0074  -  0.909 
2 

Ox6  -f  Ox1  -  0.909  -  0.028  -  0.881 

0  ^4. 
Oz7  -  j=~  0.881  -  0.309        and        Ox,  -  0.572 

These  calculated  values  may  now  be  tabulated. 
Tabulating  for  t  —  1  : 


Comp. 

Vy 

Ox 

5 

1.159 

0.572 

7 

0.1042 

0.309 

10 

0.00095 

0  0283 

The  value  of  Vyw  is  calculated  using 


(0.1042)(0.10)  -  0.00095 


Thus,  0.00095  -  0.00013  -  0.00082,  which  checks  the  assumptions  made  for  the 
top  plate.    XIQ/XI  —  0.0283/0.572  «•  0.0495,  which  checks  the  assumption  used 
to  calculate  r. 
Plate  t  -  2: 


For  C6, 
For  Or, 


-  0.3215  +  0.0389  +  0.0148  +  0.572  -  0.947 
Vyt-*  -  0.00656  -f  0.309  »  0.3155 

5  -•TOT-1'"8 


SIMULTANEOUS  RECTIFICATION  AND  CHEMICAL  REACTION    367 


Assume  that  X 


«  0.0986. 


~  -  0.0826  +  0.1475(1.998)  +  0.144(0.0986) 


0.391 


Assume  that  0.0008  more  mols  of  Cio  are  in.  the  vapor  from  plate  (t  —  2)  than  in 
the  vapor  to  this  plate.     Therefore, 


Ozio  -  0.0283  4-  0.0014  -  0.0008 
On  -  0,909  -  0.0014  -  0.908 
Ox,  +  Ox,  -  0.908  -  0.029  -  0.879 


0.0289 


Ox,  - 
Tabulating  for  t  —  2: 


0.293 


and        Ox,  -  0.586 


Comp. 

7y 

Ox 

5 

0.947 

0.293 

7 

0  3155 

0  586 

10 

0.00155 

0.0289 

72/1 


(°-10)(°-3155)  -  0.00155 


The  Cio  condensing  on  plate  (<  —  1)  -  0.00155  -  0.00095  -  0.0006,  which 
checks  the  assumption  used  in  finding  Oxio  for  plate  (t  —  l). 

XIQ/XS  »  0.0289/0.293  «  0.0986,  which  checks  the  value  used  to  calculate  n_2. 
Plate  i  -3  -  feed  plate: 

The  ratio  of  x*/xi  is  approaching  that  in  the  feed,  and  trial  calculations  show  that 
plate  (t  —  3)  should  be  the  feed  plate.  To  calculate  0  values  for  this  plate,  the 
value  of  F  must  be  known.  To  find  F,  assume  that  0.0571  mol  of  the  Cs  fed  to  the 
tower  polymerizes.  Over-all  balances  give 


F  -  W  +  D 


+  0.357 


By 


B  balance,  assuming  xw  **  0.004, 

0.20F  -  0.3215  4-  0.004TF  +  0.057  -  0.004TT  -f  0.3786 
Solving, 

F  -  1.915        and        W  **  1.56 
For  C5  +  2Cio, 

TF&TT  -  1.915(0.20)  -  0.3215  -  0.0615 
For  CT, 

Wxw  -  1.915(0,80)  -  0.00656  -  1.526 


368 


FRACTIONAL  DISTILLATION 


Calculations  for  plate  (t— 3),  the  feed  plate: 
For  C8, 

Vyt-i  -  0.3215  4  0.293  4  0.0389  4  0.0148  4  0.0028  »  0.671 
For  C7, 


<  0.00656  4  0.586 
n  0.5925 


0.5925 


x& 


0.671 


5.30 


Assume 


—  -  0.0661;        -  -  0.0826  4  0.1475(5.30)  4  0.144(0.0661)  -  0.875 

#6  T 


(4.25)10- 


0.000555;        |  =  0.000277 


^~*         (0.875)2 
Assume  that  0.002  mol  of  Cio  condenses  on  t— 3,  then 

0z10  =  0.0289  4  0.00027  4  0.0002  -  0.0293 
Ot-t  -  0.908  -  0.0003  4  1.915  =  2.823 
OzB  4  Ox-i  **  2.823  -  0.029  -  2.794 
2.794 


^  6W 
Tabulating  for  t-3: 


••  0.443        and        Oz7  =  2.351 


Comp. 

Vy 

Ox 

5 

0  671 

0  443 

7 

0  5925 

0.351 

10 

0  00074 

0  0293 

(0.5925)  (0.10)  -  0.000739 


Mols  Cio  vaporizing  from  J-2  -  0.00155  -  0.00074  =  0.0008 
Checking  the  assumption  used  in  calculating  Ox™  for  plate  (£—2), 


xio       0.0293 
xi        0.443 


=  0.0661 


which  checks  the  assumption  used  to  find  n_3. 

The  calculations  for  the  succeeding  plates  are  carried  out  in  the  same  manner  and 
the  results  are  given  in  the  following  table: 


Comp. 

(VJ0I-4 

(0aO,-4 

(Vy)t~* 

(Ox)t-t 

(Vy)*-* 

(OaOi-, 

5 

0.439 

0.227 

0.222 

0.0956 

0.0912 

0.0357 

7 

0.825 

2  567 

1.041 

2  697 

1  171 

2.757 

10 

0  00095 

0  0295 

0  00114 

0.0296 

0.0013 

0.0306 

P 

0.00013 

Neglected 

Neglected 

SIMULTANEOUS  RECTIFICATION  AND  CHEMICAL  REACTION    369 


Comp. 

(7y)« 

Wxw 

Xw 

5 

0.0313 

0.00646 

0.004 

7 

1.231 

0  526 

,0.978 

10 

0  0023 

0  0283 

0.018 

The  value  of  0.057  mol  polymerized  is  approximately  correct,  and  the  per  cent  of 
cyclopentadiene  polymerized  is 

0.057(100) 


1.915(0.20) 


1.49  per  cent 


The  number  of  plates  required  is  seven. 

The  solution  required  a  large  amount  of  trial  and  error.    An  approximate  value 
could  be  obtained  by  treating  the  mixture  as  a  binary  of  Ct>  and  C?  on  the  usual  y,x 


01     02     03     0.4     0.5    06    0.7    0.8     0.9    10 

x-mol  fraction  cyclopentadiene  in  liquid 
FIG.  13-1.     Fractionation  of  cyclopentadiene. 

diagram  and  then  calculating  the  amount  of  polymerization  from  the  plate  com- 
position so  obtained.  Such  a  diagram  is  shown  in  Fig.  13-1,  and  six  plates  are 
required.  The  plates  are  fewer  because  in  the  actual  case  the  polymerization 
decreases  the  effective  reflux  ratio  for  the  C5  +  C7.  The  polymerization  was 
calculated  neglecting  the  Cio  in  the  liquid  and  was  found  0.057  mol  of  Cs.  This 
value  agrees  well  with  the  value  found  by  the  plate-to-plate  calculations. 


CHAPTER  14 
BATCH  DISTILLATION 

Continuous  operation  is  normally  employed  when  the  material  to  be 
distilled  is  large  in  quantity  and  is  available  at  a  reasonably  uniform 
rate.  Under  such  conditions,  it  is  usually  cheaper  than  batch  dis- 
tillation, but  there  are  a  large  number  of  cases  which  are  not  suited 
to  continuous  operation  and  which  are  handled  on  the  batch  basis. 

A  batch  distillation  with  rectification  involves  charging  the  still  with 
the  material  to  be  separated  and  carrying  out  the  fractionation  until 
the  desired  amount  has  been  distilled  off.  The  overhead  composition 
will  vary  during  the  operation,  and  usually  a  number  of  cuts  will  be 
made.  Some  of  the  cuts  will  be  the  desired  products,  while  others  will 
be  intermediate  fractions  that  can  be  recycled  to  subsequent  batches  to 
obtain  further  separation. 

The  equipment  employed  and  the  method  of  operation  are  similar 
for  batch  and  continuous  distillation,  but  in  the  latter  the  mathematical 
analysis  is  based  on  the  ^ssujnoptimi  that  in  all  portions  joj^e 
the  compositions  and  flowlra^^  tkpyfij 


^ 

tion.  TEeie^ohditiOM^o^  n^*"af3>p1y^t  o  aT  batch  9Tstillation,  and 
TJecause  of  the  continuous  variations  involved,  it  must  be  analyzed  on 
a  differential  basis.  As  a  result,  the  calculations  are  much  more  diffi- 
cult, and  satisfactory  design  methods  have  been  developed  for  only  a 
few  simple  cases. 

BINARY  MIXTURES 

The  batch  distillation  of  binary  mixtures  will  be  considered  fpr  the 
cases  of  (1)  no  rectification,  (2)  rectification  without  liquid  holdup  in 
the  column,  and  (3)  rectification  with  holdup. 

No  Rectification.  Batch  distillation  without  rectification  cor- 
responds to  the  simple  distillations  of  Chap.  6,  and  the  calculations  of 
the  concentrations  as  a  function  of  the  amount  distilled  can  be  made 
by  Eqs.  (6-3)  and  (6-7). 

Rectification  without  Liquid  Holdup  in  Column.  Finite  Reflux  Ratio, 
In  this  case,  it  is  assumed  that  the  distillation  is  carried  out  with  a 
fractionating  column,  that  the  holjaqxQf  liquid  in  the  column  is  negligi- 

-  —  ~**  '      ~"~       •  —  —  —  *~- 


BATCH  DISTILLATION 


371 


ble  in  comparison  to  the  liquid  in  thej^ysiem,  and  that  the  rate  of 
£ha5^^  plate  is  negligible  in  com- 

parison to  the  rate  of  flow  of  that  component  through  the  plate.  Thus 
the  change. in  the  quantity  of  each  component  in  the  column  can  be 
neglected  in  the  differential  material  balances. 

Consider  the  system  shown  in  Fig.   14-1.    On  the  basis  of  the 
assumptions  made,  an  over-all  differential 
material  balance  gives 

dD  -  -dL 

where  L  represents  the  mols  of  liquid  in 
still,  and  a  component  balance  gives 

=  -~d(LxL) 

=  — L  dxL  —  XL  dL 

=  — L  dxL  +  XL  dD 
dD  _      dxL 
L 


XL  —  XD 


(14-1) 


This  equation  is  equivalent  to  the  Ray- 
leigh  equation  but  differs  in  that  the  denom- 
inator is  XL  —  XD  instead  of  XL  —  VL.  The 
integration  of  Eq.  (14-1)  involves  determin- 
ing a  relation  between  XL  and  XD.  Neglecting 
the  rate  of  change  of  holdup  of  a  component 
on  the  plates  and  in  the  condenser,  a  balance 
between  the  n  and  n  +  1  plates  gives 

Vn 


FIG.  14-1.     Schematic  diagram 
of  batch  distillation  system. 


At  any  time  these  equations  are  identical  to  those  for  the  continuous 
distillation  and  can  be  applied  to  determine  the  relation  between  XL 
and  XD  for  that  instant.  By  applying  the  equations  repeatedly,  the 
value  of  XL  —  XD  can  be  obtained  as  a  function  of  XL  and  the  integration 
of  Eq.  (14-1)  performed.  There  are  a  number  of  ways  such  a  dis- 
tillation can  be  made,  but  the  two  most  common  cases  involve  (1) 
operating  at  constant  reflux  ratio  and  taking  cuts  that  average  the 
desired  composition  and  (2)  operating  at  variable  reflux  ratio  to  give 
a  constant  product  composition  while  making  the  desired  product. 
In  the  first  case  the  value  of  (On/V^i)  will  remain  constant  during  the 
distillation,  and  a  series  of  lines  of  this  slope  can  be  drawn  on  the  usual 
y,x  diagram  for  various  assumed  values  of  XD,  and  the  value  of  XL  can 
be  determined  by  stepping  down  each  line  the  number  of  theoretical 


372 


FRACTIONAL  DISTILLATION 


FIG.  14-2.     Constant  reflux  ratio  case. 


FIG.  14-3.     Constant  distillate  composition  case. 

plates  equivalent  to  the  column.  This  procedure  is  illustrated  in 
Fig.  14-2  for  a  column  and  still  equivalent  to  four  theoretical  plates. 
In  the  second  case  the  value  of  XD  is  fixed,  a  series  of  operating  lines  of 
different  slopes  is  drawn  through  it,  and  the  plates  are  stepped  off  on 
each  line  to  determine  the  value  of  XL.  This  procedure  is  illustrated  in 
Fig.  14-3.  By  these  procedures,  Eq.  (14-1)  can  be  integrated,  giving 
the  relation  between  the  amount  distilled  and  the  composition  of  the 


BATCH  DISTILLATION  373 

liquid  remaining  in  the  still.  In  the  second  case  with  XD  constant,  the 
integration  of  Eq.  (14-1)  amounts  to  only  simple  material  balance,  and 
the  data  from  Fig.  14-3  are  not  needed  for  this  purpose.  In  addition 
to  the  information  obtained  by  integration  of  E(j.  (14-1),  it  is  frequently 
necessary  to  have  data  on  the  vapor  required  or  the  average  composi- 
tion of  a  fraction  produced. 

For  the  constant  reflux  case,  the  calculation  of  the  vapor  require- 
ment can  be  made  as  follows: 


F  =  ^  +  ljD  (14-2) 

and  the  average  composition  of  any  fraction  is 

vi/av    ~~~"    " '"""  7- 

For  the  variable  reflux  case, 

dV  =  dO  +  dD 


and,  by  material  balance, 

LXL  = 

r       Lo(&o  - 


XL  —  XD 


where  L0  =  original  mols  in  still 

x0  =  original  composition  of  liquid  in  still 
Substituting  these  values  in  Eq.  (14-1)  gives 

—          •b0(x0  —  Xp)  dXL 


Form  Fig.  14-3  the  relation  between  XD,  XL,  and  the  slope  of  the 
operating  line,  dO/dV,  can  be  obtained  and  Eq.  (14-4)  integrated  to 
give  the  vapor  requirement. 

The  use  of  these  methods  will  be  illustrated  by  the  following  example. 

Batch  Fractionation  of  a  Binary  Mixture.  An  equimolal  mixture  of  A  and  B  is 
to  be  fractionated  in  a  batch  column  equivalent  to  three  theoretical  plates  plus  a 
still.  The  still  will  operate  at  atmospheric  pressure,  with  a  total  condenser,  and  the 


374 


FRACTIONAL  DISTILLATION 


holdup  in  the  column  and  condenser  is  negligible.    The  company  desires  to  obtain 
an  overhead  A  product  containing  95  moi  per  cent  A,  and  two  methods  of  operation 
have  been  suggested:  (1)  Operate  the  column  at  a  constant  reflux  ratio  (0/D) 
equal  to  5.0  and  continue  the  distillation  until  the  average  composition  of  the 
distillate  is  95  per  cent  A;  (2)  Operate  the  column  at  a  variable  reflux  ratio  to 
give  a  distillate  of  constant  composition. 
Using  data  and  notes  given  below,  calculate: 
1.  For  Method  1, 

a.  The  mol  per  cent  of  the  original  charge  to  the  still  that  can  be  obtained  as 

the  95  per  cent  distillate. 

6.  The  mols  of  vapor  per  100  mols  of  original  charge  to  obtain  the  distillate 
of  Part  a. 


1.U 

0.9 
0.8 
07 
|0.6 

<0.5 
§ 

^04 

J0.3 
02 

at 
°( 

^ 

•fit) 

/ 

sf 

> 

^ 

/ 

/ 

^ 

/ 

/ 

'/ 

' 

/ 

/ 

/ 

V 

/. 

y 

7 

// 

/ 

/ 

/ 

A 

/ 

/ 

£ 

)      at     0,£     03     04    05     0.6     017     0.8    0,9     t.( 
Mol  fraction  A  in  liquid 

FIG.  14-4. 

2.  For  Method  2,  the  mol  per  cent  of  the  original  charge  to  the  still  that  can  be 
obtained  as  the  95  per  cent  distillate,  using  a  total  vapor-to-charge  ratio  equal  to 
that  of  Part  1,  6. 

Assume  «AB  is  constant  at  2.5. 

Solution.  By  trial  and  error  the  relations  between  XL  and  XD  could  be  deter- 
mined analytically  by  Eq.  (7-62)  because  the  relative  volatility  is  constant  and  the 
usual  simplifying  assumptions  apply.  However  a  y,x  diagram  is  probably  simpler 
and  will  be  employed.  The  mol  fractions  of  component  A  will  be  used  in  the 
calculations. 

The  equilibrium  curve  was  calculated  from  the  relative  volatility  and  is  given 
in  Fig.  14-4. 

Method  1.  The  slope  of  the  operating  line  -  0/V  «•  %  «  0,833.  The  rela- 
tion between  any  x&  and  the  corresponding  XD  is  found  by  taking  four  steps*  from 


BATCH  DISTILLATION 


375 


XD  on  the  operating  line.    The  case  for  XD  =*  0.94  is  shown  in  Fig.  144  and  gives 
XL  =  0.41.    The  values  for  other  cases  are  given  in  Table  14-1. 

TABLE  14-1 


XD 

XL 

1 

lnL< 

K 

w. 

XL  —  XD 

ln  L 

0  98 

0  658 

~3  11 







0.96 

0.490 

-2  13 

0  022 

0  978 

0.962 

0  94 

0,410 

-1  89 

0.182 

0  834 

0  955 

0  92 

0.344 

-1  73 

0  299 

0  742 

0  948 

0.90 

0.293 

-1.65 

0  379 

0.678 

0  937 

The  values  of  I/ (XL  —  XD)  are  plotted  as  a  function  of  XL,  and  the  graphical 
integration  is  performed  from  XL  =  0.5  to  XL-  The  resulting  values  of  the  integral 
are  equal  to  In  (L0/L)  and  are  tabulated  in  the  fourth  column  of  the  table.  The 
fifth  column  gives  the  values  of  L/L0  and  the  average  composition  of  all  of  the 
distillate  from  the  start  of  the  distillation  as  calculated  by  Eq.  (14-3).  By  plotting 
the  average  value  of  XD  vs.  (L/L0),  it  is  found  that  (£z>)av  **  0.95  for  (L/L0)  «•  0.82. 
Therefore,  mol  per  cent  of  original  charge  recovered  as  95  per  cent  distillate  is  18.0. 


For 
and 


Lo  -  100  mols,  D  -  18 

V  -  total  vapor  =  0  +  D  -  6Z>  -  6(18) 
-  108  mols 


Method  2.  In  this  case  it  is  necessary  to  calculate  the  per  cent  recovery  of  the 
original  charge  as  95  per  cent  distillate  for  a  total  heat  supply  equal  to  108  mols  of 
vapor  when  operating  on  the  variable  reflux  basis.  This  result  can  be  evaluated  by 
integrating  Eq.  (14-4).  For  various  assumed  values  of  the  slope  of  the  operating 
line  dO/dV,  the  lines  were  drawn  through  XD  =  0.95  and  the  plates  stepped  down 
to  obtain  XL-  The  values  are  summarized  in  Table  14-2  for  L0  =  100. 

TABLE  14-2 


dO 
dV 

XL 

100(0.5  -  0.95) 

F 

(0.95  -  xL)*(l  -  ^) 

1.0 

0.330 

00 

00 

0.95 

0  360 

-2,590 

176  2 

0  90 

0.390 

-1,430 

121.5 

0  85 

0.435 

-1,130 

65.7 

0.80 

0.480 

-     968 

19.1 

0.75 

0.510 

-;     937 

(•—  «• 

376  FRACTIONAL  DISTILLATION 

The  values  given  in  the  third  column  of  Table  14-2  were  plotted  vs.  XLt  and  the 
area  under  the  curve  from  XL  —  0.5  to  XL  is  equal  to  the  total  vapor  necessary  ta 
reduce  the  still  concentration  to  XL.  A  plot  of  XL  against  V  gives  XL  =  0.398  at 
V  -  108  and 

0.398L  +  0.95  (L0  -  L)  -  0.5L, 
~  -  0.815 

LIO 

Recovery  as  95  per  cent  distillate  =  18.5  per  cent  of  original  charge.  This 
would  indicate  that  Method  2  gave  a  slightly  better  recovery  for  the  same  heat 
consumption,  but  the  accuracy  of  the  calculations  is  not  sufficient  to  make  the 
difference  significant.  Method  1  would  be  more  practical  from  an  operating 
viewpoint. 

Total  Reflux.  Limiting  conditions  can  be  calculated  for  batch  dis- 
tillation that  are  useful  for  orientation  purposes.  The  total  reflux 
limit  applies  to  the  constant  reflux  category  and  can  be  used  to  deter- 
mine the  minimum  number  of  theoretical  plates  necessary  to  give  the 
desired  product  recovery  or  to  determine  the  maximum  possible  recov- 
ery for  a  given  number  of  plates.  Consider  an  equimolal  mixture  of 
A  and  B  having  constant  relative  volatility  of  2,  and  make  the  usual 
simplifying  assumptions.  The  desired  distillate  is  to  contain  95  mol 
per  cent  A,  and  the  limits  to  be  estimated  are  (1)  the  number  of  theo- 
retical plates  for  a  50  per  cent  recovery  of  A  in  the  distillate  and  (2)  the 
maximum  per  cent  recovery  of  A  with  a  tower  equivalent  to  five  theo- 
retical plates. 

For  total  reflux  and  constant  relative  volatility,  by  Eq.  (7-53), 


XD  "  * 


Substituting  this  value  in  Eq.  (14-1), 

^  » fa* (14-6) 

L  aN+lXL 

\QL         ~— "  L)XL    I    J- 

This  equation  is  identical  with  Eq.   (6-5)  except  that  a**1  has 
replaced  a,  and  by  integration  from  x0,  L0  to  x,  I/, 


BATCH  DISTILLATION  377 

for  50  per  cent  recovery  of  A, 

(L.  -  L)0.95  =  0.5Le(0.5) 
-  0.737 


-  °'339 

Using  these  values  with  Eq.  (14-7)  gives  TV  =  3.8  theoretical  plates. 
For  maximum  per  cent  recovery  with  five  theoretical  plates, 

,    L  1      ,        3(0.5)       ,  .      0.5 

-          -  ' 


*"  T        96  —  i      n  *n  —  ^   '  "x 

JL/o          A     —    1         U.O^l   —  X) 

and 

LZ  +  0.95(Lo  -  L)  =  0.5L0 
L  0.45 


Lo      0.95  -  3 
Solving  these  equations  simultaneously  gives 

~  =  0.492,        x  =  0.0352 
L/o 

-P                   0.5L0  -  0.0352L  ^  lrkA 
Recovery  = 7r^? X  100 

U.Olvo 

=  96.3  per  cent 

It  will  be  noted  that  the  per  cent  recovery  increases  rapidly  as  the 
theoretical  plates  are  increased  in  number. 

If  the  relative  volatility  is  not  constant,  the  total  reflux  condition 
can  be  solved  using  the  y  —  x  line  as  the  operating  line  and  determining 
XL  as  a  function  of  XD  graphically. 

Minimum  Vapor  Requirements.  Another  limit  that  is  instructive  is 
the  minimum  vapor  requirement  for  a  given  separation  and  recovery. 
This  limit  will  correspond  to  the  minimum  reflux  ratio  condition  for 
continuous  distillation  and  will  require  an  infinite  number  of  theoretical 
plates.  For  this  case,  the  operating  line  must  contact  the  equilibrium 
curve  at  some  point.  For  the  general  case,  the  limit  can  be  deter- 
mined graphically,  for  either  the  constant  or  variable  reflux  conditions, 
by  drawing  such  operating  lines  and  obtaining  XL  as  a  function  of  XD  or 
dO/dV  and  integrating  Eqs.  (14-1)  and  (14-4). 

The  general  principles  can  be  illustrated  analytically  for  the  case  of 
constant  relative  volatility.  For  the  constant  reflux  method,  the 
operating  line  will  intersect  the  equilibrium  curve  at  either  XD  =  1  or 


378  FRACTIONAL  DISTILLATION 

at  x  *•  XL.  The  overhead  product  will  be  pure  volatile  component, 
Xj>  ••  1,  until  XL  decreases  to  a  value  such  that  the  operating  line 
through  XD  »  1  intersects  the  equilibrium  curve  at  the  composition  of 
the  still;  for  smaller  values  of  XL>  the  intersection  will  be  at  #L.  This 
limiting  value  of  XL  can  be  calculated  by 

0  _  i  -  yl 
V  ~  i  -  x* 

where  x*  =*  value  of  XL  such  that  operating  line  intersects  equilibrium 

curve  at  XD  =  1  and  x* 
y*  =  vapor  in  equilibrium  with  xj 
For  constant  relative  volatility, 


L  ~  o/v(«  -  i) 

for  values  of  XL  equal  to  or  greater  than  #*,  XD  will  be  equal  to  unity. 
For  values  less  than  x%, 


With  constant  relative  volatility,  substituting  for  yL  gives 

a  0~ 


(a  - 


[(a  -  l)afc  +  1]     1  -  rr 


This  relation  ean  be  combined  with  Eq.  (14-1)  and  integrated  from 
%  to  x  and  L*  to  L  to  give 


which  is  the  same  as  Eq.  (6-5)  except  for  the  (1  —  0/V)  term.    This 
equation  can  be  rewritten 


(14-12) 


BATCH  DISTILLATION  379 

The  average  composition  for  the  distillate  from  L0  to  L  is 


and  the  total  vapor  generated  from  L0  to  L, 


By  a  similar  procedure  an  equation  can  be  developed  for  the  variable 
reflux  case. 


V       XQ-  XD  [2(«  -  1)  +  (1  +  XD)  ,    (I  -  X)(XQ  -  XD) 
Lo       2(«-l)L  XD-I  m(l-x0)(x-xD) 


H^l 

/  J 


Minimum  Vapor  Requirements  for  Batch  Distillation.  Consider  the  batch  dis- 
tillation of  an  equimolal  mixture  of  A  and  B.  The  relative  volatility,  <XAB,  is 
constant  at  2;  and  the  average  distillate  is  to  be  95  mol  per  cent  A.  Calculate  the 
minimum  mols  of  vapor  for  50  mol  per  cent  recovery  'of  A  in  the  distillate  for  both 
the  constant  reflux  ratio  and  the  variable  reflux  ratio  cases. 
Solution 

0.95(L0  -  L)  »  0.25L, 

T-  -  0.737 

Lo 

Final  liquid  composition,  xL  «  0.25/0.737  —  0.339.  For  the  constant  reflux 
method,  assume  that  the  value  of  0/D  is  such  that  x%  will  be  less  than  0.5;  then 
by  Eq.  (14-8) 

*       1  -  0/7      D 

XL  .  ___  _  ^ 

By  over-all  material  balance 

L*       I  -  x0  0.5 


Lo       1  -  *J      1  -  D/0 

For  the  portion  of  the  distillation  when  the  liquid  in  the  still  is  decreasing  from 
x*L  to  0.339,  Eq.  (14-12)  gives 


\  -  1  474  (l  -  -^  -  fl"i>/Q  [0-880(1  - 
V       ^'^V        O/       t    0.661     L     D/0(0.661)       / 

Solution  of  this  relation  by  trial  and  error  gives  0/D  **  2.12.    This  value 
gives  #J  «  0.472,  indicating  that  the  assumption  made  was  correct.    Then,  by  • 


380  FRACTIONAL  DISTILLATION 

Eq.  (14-14), 

~  -  (2.12  +  1)(0.263)  -  0.82 


For  variable  reflux  operation, 

,    0.661  (-0.45)       .     (0.339)*(-  0.45)  (0.5)  1  _  ft  7fi7 
0*5(-0.611)  (0.5)2(  -0.611)  (0.661)  J  ~~  U*7b7 


V_       -0.45  [2  +  1.95,    0.661  (-0.45) 
Lo  "      2      L  .-0.05 


As  was  found  in  the  example  page  376,  the  variable  reflux  method  requires  less 
vapor  for  a  given  separation  than  the  constant  reflux  method. 

Rectification  with  Liquid  Holdup  in  Column.  In  continuous  distilla- 
tion holdup  does  not  harm  the  degree  of  separation  obtainable;  in  some 
cases  it  may  be  an  advantage  in  that  it  gives  "flywheel"  action  and 
tends  to  smooth  out  the  operation.  The  effect  of  column  holdup  in 
batch  distillation  is  not  completely  clear  because  a  satisfactory  method 
of  analyzing  the  operation  has  not  been  developed.  Rose,  Welshans, 
and  Long  (Ref  .  3)  present  approximate  equations  for  total  reflux,  and 
Colbura  and  Stern  (Ref.  1)  have  discussed  the  case  of  finite  reflux. 
Rose  et  al.  concluded  that  holdup  was  detrimental  at  total  reflux,  and 
Colburn  and  Stern  indicated  that  holdup  in  some  cases  improved  the 
sharpness  of  separation  as  compared  to  no  holdup.  In  order  to  inves- 
tigate this  difference  of  opinion,  a  number  of  cases  were  studied  analyti- 
cally, and  holdup  was  not  advantageous  in  any  of  them.  For  the  same 
number  of  theoretical  plates,  reflux  ratio,  and  total  mols  vaporized, 
the  no-holdup  system  was  superior  in  all  the  cases  tried.  It  is  believed 
that  the  improved  results  reported  with  column  holdup  were  due  to  the 
fact  that,  for  the  starting  condition,  a  concentration  gradient  was 
already  established  in  the  rectifying  column  which  had  required  the 
expenditure  of  considerable  vapor  which  should  be  but  was  not  included 
in  the  evaluation.  The  amount  of  vapor  required  to  establish  this 
concentration  gradient  can  be  estimated  from  Fig.  18-3.  For  a  sys- 
tem with  a  relative  volatility  of  2.0,  the  "prerun"  vapor  would  be 
about  eight  times  the  column  holdup.  Thus  a  column  holdup  equal 
to  15  per  cent  of  the  charge  to  still  would  require  a  prerun  vapor  gen- 
eration of  approximately  1.2  times  the  charge  to  the  still.  If  the 
column  holdup  was  obtained  by  charging  it  with  the  fresh  still  liquid 
and  including  all  vapor  from  the  beginning  of  the  distillation,  the 
holdup  results  would  have  been  less  favorable.  Alternately,  if 
the  no-holdup  case  was  given  the  extra  vapor  corresponding  to  the 
"tune-up"  period  for  the  holdup  case,  it  would  be  more  favorable. 
It  is  concluded  that  in  general  holdup  in  the  rectifying  column  is  unde- 
sirable in  batch  distillation. 


BATCH  DISTILLATION  381 

There  is  an  intermittent  type  of  batch  operation  for  which  column 
holdup  has  an  apparent  advantage.  In  this  case  the  column  is  run  at 
total  reflux,  and  the  top  plates  are  filled  with  liquid  rich  in  the  more 
volatile  component.  Product  is  then  withdrawn  at  a  high  rate  for  a 
short  time  after  which  the  column  is  put  back  on  total  reflux  to  reestab- 
lish a  concentration  gradient.  Most  of  the  product  withdrawn  during 
the  short  time  interval  comes  from  the  accumulation  of  volatile  com- 
ponent on  the  top  plates.  For  this  type  of  operation,  improved 
results  would  be  obtained  without  liquid  holdup  in  the  column,  but 
with  a  reservoir  in  the  reflux  line.  The  column  would  operate  at  total 
reflux  until  the  liquid  in  the  reservoir  was  rich  in  the  volatile  compo- 
nent. This  liquid  would  then  be  withdrawn  completely,  and  the  col- 
umn returned  to  total  reflux  operation  to  prepare  the  next  fraction. 
This  type  of  operation  is  frequently  convenient  for  laboratory  distilla- 
tions but  is  not  often  advantageous  for  large-scale  operation. 

A  detailed  differential  stepwise  integration  for  a  batch  distillation 
with  holdup  can  be  made,  but  the  time  and  effort  involved  are  usually 
not  justified  by  the  value  of  the  result.  The  method  is  useful  in 
developing  the  principles  of  batch  distillation  with  holdup.  The  fol- 
lowing section  will  consider  the  basic  differential  equations  which  are 
helpful  in  obtaining  a  qualitative  picture  of  the  process.  By  an  analy- 
sis similar  to  that  for  Eq.  (14-1), 

xDdD  =  -d(LxL)  -  d(Hxff)  (14-15) 

where  H  =  holdup  in  column 

XH  =  average  composition  of  holdup 

The  integration  of  this  equation  requires  a  relationship  between  HXH 
and  L  and  XL.  This  relationship  is  very  complex  and  has  not  been 
expressed  in  a  form  suitable  for  direct  integration.  It  has  been  cus- 
tomary to  simplify  this  equation  by  assuming  that  H  is  constant.  In 
some  cases  this  may  be  a  reasonable  assumption,  but  on  a  molal  basis 
it  could  be  greatly  in  error.  For  example  in  the  distillation  of  an  iso- 
propanol-water  mixture,  a  given  volume  of  holdup  could  give  a  fourfold 
variation  in  molal  holdup.  If  Eq.  (14-15)  is  applied  on  a  weight  and 
weight  fraction  basis,  the  variation  of  holdup  during  the  distillation 
will  be  less,  but  it  can  still  be  large.  Analysis  based  on  the  constancy 
of  H  will  be  only  approximate  for  most  cases. 

An  operating  line  for  this  case  can  be  written  by  taking  a  material 
balance  on  one  of  the  components, 

yn  dVn  »  xn+i  dOn+i  —  d(LxL)  -  d(HmxHm)  (14-16) 


382  FRACTIONAL  DISTILLATION 

or 


dVn  -  xn+i  dO^i  +  XD  dD  -  d(HnVan)  (14-17) 


where  Hmxffm  and  HnXan  equal  holdup  of  component  for  plates  from 
still  to  plate  n,  and  for  plate  n  +  I  to  top  of  column,  respectively. 

Because  the  last  terms,  d(Hnpcan?)  and  d(JHn##n),  vary  from  plate  to 
plate,  the  operating  line  is  curved  and  cannot  be  evaluated  without  a 
knowledge  of  the  variation  of  these  terms.  At  total  reflux,  these  terms 
are  negligible  in  comparison  to  the  first  two  terms,  and  the  operating 
line  is  the  y  =  x  diagonal  with  holdup  the  same  as  it  was  for  no  holdup, 
but  for  finite  reflux  ratio  the  operating  line  with  holdup  is  curved  and 
difficult  to  establish  on  the  y,x  diagram. 

When  a  batch  distillation  is  carried  out  by  the  variable  reflux  ratio 
method  to  give  a  constant  value  of  XD,  and  the  distillation  is  continued 
until  the  reflux  ratio  is  essentially  total  reflux,  the  amount  of  holdup  in 
the  column  at  the  end  of  the  distillation  can  be  easily  calculated  by 
using  the  y  =  x  line  as  the  operating  line.  Such  a  procedure  gives  the 
composition  of  the  liquid  on  each  plate,  and  a  correction  can  be  applied 
for  effect  of,  the  holdup  on  the  percentage  yield  of  a  given  fraction. 

Batch  Rectification  of  Binary  Mixture.  Assume  that  the  equimolal  mixture 
already  considered,  page  376,  is  to  be  distilled  in  this  manner,  and  it  is  desired  to 
evaluate  the  effect  of  the  column  holdup. 

Solution.    Basis:  100  mols  liquid  charged  to  still,  no  liquid  in  column  at  start. 

At  the  end  of  the  distillation  the  results  will  correspond  to  total  reflux  and,  by 
Eq.  (14-5), 


2«(0.05)  +  0.95 
-  0.229 

By  material  balance, 

5 

0.5(100)  -  0.95D  +  0.229L  -f  Y  (hx) 

l 
5 

100  -  D  -f  L  +  Y  h 
T 


5 
w* 


here  }  (hx)  —  total  mols  of  A  in  the  holdup  on  the  plates  at  total  reflux 

5 

y  (h)  ••  total  mols  of  holdup  on  the  plates 


BATCH  DISTILLATION  383 

5 


For  the  case  in  which  the  total  holdup  per  plate  is  constant,  the  value  h  y  x 

can  be  obtained  approximately  in  a  mathematical  form  (Ref.  2),  but  it  is  just  as 
easy  in  most  cases  to  evaluate  it  plate  by  plate.  In  the  present  case,  assuming  the 
holdup  per  plate  is  constant, 

)**  h(xi  H-  xz  +  x*  +  #4  +  *e) 

-  M0.371  -f  0.541  -f  0.701  +  0.823  +  0.903) 

-  3.35/1 
and 

50  -  0.95D  +  0.229(100  -  D  -  5A)  +  3.3 
27.1  -  2.205fc  -  0.721D 

n      27.1  -  2.205fc 


If  the  holdup  per  plate  were  2  per  cent  of  the  original  charge  to  the  still,  then 

h  -  2 
and  the  mols  of  the  distillate  would  be 

n  -  27.1  -  2.205(2) 

0.721 
-  31.5 

as  compared  to  D  =  37.5  for  no  holdup.  If  the  value  of  holdup  per  plate  was  as 
high  as  12.3  per  cent  of  the  charge,  no  95  per  cent  distillate  could  be  produced  in  the 
5-plate  column.  If  the  percent  of  A  in  the  original  feed  had  been  smaller,  the 
effect  of  holdup  would  have  been  even  more  serious. 

A  similar  analysis  has  been  applied  by  some  writers  to  the  case  where  the  dis- 
tillation has  been  discontinued  at  a  finite  reflux  ratio.  The  usual  straight  operat- 
ing line  was  employed  to  determine  the  plate  composition  at  the  end  of  the  distilla- 
tion, but  in  view  of  the  fact  that  the  operating  line  is  curved  for  such  a  case,  the 
calculations  are  probably  of  little  value.  If  the  operating  lines  corresponding  to 
Eq.  (14-16)  or  (14-17)  could  be  evaluated  for  the  end  of  the  distillation,  then  an 
analysis  of  the  effect  of  holdup  could  be  made. 

MULTICOMPONENT  MIXTURES 

The  state  of  the  art  for  the  batch  distillation  of  multicomponent 
mixtures  is  even  less  satisfactory  than  for  binary  mixtures,  and  except 
for  total  reflux,  no  accurate  and  practical  method  is  available  even 
without  liquid  holdup  in  the  column.  This  difficulty  arises  from  the 
fact  that,  starting  with  a  given  liquid  composition  in  the  still,  it  is 
possible  to  calculate  the  equilibrium  vapor  leaving  the  still,  but  the 
composition  of  the  overflow  to  the  still  from  the  first  plate  cannot  be 
calculated  without  knowing  the  composition  of  the  distillate  leaving 
the  system.  In  a  binary  system,  it  is  possible  to  choose  the  composi- 


384  FRACTIONAL  DISTILLATION 

tion  of  the  distillate  arbitrarily  and  calculate  back  to  the  liquid  in  the 
still,  and  the  liquid  composition  in  the  actual  distillation  must  pass 
through  this  condition.  However,  in  a  multicomponent  mixture,  the 
still  composition  calculated  in  that  manner  for  an  assumed  overhead 
product  will  probably  not  occur  in  an  actual  distillation.  For  exam- 
ple, consider  the  distillation  of  a  mixture  of  components  A,  B,  and  C 
and  assume  an  overhead  composition  of  XDA.J  #Z>B,  and  XDC.  For  these 
assumed  values,  the  calculations  can  be  carried  down  the  column  for 
a  given  reflux  ratio  assuming  no  holdup,  and  values  of  #LA,  #LB,  and 
XLC  will  be  obtained.  However,  in  the  actual  distillation  when  com- 
ponent A  has  the  value  XL±,  it  is  very  improbable  that  the  ratio  of  B  to 
C  will  be  the  same  as  calculated  on  the  basis  of  the  assumed  overhead 
composition.  By  a  laborious  trial-and-error  procedure,  a  consistent 
calculation  could  be  made,  but  it  is  doubtful  that  it  would  justify  the 
effort.  A  laboratory  distillation  would  probably  give  a  better  and 
cheaper  evaluation. 

As  a  guide  to  the  characteristics  of  multicomponent  batch  distilla- 
tion, the  case  of  (1)  total  reflux  with  no  liquid  holdup  in  the  column 
and  (2)  finite  reflux  ratio  with  no  liquid  holdup  by  an  approximate 
method,  will  be  considered. 

Rectification  without  Liquid  Holdup  in  Column.  Total  Reflux.  For 
the  case  of  no  liquid  holdup,  Eq.  (14-1)  applies  to  each  of  the  com- 
ponents, but  the  difficulty  involves  the  evaluation  of  XD  as  a  function 

of  XL. 

In  this  case  the  stepwise  calculations  can  be  carried  out  starting  at 
the  still,  because  the  composition  of  the  distillate  does  not  affect  the 
operating  line.  Thus  for  a  given  XL  the  value  of  XD  can  be  calculated, 
but  the  calculation  is  still  difficult  for  the  general  case  because  the  con- 
centration pattern  followed  by  the  liquid  in  the  still  is  unknown.  The 
pattern  can  be  approximated  by  a  stepwise  integration,  but  the  calcu- 
lations are  tedious.  The  general  principles  will  be  developed  for  the 
case  in  which  all  the  relative  volatilities  remain  constant,  because  in 
this  case  direct  integration  is  possible.  It  is  simpler  to  apply  the  rela- 
tive volatility  form  of  the  simple  distillation  of  Eq.  (6-6)  than  to  use 
Eq.  (14-1).  Thus,  for  any  two  components  of  a  multicomponent  mix- 
ture at  total  reflux, 


BATCH  DISTILLATION  385 

and,  for  a  column  with  a  total  condenser, 


(£).-(£),-*•(). 


. 

Let  a  differential  amount,  dV,  be  vaporized  and  set 

ywdV  =  —  dA 
y^T  dV  =  -dB 

where  A,  B  are  mols  of  A  and  B  in  still,  respectively. 


(14-19) 
integrating  from  A0to  A,  B0to  B 


A       /B\a 

T.  ~  (I.) 

Likewise, 


and  the  same  for  the  other  components. 

For  a  given  fraction  of  A  vaporized,  it  is  possible  to  calculate  (1)  the 
fraction  of  all  components  vaporized,  (2)  the  composition  of  the  liquid 
remaining  in  the  still,  (3)  the  instantaneous  composition  of  the  distil- 
late, and  (4)  the  average  composition  of  all  of  the  distillate. 

Batch  Rectification  of  Multicomponent  Mixture  at  Total  Reflux.  As  an  example, 
consider  the  fraction ation  of  an  equimolal  mixture  of  A,  B,  C,  and  D  at  total  reflux 
with  relative  volatility  «AB  =  2.0,  «AC  —  4.0,  and  a  AD  =  8.0.  The  column  is 
equivalent  to  three  theoretical  plates  and  holdup  of  liquid  will  be  neglected. 

Solution.     Basis:  100  mols  originally  charged  to  still. 

a.  First  distillate  composition: 

^  .  2*  -  16 

2/B 

^  -  4*  -  256 

yc 

^  -  84  «  4,096 

2/D  * 

2/A  4-  2/B  -f  yc  +  2/D  =*  1.0 

4.096  rt  n0o 

"A  '  4,096+256  +  16  +  1  =  °'938 

#B  «•  0.0586 
yG  **  0.00366 
2/D  »  0.00023 


386  FRACTIONAL  DISTILLATION 

b.  50  per  cent  of  A  distilled: 
B 


Bo 
C 


*  0.5Ho  »  0.958 

53*»*  -  0.9973 
-£  «  0.51/4'09'  -  0.99983 


The  compositions  of  (1)  liquid  remaining  in  still,  (2)  average  distillate,  and  (3) 
instantaneous  distillate  are  given  in  Table  14-3. 

TABLE  14-3 


(D 

(2) 

(3) 

Mols 

XL 

XD(*V) 

XD 

A 

0.125 

0  145 

0.922 

0.900 

B 

0.24 

0.278 

0.074 

0.0931 

C 

0.2494 

0.288 

0.004 

0.0061 

D 

0.25 

0,289 

— 

0.004 

0.864 

By  repeating  this  type  of  calculation,  the  values  of  the  instantaneous  distillate 
composition  given  in  Fig.  14-5  were  obtained. 


0.4  0.6  0,8  1.0 

Mo!  fraction  distilled 
FIG.  14-5.     Fractionation,  curves  for  multicomponent  mixture  at  total  reflux. 


3ATCH  DISTILLATION 


387 


In  the  example  on  p&ge  385,  about  60  per  cent  of  the  original  charge 
could  be  obtained  as  fractions  containing  one  of  the  components  in  at 
least  85  concentration.  It  will  be  noted  that  the  least  volatile  material 
can  be  obtained  in  the  highest  purity  because  \t  is  taken  over  when  the 
still  contains  the  l<*ast  amount  of  other  components.  By  operating 
an  inverted  batch  'Jolumn  as  shown  in  Fig.  14-6,  it  is  possible  to  remove 
the  less  volatile  components  and  obtain  the  most  volatile  material  in 
high  purity.  In  this  case,  the  batch  is  charged  to  the  reservoir,  and 


Reservoir 


Vaporizer*       ^Separator 
FIG.  14-6.     Inverted  batch  distillation  system. 

liquid  is  continuously  added  to  the  top  of  the  column  from  this  tank. 
The  liquid  from  the  bottom  of  the  column  is  partly  vaporized,  and  the 
unvaporized  portion  is  removed  as  product.  The  vapor  is  passed  back 
up  the  column  to  strip  out  the  more  volatile  components,  and  the 
overhead  vapor  is  condensed  and  returned  to  the  reservoir.  The  most 
volatile  component  collects  in  the  reservoir  and  is  obtained  as  the  last 
fraction  after  essentially  all  the  other  components  have  been  elimi- 
nated. An  intermediate  fraction  can  be  increased  in  concentration 


388 


FRACTIONAL  DISTILLATDN 


by  fractionating  first  in  a  normal  batch  operation  to  remove  the  lighter 
components  and  then  in  an  inverted  batch  unit  &o  remove  the  heavy 
constituents;  or  both  operations  can  be  combinec  -  simultaneously  tak- 
ing the  lighter  components  overhead  and  the  he&  Wer  components  out 
the  bottom  with  the  reservoir  between.  This  combined  operation  can 
save  heat  but  requires  additional  equipment. 

Finite  Reflux  Ratio.  It  was  pointed  out  on  page  384  that  this  case 
was  difficult  because,  if  the  plate-to-plate  calculations  were  carried 
down  from  the  top  for  an  assumed  overhead  product,  the  still  compo- 
sition would  probably  not  correspond  to  any  liquid  composition 
encountered  in  the  actual  distillation.  Starting  with  a  given  liquid 
composition  in  the  still,  it  is  possible  to  make  calculations  for  an 
assumed  overhead  composition.  If  the  calculated  overhead  composi- 
tion checks  the  assumed  value,  this  gives  one  set  of  corresponding 
distillate  and  still  compositions.  Thus  a  laborious  trial-and-error  pro- 
cedure is  necessary  to  obtain  this  one  set  of  values,  but  the  design  cal- 
culation requires  a  number  of  such  sets  for  liquid  compositions  that 
follow  a  definite  pattern.  Thus  the  overhead  vapor  corresponding  to 
the  original  charge  could  be  calculated  by  the  above  trial-and-error 
procedure  and  a  small  increment  of  this  composition  removed,  leaving 
a  new  liquid  composition.  By  a  large  number  of  such  small  steps,  the 
distillation  curve  can  be  established,  but  making  the  trial-and-error 
plate-to-plate  calculation  for  each  step  is  almost  prohibitive. 

The  total  reflux  analysis  of  the  preceding  section  was  relatj^ely 
simple,  and  an  approximate  method  for  finite  reflux  ratios  can  be 
developed  in  an  analogous  manner.  For  the  operating  lines  for  any 
two  components,  the  liquids  on  a  plate  are  related  to  the  vapors  from 
below  by 


1  + 


D 


O 


»»)•(£) 

CM. 


n+1 


(14-21) 


n+l 


Thus,  for  a  batch  distillation, 


'B/T 


(14-22) 


(14-23) 


BATCH  DISTILLATION 


389 


for  a  total  condenser, 


These  equations  are  the  same  as  some  of  those  presented  in  Chap. 
12,  page  341,  for  the  approximate  design  of  contiguous  columns. 
Using  an  average  value  of  a/j3  gives 


(14-25) 


For  the  average  value  it  is  possible  to  use  either 


'a\  GLT  +  (CLL/PL) 

2 


or  the  ratio  of  (a^/^v)  calculated  as 

OiT   + 


ftnr    = 


Equation  (14-25)  is  analogous  to  Eq.  (14-18)  and  the  resulting 
integrated  form  is  equivalent  to  Eq.  (14-20)  except  that  a  is  replaced 
by  (a/ft). 

The  evaluation  of  £L  requires  a  trial-and-error  procedure  that  can 
be  best  illustrated  by  the  following  example. 

Batch  Rectification  of  Multicomponent  Mixture  at  Finite  Reflux  Ratio.  The 
example  given  on  page  385  will  be  repeated  for  a  constant  reflux  ratio,  O/D,  equal 
to  5.0.  The  tower  will  operate  with  a  total  condenser,  and  the  usual  simplifying 
assumptions  are  made.  The  calculation  for  the  initial  vapor  overhead  is  given  in 
Table  14-4. 

TABLE  14-4 


0L  (rela- 

Comp. 

Assumed 

XD 

tive  to 
compo- 

a (relative  to 
component  D) 

/3av 

(a. 

G)~ 

XD  oalo 

nent  D) 

A 

0  85 

1  68 

8 

1  34 

5  96 

315 

0.851 

B 

0.14 

1.112 

4 

1.06 

3.78 

50.8 

0  138 

C 

0.01 

1.008 

2 

1  00 

2.0 

4  0 

0.011 

D 

0.0 

1.0 

1 

1.00 

1.0 

0.25 

0  0006 

390 


FRACTIONAL  DISTILLATION 


The  second  column  Table  14-4  gives  the  assumed  values  of  the  overhead  com- 
position, and  the  third  column  is  the  PL  relative  to  component  D  calculated  from 
the  assumed  XD  values,  for  example,  for  component  A, 

,    .   D  (XA.)D 


PL  - 


1 


0 


This  is  not  the  correct  value  of  PL  which  should  be  calculated  using  XL+I  values 
instead  of  XL  values.  For  purposes  of  simplification  this  concentration  in  the 
still  was  employed. 

The  fifth  column  gives  the  values  of  /3av,  which  divided  into  the  relative  volatili- 
ties are  presented  in  the  sixth.  These  values  of  (a//3)  can  be  used  with  the  liquid 
composition  to  calculate  the  distillate  concentrations.  The  procedure  used  in 
column  seven  is  the  same  as  Eq.  (9-3).  The  individual  values  of  (<*/|8)4  XL  are 
divided  by  370.05  »  S(a/0)4  XL  to  give  XD.  The  calculated  values  are  close  enough 
to  the  assumed  value,  and  no  retrial  is  necessary  < 

Equation  (14-20)  using  (a//3)av  instead  of  a  cannot  be  applied  over  a  wide  range 
due  to  the  variation  of  the  exponent  as  the  distillation  proceeds,  but  it  can  be  used 
over  a  limited  region  averaging  the  values  of  (a//3)av  for  the  two  ends  of  the  range. 
For  example,  assume  that  the  first  increment  is  to  remove  40  per  cent  of  the  A,  i.e., 
at  the  end  of  the  step,  FA  =•  A/A0  «  0.6.  In  order  to  carry  out  the  calculations, 
the  values  of  (<x/£)av  corresponding  to  this  final  condition  are  assumed.  These 
calculations  are  summarized  in  Table  14-5. 

TABLE  14-5 


Comp. 

As- 
sumed 

(<*/0)av 

F 

Mols  re- 
maining 
in  still 
-  Q.25F 

XL 

(D> 

XD 

PL 

/3av 

(*} 

V/Vav 

A 

5.5 

0.6 

0.15 

0.171 

155.5 

0  762 

1  89 

1.445 

5.53 

B 

3.6 

0.915 

0.229 

0.261 

43.8 

0.215 

1  165 

1.083 

3.68 

C 

1.95 

0.993 

0.248 

0.283 

4.1 

0.020 

1.014 

1.007 

1.98 

D 

1.0 

1.0 

0  25 

0.285 

0.285 

0.001 

1.001 

1.0 

1.0 

0.877 

203.69 

The  assumed  values  of  (<*/j8)av  are  given  in  the  second  column,  and  averaging 
these  with  those  of  Table  14-4  gives  values  that  are  used  to  calculate  the  values  of 
F  by  Eq.  (14-20).  At  the  start  of  the  distillation  (a/0)av  for  component  A  was 
5.96,  and  the  assumed  value  is  5.5  giving  an  arithmetic  value  of  5.73  over  the  incre- 
ment. The  corresponding  value  for  component  B  is  3.69  and 


•f*  »  0.915 


BATCH  DISTILLATION 


391 


A  similar  procedure  is  employed  for  the  other  components.  The  composition 
of  the  liquid  in  the  still  is  calculated  fronn  the  values  of  the  fractions  unvaporized, 
F,  and  is  given  in  the  fifth  column  of  the  table.  The  following  column  contains 
the  values  of  (a/j9)av  of  the  second  column  raised  to  the  fourth  power  and  multi- 
plied by  the  mol  fraction  in  the  liquid.  The  XD,  PL,  0av,  and  («/0)av  values  are 
calculated  in  the  same  manner  as  in  Table  14-4.  The  calculated  values  of  («/0)av 
are  in  good  agreement  with  the  assumed  values.  The  calculations  for  the  next 
increment  are  made  in  a  similar  manner  using  F'  as  the  fraction  unvaporized  for 
the  increment.  The  results  are  given  in  Table  14-6. 

TABLE  14-6 


Comp. 

As- 
sumed 

(«/W.T 

F' 

F 

Mols  re- 
maining 
in  still 
-  0.25F 

XL 

ex.- 

XD 

PL 

&w 

(*) 

W»v 

A 

5.0 

0.5 

0  3 

0.075 

0.098 

60.6 

0  58 

2.19 

1.595 

5.01 

B 

3.5 

0.86 

0.786 

0  197 

0  257 

38.9 

0.372 

1.29 

1.145 

3.49 

C 

1.97 

0  986 

0.98 

0  245 

0  320 

4.77 

0  0456 

1.028 

1.014 

1.97 

D 

1.0 

0.999 

0  999 

0  25 

0  325 

0.325 

0  003 

1  002 

1.001 

1.0 

0.767 

In  this  case  the  values  of  («/0)av  were  assumed,  and  F'  for  A  was  taken  as  0.5. 
This  latter  value  indicates  that  the  increment  is  to  reduce  the  mols  of  A  in  the  still 
left  after  the  step  of  Table  14-5  by  one-half.  The  values  of  F'  for  the  other  com- 
ponents were  calculated  from  F^  ==  0.5  and  the  average  of  the  assumed  value  of 
(a//J)av  of  Table  14-6  and  the  calculated  values  given  in  the  last  column  of  Table 
14-5.  The  values  of  F  were  obtained  by  multiplying  F'  by  the  corresponding  F 
value  of  the  preceding  table.  The  rest  of  the  calculations  are  made  in  the  same 
manner  as  t}iose  of  Table  14-5. 

The  values  of  yr  calculated  by  this  method  are  plotted  in  Fig.  14-7  as  a  function 
of  mol  fraction  of  the  original  charge  distilled.  The  total  reflux  results  of  Fig. 
14-5  are  also  included  in  the  dotted  curves  for  comparison.  As  would  be  expected, 
the  separation  with  the  finite  reflux  ratio  is  not  so  sharp  as  for  total  reflux.  This 
decrease  in  the  degree  of  separation  is  particularly  marked  for  the  intermediate 
fraction  and  is  least  serious  in  the  case  of  the  least  volatile  fraction.  The  inverted 
and  split-towers  systems  discussed  for  total  reflux  would  be  effective  in  increasing 
the  purity  of  any  given  fraction. 

This  approximate  method  is  most  accurate  for  high  values  of  (0/D)  and  rela- 
tively few  total  plates.  It  reduces  to  the  total  reflux  method  for  (0/D)  «•  *>. 
When  the  number  of  plates  is  large,  they  tend  to  pinch  at  the  equilibrium  curve 
and  the  arithmetic  average  for  ft  is  not  satisfactory. 

Finite  Reflux  Ratio  with  Column  Holdup.  The  conditions  in  this 
case  are  similar  to  those  discussed  on  page  380  for  binary  mixtures. 
The  operating  lines  for  each  component  are  curved  except  for  total 


392 


FRACTIONAL  DISTILLATION 


reflux,  and  it  is  very  difficult  to  evaluate  the  position  of  these  lines. 
The  total  reflux  relations  given  on  page  382  for  a  binary  mixture  do  not 
apply  exactly  in  this  case  because  it  was  assumed  that  the  unit  was 
operating  with  a  variable  reflux  ratio,  producing  a  constant  overhead 
composition.  In  the  case  of  a  multicomponent  mixture,  it  is  generally 


08 


t.O 


02  0.4  0.6 

Mol  fraction  distilled 
FIG.  14-7.    Fractionation  curve  for  multicomponent  mixture  at  finite  reflux  ratio. 

impossible  to  keep  the  overhead  composition  constant  for  all  com- 
ponents as  product  is  withdrawn.  In  certain  cases,  the  distillate  may 
be  reasonably  constant,  and  a  similar  analysis  can  be  made. 

The  design  methods  for  batch  distillation  allowing  for  liquid  holdup 
in  the  column  are  very  unsatisfactory,  and  it  is  a  field  that  should  be 
actively  studied  in  view  of  the  importance  of  the  operation.  Improve- 
ments in  the  design  calculations  for  multicomponent  mixtures  with  no 
liquid  holdup  in  the  column  are  also  needed. 

References 

1.  COLBUKN  and  STERN,  Trans.  Am.  Inst.  Chem.  Engrs.,  37,  291  (1941). 

2.  EDGEWORTH-JOHNSTONE,  Ind.  Eng.  Chem.,  35, 407  (1943);  36,  482, 1068  (1944). 

3.  ROSE,  WELSHANS,  and  LONG,  Ind.  Eng.  Chem.,  32,  673  (1940). 


CHAPTER  15 
VACUUM  DISTILLATION 

Reduced  pressures  are  frequently  used  in  distillation  processes  for 
the  purpose  of  lowering  the  temperature  required.  This  is  frequently 
important  in  the  distillation  of  organic  substances  that  are  thermally 
unstable  and  would  decompose  if  boiled  at  normal  pressures.  In  addi- 
tion to  the  reduced  thermal  degradation,  the  lower  temperature  often 
modifies  the  relative  volatility  or  the  degree  of  separation  involved. 

In  the  design  methods  considered  in  the  preceding  chapters,  vapor- 
liquid  equilibrium  was  used  as  a  basic  criterion.  The  equilibrium  con- 
dition between  a  vapor  and  a  liquid  is  a  dynamic  condition  in  which 
equal  numbers  of  molecules  of  each  species  enter  and  leave  the  liquid 
per  unit  time.  Langmuir  (Ref.  5)  derived  an  expression  for  the  rate 
of  this  interchange  based  on  kinetic  theory  considerations.  The  num- 
ber of  molecules  striking  a  unit  of  surface  per  unit  time  for  a  perfect 
gas  is 

Pu  /i  c    1  \ 

(15-1) 


where  P  =  pressure 

u  =  average  velocity  of  molecules 
E  =  gas-law  constant 
T  =  temperature 

n  =  number  of  molecules  striking  a  unit  area  per  unit  time 
The  arithmetical  average  velocity  of  the  molecules  is 

(15-2) 

where  M  is  the  molecular  weight,  and  the  mass,  m,  striking  a  unit  area 
per  unit  time  is 


Langmuir  has  investigated  the  fraction  of  the  molecules  striking  a  sur- 
face that  rebound,  and  his  results  would  indicate  that  Essentially  all 
the  molecules  striking  the  surface  would  enter  and  not  be  reflected. 


394  FRACTIONAL  DISTILLATION 

Assuming  that  none  are  reflected,  Eq.  (15-2)  can  be  used  to  calculate 
the.  rate  of  evaporation  from  a  liquid  at  equilibrium  with  a  vapor. 
The  same  relation  has  been  used  to  predict  the  rate  of  evaporation 
from  a  liquid  even  when  the  vapor  is  not  in  equilibrium  with  a  liquid. 
This  may  be  approximately  true  for  the  nonequilibrium  case,  but  there  is 
undoubtedly  some  interference  of  the  molecules  in  the  vapor  with  those 
evaporating,  and  the  use  of  Eq.  (15-3)  to  calculate  the  absolute  rate  of 
evaporation  probably  gives  results  that  are  somewhat  low.  For  the 
purposes  of  this  discussion,  Eq.  (15-3)  will  be  used  as  the  basis  of  esti- 
mating the  absolute  evaporation  rate  from  a  liquid. 

A  consideration  of  Eq.  (15-3)  indicates  that  the  vapor  pressure  is  the 
most  important  factor  in  determining  the  rate  of  evaporation.  Molecu- 
lar weight  and  temperature  are  of  less  importance.  The  same  reason- 
ing that  was  used  to  develop  this  equation  can  be  applied  to  each  com- 
ponent in  a  mixture,  giving  for  component  A, 


It  is  instructive  to  compare  this  absolute  rate  of  evaporation  with 
the  rate  of  mass  transfer  obtained  in  an  atmospheric  pressure  distilla- 
tion of  benzene  and  toluene.  Assume  that  the  liquid  phase  is  equi- 
molal  in  benzene  and  toluene  and  that  a  vapor  bubble  %  in.  in  diam- 
eter changes  10  mol  per  cent  in  composition  for  a  0.1-sec.  contact  with 
the  liquid.  This  liquid  composition  corresponds  to  an  equilibrium 
temperature  of  92.4°C.  and  an  equilibrium  vapor  composition  of  0.71 
mol  fraction  benzene.  By  Eq.  (15-3)  the  rate  of  evaporation  of  ben- 
zene is 

78 


/ 2(3.1416)  (8.316  X  107)  (365.4) 
=  14.5  g.  per  sq.  cm.  per  sec. 
=  53.5  tons  per  sq.  ft.  per  hr. 

A  similar  calculation  for  toluene  gives  6.4  g.  per  sq.  cm.  per  sec. 
In  the  actual  experiment  it  is  assumed  that  the  exchange  is  equimolar, 
and  the  rate  of  exchange  is 


m' 


r/2.54V.ir         273         1/O.A 
L\  2  /  6  J  [22,400(365.4)  JVO.l/ 

••  3.5  X  10~5  g.  mols  per  sq.  cm.  per  sec. 

e  2.7  X  10~8  g.  benzene  per  sq.  cm.  per  sec. 

o  3.2  X  10~8  g.  toluene  per  sq.  cm.  per  sec. 


VACUUM  DISTILLATION  395 

The  actual  rates  of  mass  transfer  are  about  0.03  per  cent  of  the  theo- 
retical evaporation  rates.  Under  the  conditions  of  mass  transfer,  the 
true  vapor-liquid  equilibria  do  not  exist  at  the  interface  because  mole- 
cules are  leaving  the  liquid  phase  at  a  different  rate  than  they  are 
returning.  In  the  case  of  benzene,  more  molecules  are  leaving  than 
return  while  for  toluene  the  reverse  is  true.  In  the  case  just  con- 
sidered, the  net  removal  from  the  interface  is  so  small  in  comparison 
to  the  interchange  rate  that  equilibrium  should  be  closely  attained  at 
the  interface.  If  the  net  removal  is  made  a  large  percentage  of  the 
interchange  rate,  normal  vapor-liquid  equilibrium  will  not  be  obtained. 
In  the  extreme  case  all  the  molecules  that  evaporate  could  be  removed, 
and  the  relative  rate  for  two  components  would  be 

(        . 

(         } 

If  an  equilibrium  vapor  were  removed,  the  ratio  of  the  components 
would  be  PA/PB,  and  the  relative  evaporation  rate  differs  by  the  molecu- 
lar weight  term.  Thus,  the  composition  of  an  equilibrium  vapor  and 
the  one  obtained  by  removing  all  the  molecules  that  evaporate  will 
generally  be  different  unless  the  molecular  weights  are  the  same.  It 
should  be  possible  in  some  cases  to  separate  normal  azeotropic  mix- 
tures by  the  evaporative  technique,  while  other  mixtures  that  would 
give  no  separation  by  this  method  could  be  handled  by  equilibrium 
vaporization.  The  azeotrope  for  the  system  ethanol-water  should 
give  a  vapor  by  the  evaporative  method  of  considerably  different  com- 
position than  the  liquid  due  to  the  high  ratio  of  the  molecular  weights. 

It  should  be  possible  to  obtain  a  vapor  composition  anywhere 
between  the  true  equilibrium  value  and  that  given  by  Eq.  (15-5)  by 
adjusting  the  net  removal  rate  relative  to  the  evaporation  rate.  A 
system  with  a  net  removal  rate  equal  to  the  evaporation  rate  has  been 
termed  "molecular  distillation"  by  Fawcett  (Ref.  2)  and  "unob- 
structed-path distillation"  by  Hickman  (Ref.  3).  Hickman  has  used 
the  term  "molecular  distillation"  for  systems  having  mean  free  path 
for  the  vapor  molecules  comparable  to  the  distance  between  the  evapo- 
rating and  condensing  surfaces.  In  this  text  the  term  molecular  dis- 
tillation will  be  applied  to  those  cases  in  which  the  net  removal  rate  is 
a  high  percentage  of  the  absolute  evaporation  rate. 

The  benzene-toluene  example  gave  low  rates  of  transfer  relative  to 
the  absolute  evaporation  rate,  because  (1)  the  interchange  process 
introduces  diffusional  resistance  and  (2)  the  vapor  approaches  equilib- 


396  FRACTIONAL  DISTILLATION 

rium  with  the  liquid  thereby  decreasing  the  net  rate  of  transfer.  In 
order  to  approach  molecular  distillation  conditions,  it  is  necessary  to 
increase  the  net  rate  at  which  molecules  are  removed  relative  to  the 
rate  at  which  they  evaporate.  This  can  be  accomplished  by  increasing 
the  removal  rate  and  decreasing  the  evaporation  rate. 

For  either  type  of  vaporization,  the  general  consideration  of  solution 
laws  apply  and  can  be  used  to  predict  the  results  of  modifying  the 
liquid  phase.  Thus  it  would  be  possible  to  modify  the  composition  of 
the  vapor  removed  in  molecular  distillation  just  as  in  azeotropic  or 
extractive  distillation. 

VACUUM  AND  STEAM  DISTILLATION 

For  pressures  down  to  about  1  mm.  Hg  abs.,  the  distillation  opera- 
tion can  be  carried  out  in  a  manner  similar  to  those  at  higher  pressures, 
and  the  problems  relate  to  reducing  the  pressure  drop  for  vapor  flow 
through  conventional  equipment.  The  pressure  drop  of  bubble-cap 
plates  can  be  reduced  to  the  order  of  2  mm.  Hg  per  plate  (see  page 
404),  and  special  spray-type  plates  can  give  pressure  drops  as  low  as 
0.5  mm.  Hg  per  plate.  With  such  contacting  units  it  is  obvious  that 
even  a  few  plates  will  necessitate  a  still  pressure  of  several  millimeters 
even  with  a  high  vacuum  at  the  condenser.  Packed  towers  can  give 
low  pressure  drops,  and  still  pressures  of  the  order  of  5  to  10  mm.  Hg  are 
obtainable  with  reasonable  tower  sizes  and  rates  of  distillation.  How- 
ever, owing  to  the  poor  vapor-liquid  contact,  they  are  not  widely  used 
for  such  operations.  The  contact  is  particularly  poor  in  this  case 
because  of  the  low  volumetric  ratio  of  liquid  to  vapor. 

Lower  distillation  temperatures  can  be  obtained  by  the  use  of  steam. 
In  this  case  the  steam  does  not  usually  condense  in  the  tower,  and  the 
plates  contain  only  the  high  boiling  organic  material.  The  steam  acts 
as  an  inert  carrier  that  is  easily  condensed  and  does  not  have  to  pass 
through  the  vacuum  pump.  In  some  cases  the  temperature  gradient 
in  the  distillation  column  may  be  so  great  that  the  steam  will  con- 
dense in  the  upper  portions,  and  it  should  be  withdrawn.  If  too  much 
water  collects  on  a  plate,  it  may  seriously  interfere  with  the  fractiona- 
tion;  if  the  water  runs  down  the  tower,  it  will  vaporize  on  the  lower 
plates  and  this  steam  recycle  in  the  tower  can  overload  the  unit  and 
seriously  interfere  with  the  fractionation.  Theoretically  it  is  possible 
to  fractionate  a  material  with  a  very  low  vapor  pressure  by  the  use  of 
steam  distillation,  but  the  steam  consumption  increases  as  the  vapor 
pressure  on  the  component  decreases.  Reduction  of  the  total  pressure 
reduces  the  steam  consumption,  but  if  the  vapor  pressure  of  the  com- 


VACUUM  DISTILLATION  397 

ponent  at  the  distillation  temperature  becomes  less  than  0.1  mm.  Hg, 
the  steam  consumption  becomes  excessive  for  most  cases.  The  rectifi- 
cation calculation  for  such  distillations  can  be  made  in  the  usual  plate- 
to-plate  manner. 

The  vapor-liquid  equilibria  at  pressures  doWn  to  1  mm.  Hg  are  not 
greatly  different  from  those  at  higher  pressure.  The  relative  volatility 
of  a  binary  system  may  either  increase  or  decrease  as  the  pressure  is 
reduced.  For  example,  in  a  mixture  of  oleic  and  stearic  acids,  oleic  is 
the  more  volatile  at  temperatures  above  100  to  110°C.,  while  below 
these  temperatures  it  is  the  less  volatile.  For  mixtures  that  obey 
Raoult's  law  the  relative  volatility  generally  increases  as  the  tempera- 
ture is  decreased  because  the  less  volatile  constituent  has  the  higher 
latent  heat  resulting  in  a  high  temperature  coefficient  of  vapor  pressure. 

The  calculations  of  the  separation  to  be  expected  as  a  function  of 
the  reflux  ratio  and  the  number  of  theoretical  plates  for  vacuum  dis- 
tillation are  completely  analogous  to  those  for  higher  pressure  opera- 
tion. The  difficult  design  problems  are  those  related  to  obtaining 
efficient  contact  between  the  liquid  and  vapor  with  the  low  pressure 
drops  available. 

MOLECULAR  DISTILLATION 

This  type  of  operation  has  been  applied  to  the  distillation  of  mate- 
rials that  have  very  low  vapor  pressures  at  the  maximum  operating 
temperature.  The  available  pressure  drops  in  such  cases  would  be  too 
low  to  obtain  practical  production  rates  in  conventional  equipment, 
but  by  operating  such  that  the  rate  of  distillation  is  approximately 
equal  to  the  absolute  evaporation  rate  of  the  liquid  reasonable  capaci- 
ties can  be  obtained.  The  most  common  method  of  obtaining  the 
molecular  distillation  conditions  is  to  carry  out  the  operation  at  a  high 
vacuum  (0.01  mm.  Hg  or  less)  and  to  place  the  condensing  surface  so 
that  it  is  parallel  to  the  evaporating  surface  and  in  close  proximity  to 
it.  The  condenser  is  operated  at  a  low  temperature  to  limit  the 
reevaporation.  In  order  to  obtain  satisfactory  absolute  rates  of  evapo- 
ration, it  has  been  found  that  as  an  approximate  rule  the  temperature 
should  not  be  lower  than  100°C.  below  the  temperature  at  which 
the  vapor  pressure  of  the  substance  being  evaporated  is  1  to  5  mm. 
Hg  abs. 

Even  with  molecular  distillation,  the  rates  of  evaporation  obtained 
are  low  when  the  vapor  pressure  is  less  than  0.01  mm.  Hg  abs.  Thus 
for  a  material  having  a  molecular  weight  of  400  and  a  vapor  pressure  of 
0.01  mm.  Hg  at  100°C.,  the  absolute  evaporation  rate  by  Eq.  (15-3) 


398  FRACTIONAL  DISTILLATION 

would  be  only  0.02  X  10~*  g.  per  sec.  per  sq.  cm.  This  is  much  lower 
than  the  rate  estimated  on  page  394  for  normal  vapor-liquid  inter- 
change. High-molecular-weight  polymers  would  have  low  evapora- 
tion rates  regardless  of  the  vacuum. 

It  should  be  possible  to  obtain  results  similar  to  molecular  distillation 
at  higher  total  pressures  by  a  high  degree  of  turbulence  in  the  space 
between  the  condenser  and  the  evaporating  liquid  in  order  to  obtain 
rapid  mass  transfer. 

In  the  high-vacuum  method  of  operating,  it  is  usually  suggested  that 
the  distance  between  the  condenser  and  the  evaporating  surface  should 
be  of  the  order  of  the  mean  free  path  of  the  molecules  in  the  vapor. 
Jeans  (Ref.  4)  gives  the  following  equation  for  the  mean  free  path 
(M.F.P.)  of  a  molecule: 


where  M.F.P.  =  mean  free  path,  cm. 

p  =  molal  density  =  molecules,  cu.  cm. 
B  1.75  X  1019(P/!T)  for  a  perfect  gas 
d  »  diameter  of  molecule,  cm.     (As  an  approximate  rule 
use  cube  root  of  6/rr  times  the  liquid  volume  per 
molecule) 

P  =  absolute  pressure,  mm.  Hg 
T  ~  absolute  temperature,  °R. 

Thus  for  a  material  that  has  a  molecular  weight  of  500  and  a  liquid 
density  of  0.9,  the  mean  free  path  at  a  pressure  of  10~3  mm.  Hg  and 
400°F.  would  be 


1023)T 
»  0.77  cm. 

However,  it  does  not  appear  to  be  necessary  to  make  the  mean  free 
path  as  large  as  the  distance  between  the  condenser  and  the  evaporat- 
ing surface  to  obtain  molecular  distillation  conditions.  Bronsted  and 
Hevesy  (Ref.  1)  obtained  separations  of  mercury  isotopes  that  cor- 
responded closely  to  molecular  distillation  rates  under  conditions  where 
the  condenser  was  separated  from  the  evaporating  mercury  surface  by 
a  distance  approximately  100  times  the  mean  free  path.  Taylor  (Ref, 
7)  distilling  petroleum  fractions  found  the  rate  of  evaporation  to  be 
independent  of  the  total  pressure  over  a  range  corresponding  to  mean 


VACUUM  DISTILLATION  399 

free  paths  of  0.01  to  10  times  the  clearance  between  the  condenser  and 
the  evaporating  surface.  He  also  found  that  noncondensable  residual 
gas  at  pressures  up  to  the  vapor  pressure  of  the  liquid  being  distilled  did 
not  materially  lower  the  distillation  rate.  Higher  residual  gas  pressure 
can  cause  appreciable  lowering  of  the  rate. 

The  most  common  type  of  high  vacuum  molecular  distillation  still  is 
the  vertical-tube  falling-film  unit,  a  schematic  diagram  of  which  is  shown 
in  Fig.  15-1.  The  liquid  to  be  distilled  is  first  degassed.  This  is  essen- 
tial if  splashing  in  the  distillation  unit  is  to  be  avoided.  This  liquid 
then  flows  down  in  a  film  on  the  outside  of  the  inner  tube  which  is 
internally  heated.  The  inner  surface  of  the  outer  tube  is  the  condenser 
which  can  be  air-  or  water-cooled.  For  a  high  rate  of  distillation,  the 
clearance  between  the  two  surfaces  should  be  relatively  small,  but  if 
they  are  too  close,  any  noncondensable  gas  released  at  the  bottom  of 
the  still  will  have  difficulty  flowing  out  of  the  unit.  A  clearance  of  6.4 
to  1.0  in.  appears  to  be  about  optimum  for  a  unit  2  to  4  ft.  long. 

In  such  a  falling-film  unit  a  molecule  moving  from  the  evaporating 
liquid  to  the  condenser  encounters  a  number  of  resistances:  (1)  diffu- 
sional  resistance  from  the  interior  to  the  surface  of  the  liquid,  (2) 
evaporational  resistance,  (3)  resistance  to  transfer  in  the  vapor,  (4) 
resistances  in  condensation. 

The  resistance  to  condensation  is  small,  and  the  resistance  to  vapor 
transfer  is  made  small  by  the  use  of  low  pressure  and  by  keeping  the 
condenser  close  to  the  evaporating  surface.  The  evaporating  rate  of 
the  molecules  in  the  surface  is  chiefly  a  function  of  temperature  which 
should  be  kept  as  high  as  possible  without  thermal  degradation  or 
bubbling  of  the  liquid  which  throws  unvaporized  material  over  to  the 
condenser.  In  most  cases,  the  limiting  rate  is  diffusion  in  the  liquid 
phase.  Owing  to  the  large  size  of  the  molecules  and  the  viscosity  of 
the  liquid,  the  mass-transfer  rate  is  very  low.  The  outer  surface  of  the 
liquid  is  depleted  of  the  more  rapidly  evaporating  molecules  a  short 
distance  from  the  top  of  the  unit,  and  the  surface  then  has  a  higher  con- 
centration of  the  less  volatile  molecules  than  the  average  composition 
of  the  liquid.  This  reduces  both  the  evaporation  rate  and  the  degree 
of  separation  obtained.  A  small  amount  of  large-size,  essentially  non- 
volatile, material  in  the  liquid  can  give  a  serious  blocking  of  the  surface. 
Owing  to  this  effect,  increasing  the  length  of  the  apparatus  does  not 
give  a  proportional  increase  in  the  amount  of  evaporation.  For  this 
reason,  the  falling-film  units  are  seldom  made  over  2  to  4  ft.  high. 
Several  methods  can  be  used  to  reduce  the  effect  of  this  surface  block- 
ing: (1)  Various  mechanical  devices  have  been  proposed  to  cause  mix- 


400 


FRACTIONAL  DISTILLATION 
V 


Vacuum 
system 


To  switch 
board 


to  aspirator  and 
pressure  control 

"f 


Pressure 

measuring 

equipment 


Oistilhte 
receiver 
FIG.  15-1.     Diagram  of  falling-film  unit. 


ing  of  the  falling  film.  (2)  The  liquid  circulation  rate  can  be  increased, 
resulting  in  a  lower  percentage  evaporation  per  pass  through  the  unit. 
A.  high  percentage  evaporation  can  be  obtained  either  by  recirculating 
the  liquid  or  by  using  several  units  in  series  with  mixing  between  each 
unit.  (3)  A  high  liquid  flow  rate  can  be  used  to  cause  the  outer  por- 
tion of  the  falling  film  to  be  in  turbulent  rather  than  laminar  flow. 
This  type  of  operation  may  also  necessitate  recirculation  or  series 


VACUUM  DISTILLATION  401 

operation  to  obtain  a  high  percentage  of  the  liquid  distilled.  (4)  The 
unit  may  be  modified  to  give  a  continually  increasing  surface.  For 
example,  the  inner  tube  can  be  made  conical  with  the  small  end  at  the 
top.  As  the  liquid  flows  down,  it  must  increase  in  area  which  will 
present  fresh  surface  for  evaporation.  Hickman  (Ref.  3)  has  devel- 
oped a  spinning-plate  type  of  film  unit  in  which  the  liquid  is  fed  at  the 
center  and  flows  across  the  spinning  plate  due  to  centrifugal  force. 
The  plate  is  heated,  and  the  condenser  is  placed  parallel  to  it.  Because 
of  the  high  centrifugal  force,  it  is  possible  to  obtain  thin  films  which 
mean  large  surface  area  per  unit  volume  of  liquid,  and  the  increasing 
diameter  of  the  plate  requires  the  formation  of  new  surface  as  the 
liquid  flows  outward.  The  spinning-plate  unit  is  effective  for  the  pur- 
pose of  increasing  evaporation  rate,  but  other  methods  would  appear 
to  be  simpler  for  large-scale  units. 

Besides  keeping  the  temperature  at  a  low  level,  molecular  distilla- 
tion holds  only  a  small  volume  of  the  liquid  at  the  evaporation  tempera- 
ture and  thereby  reduces  the  thermal  degradation.  The  spinning- 
plate  type  of  still  is  particularly  effective  in  this  respect  because  of  the 
thin  film  obtained. 

The  thermal  efficiency  of  a  molecular  distillation  is  low.  Fawcett 
(Ref .  2)  has  given  a  heat  balance  on  a  unit  distilling  triolein  at  240°C. 
with  the  condenser  at  25°C.  The  data  are  summarized  in  the  follow- 
ing table: 

Per  Cent 

Preheat  of  liquid . .  .8 

Radiation ...  ...   59 

Latent  heat  of  evaporation  9 

Conduction ...  24 

Only  17  per  cent  of  the  total  heat  is  usefully  employed,  the  other  83 
per  cent  is  lost  by  heat  transfer.  The  loss  by  radiation  could  be 
reduced  somewhat  by  increasing  the  temperature  of  the  condenser,  but 
this  might  reduce  the  effectiveness  of  the  distillation. 

Another  drawback  to  molecular  distillation  is  the  fact  that  an  effec- 
tive rectification  system  has  not  been  developed.  Greater  separations 
than  are  equivalent  to  a  single  distillation  stage  have  been  obtained  by 
repeated  distillations  in  the  manner  described  on  page  102,  but  they 
are  tedious  and  difficult  to  perform. 

Schaffner,  Bowman,  and  Coull  (Ref.  6)  have  described  a  vertical 
falling-film  multiple  distillation  column  that  can  be  employed  for  frac- 
tionation  under  vacuum.  The  wall  is  made  up  of  short  sections  with 


402  FRACTIONAL 

alternate  sections  being  heated  and  cooled.  Vapor  condenses  on  the 
cooled  section  and  flows  down  to  the  heated  section  below  where  it  is 
partially  vaporized,  and  the  action  is  repeated  on  succeeding  sections. 
By  this  series  of  partial  condensations  and  vaporizations,  an  enrich- 
ment of  the  vapor  in  the  more  volatile  component  is  obtained.  Because 
vapor  must  flow  between  sections,  molecular-type  distillation  is  not 
possible.  Such  a  unit  is  very  sensitive  to  operating  conditions,  and 
the  heat  added  and  removed  in  the  successive  sections  must  be  well 
balanced  or  the  vapor  will  all  condense  or  the  liquid  will  all  vaporize. 
Owing  to  the  temperature  changes  during  distillation,  the  heat  supplies 
will  need  frequent  readjustments  to  obtain  optimum  operation. 

The  very  low  pressures  involved  almost  preclude  an  effective  contact 
similar  to  that  obtained  in  normal  rectification.  It  would  appear  that 
the  use  of  higher  pressures  with  a  high  degree  of  turbulence  to  obtain 
the  necessary  rate  of  distillation  at  the  low  temperatures  involved 
would  offer  better  possibilities  for  rectification  than  the  use  of  high 
vacuum.  By  the  use  of  a  low  molecular  weight  gas,  such  as  hydrogen, 
to  maintain  the  pressure,  it  should  be  possible  to  obtain  high  transfer 
rate  with  a  moderate  degree  of  gas  turbulence. 

Because  of  the  high  cost  per  unit  of  product  distilled,  the  use  of 
molecular  distillation  has  been  limited  in  its  application  to  the  separa- 
tion of  relatively  expensive  materials  that  are  sensitive  to  thermal 
degradation. 

References 

1.  BRdNSTED  and  HEVESY,  Phil  Mag.  (VI),  43,  3  (1922). 

2.  FAWCETT,  /.  Soc.  Chem.  Ind.,  58,  43  (1939). 

3.  HICKMAN,  Chem.  Rev.,  34,  51  (1944). 

4.  JEANS,  "An  Introduction  to  the  Kinetic  Theory  of  Gases,"  pp.  131-155,  The 
Macmillan  Company,  New  York,  1940. 

5.  LANGMTJIB,  Phys.  Rev.,  8, 149  (1916). 

6.  SCHAFFNER,  BOWMAN,  and  COULL,  Trans.  Am.  Inst.  Chem.  Engrs.j  39, 77  (1943) . 

7.  TAYLOR,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1938. 


CHAPTER  16 
FRACTIONATING  COLUMN  DESIGN 

The  design  calculations  given  in  the  previous  chapters  involved  the 
mathematical  problem  of  determining  the  number  of  theoretical  plates 
and  did  not  consider  whether  or  not  such  performance  could  be 
obtained.  This  chapter  will  consider  the  mechanical  design  problems 
encountered  in  attempting  to  attain  the  desired  degree  of  contact 
between  the  vapor  and  liquid  and  the  necessary  fluid  and  vapor  flow. 
The  fractionating  column  must  bring  the  liquid  and  vapor  into  counter- 
current  contact  (or  some  approximation  of  this  type  of  flow),  and  it 
must  be  constructed  to  furnish  the  necessary  pressure  drops  (or  liquid 
heads)  to  give  the  desired  flow  conditions. 

The  liquid  gradients  and  the  pressure  drops  encountered  by  the 
liquid  and  the  vapor  are  important  for  several  reasons: 

1.  The  conditions  may  be  such  that  the  liquid  will  not  flow  down  the 
column,  countercurrent  to  the  upflow  of  the  vapor.    Such  action  pre- 
venting rectification  is  termed  "flooding"  or  "priming."    Any  column 
has  a  maximum  operating  capacity,  above  which  this  condition  will  be 
encountered,  but  the  capacity  of  a  column  of  given  diameter  with  fixed 
plate  spacing  will  be  a  function  of  the  design. 

2.  The  contact  between  the  vapor  and  liquid  is  a  function  of  the 
liquid  gradients  and  the  gas  pressure  drops. 

3.  The  pressure  drop  per  plate  is  of  vital  importance  in  vacuum 
rectification. 

BUBBLE-CAP  PLATES 

The  bubble-cap  plate  is  the  most  common  vapor-liquid  contacting 
device  employed  for  fractional  distillation.  Its  purpose  is  to  bring  the 
vapor  and  liquid  into  intimate  contact  so  that  the  necessary  mass 
transfer  can  be  effected.  This  requires  means  for  bringing  the  liquid 
down  the  column  from  plate  to  plate,  across  the  plates,  and  into  contact 
with  the  vapor.  The  design  of  such  a  unit  involves  obtaining  the 
desired  flow  conditions  of  the  liquid  and  vapor  and  the  contact  between 
the  two.  Figure  16-1  gives  a  schematic  drawing  of  the  cross  section  of 
a  bubble-cap  plate.  The  liquid  flows  onto  a  plate  from  the  down  pipes, 
flows  across  the  plate  contacting  the  vapor,  flows  over  the  outlet  weir 

403 


404 


FRACTIONAL  DISTILLATION 


and  through  the  down  pipe  to  the  plate  below.  The  vapor  flows  up 
through  the  liquid  on  the  plate  and  to  the  space  above.  .For  these 
flows  to  follow  the  desired  pattern,  the  necessary  pressure  drops  and 

hydraulic  heads  must  be  available. 
The  liquid  meets  resistance  in 
the  down  pipes,  in  flowing  across 
the  plate,  and  in  flowing  over  the 
weirs.  The  frictional  resistance  in 
the  down  pipes  is  handled  by  mak- 
ing them  of  adequate  cross  section 
and  height  to  take  the  liquid  load. 
For  a  given  cross  section,  in  general 
the  liquid-handling  capacity  of  the 
down  pipes  will  increase  with  in- 
creasing height  but  it  is  desirable 
to  keep  the  height  low  in  order  to 
reduce  the  plate  spacing.  On  the 
other  hand,  down  pipes  of  large 
cross  section  reduce  the  available 
plate  area  for  vapor  liquid  contact.  In  flowing  across  the  plate  the 
liquid  decreases  in  depth  owing  to  the  frictional  and  kinetic  effects 
giving  the  so-called  hydraulic  gradient.  An  overflow  weir  is  employed 


FIG.     16-1.     Schematic    cross-sectional 
diagram  of  bubble-cap  plate  column. 


Tffff 


Riser  ->- 


A 


gPHrtj*^  Bottom  of  cap 


/////////77/7///7/^  ~-  Pi»  t* 


©  Vapor 

FIG.  16-2.     Cross  section  of  bubble  cap. 

to  maintain  the  liquid  level  at  approximately  the  desired  level.    These 
various  factors  are  considered  in  detail  in  later  paragraphs. 

Pressure  Drop  for  Vapor  Flow.  The  vapor  meets  its  main  resistance 
in  passing  through  the  bubble  cap  and  the  liquid  on  the  plate.  Con- 
sider the  section  of  a  cap  shown  in  Fig.  16-2.  The  vapor  from  the  plate 


FRACTIONATING  COLUMN  DESIGN  405 

below  enters  the  riser  at  (1)  and  encounters  a  pressure  drop  due  to  the 
reduction  in  cross  section.  There  is  a  frictional  drop  in  the  riser  from 
(1)  to  (2),  a  reversal  loss  (2)  to  (3)  between  the  top  of  the  riser  and  the 
cap,  and  then  a  frictional  drop  against  the  ca£  from  (3)  to  (4).  The 
vapor  then  passes  through  the  slot  and  up  to  the  vapor  space  above 
the  liquid.  It  is  convenient  to  group  these  into  three  pressure  drops: 
the  pressure  drop  through  the  riser  and  cap,  hc]  the  loss  in  pressure  in 
flowing  through  the  slots,  hs;  and  the  pressure  drop  due  to  the  liquid 
head  above  the  slots,  hL. 

Pressure  Drop  through  Risers  and  Cap.  This  loss  is  chiefly  a  kinetic 
velocity  effect  due  to  the  changing  cross-sectional  areas.  The  pressure 
drop  in  inches  of  the  liquid  equivalent  to  the  kinetic  head  is 

fc,-^*  (16.1) 

20C    PL  ^          ' 

where  ha  =  kinetic  head,  inches  of  liquid  of  density  pL 

VR  =  maximum  velocity  in  riser,  between  top  of  riser  and  cap, 

or  in  annulus  between  riser  and  cap,  f.p.s. 
gc  =  conversion  constant 

=  32.2[(ft.)(lb  force)]/[(sec.2)(lb.  mass)] 
pv  =  density  of  ga&,  same  units  as  pL 

The  actual  loss  in  pressure,  hc,  from  (1)  to  (4)  should  be  a  function  of 
hH.  The  data  of  Mayer  (Ref.  24),  Schneider  (Ref.  28),  and  Dauphin6 
(Ref  .  5)  on  several  3-,  4-,  and  6-in.-diam.eter  caps  with  risers  from  2  to  4 
in.  in  diameter  gave  ratio  of  hc/hn  from  4.7  to  6.3.  Souders  (Ref.  34) 
gave  results  indicating  a  ratio  of  2.9,  Kirkbride  (Ref.  20)  recommended 
3.2,  and  Edminster  (Ref.  7)  suggested  7.8  but  included  the  pressure 
drop  through  the  slots.  In  view  of  the  extensive  data  by  the  first  three 
investigators,  a  value  of  the  ratio  equal  to  6.0  will  be  used  giving 


hc  =  1.1F|  (16-2) 

PL 

Pressure  Drop  through  Slots.  The  pressure  drop  through  the  slot, 
h8j  is  evidenced  by  the  liquid  level  in  the  cap  being  lower  than  the  top 
of  the  slots.  The  value  of  ha  will  be  taken  as  equal  to  the  difference  in 
the  pressure  of  the  vapor  in  the  cap  at  position  (4)  and  the  liquid  out- 
side the  cap  at  the  top  of  the  slot.  The  slot  action  varies  with  the  rate 
of  flow.  At  low  rates  of  flow  an  intermittent  type  of  bubbling  action  is 
obtained.  Owing  to  the  surface  tension,  the  pressure  within  the  cap 
rises  until  the  liquid  under  the  cap  is  depressed  an  appreciable  distance 
below  the  top  of  the  slots.  When  the  pressure  is  sufficient  to  overcome 


406  FRACTIONAL  DISTILLATION 

the  surface  tension,  there  is  a  rapid  flow  of  vapor  reducing  the  pressure, 
and  the  cycle  is  then  repeated.  Thus  a  pressure  drop  across  the  plate 
greater  than  the  height  of  the  liquid  above  the  slots  is  necessary  to 
initiate  vapor  flow.  At  higher  rates  of  vapor  flow,  bubbling  becomes 
continuous,  and  the  pressure  drop  through  the  slots  becomes  greater 
than  that  necessary  to  initiate  flow.  At  still  higher  vapor  rates,  the 
vapor  blows  open  channels  through  the  liquid.  For  all  cases,  the 
pressure  drop  through  the  slots  is  greater  than  for  flow  through  the 
slots  on  a  dry  plate.  The  calculation  of  h,  is  also  complicated  because 
the  velocity  of  the  vapor  in  the  cap  approaching  the  slots  may  be  equal 
to  or  greater  than  the  slot  velocity,  and  a  simple  orifice-type  equation 
is  not  applicable. 

A  definite  value  of  h.  is  necessary  to  initiate  flow.  It  then  increases 
slowly  with  the  rate  of  vapor  flow  as  the  slot  opening  increases,  and 
when  the  slots  are  completely  open,  the  pressure  drop  increases  rapidly 
with  increase  in  vapor  rate  through  the  slots. 

For  values  of  he  less  than  the  height  of  the  slot,  the  data  of  Carey 
(Ref.  3),  Griswold  (Ref.  14),  Mayer,  Schneider,  and  Dauphin^  can  be 

correlated  by 

.  0>67 

h,  =  0.12  i  +  K  (hfV.  J— ^-)  (16-3) 

PL  \  \PL  —  pv/ 

where  h8  —  slot  opening  or  pressure  drop  through  slots,  in. 
7  =  surface  tension,  dynes/cm. 
PL  =  liquid  density,  Ib.  per  cu.  ft. 
pv  =  vapor  density,  Ib.  per  cu.  ft. 
fe*  =  slot  height,  in. 

7,  «  velocity  based  on  total  slot  area,  f.p.s. 
K  =  see  Table  16-1 

^-^  =  cu.  ft.  per  sec.  per  total  ft.  of  slot  width 
12 

The  variation  in  the  values  of  K  is  believed  to  be  due  mainly  to  the 
effect  of  the  ratio  of  velocity  in  the  cap  to  slot  velocity.  In  the  case  of 
the  small  caps,  this  ratio  was  high,  and  it  is  suggested  that  for  other 
cases  the  following  relation  be  employed. 


The  pressure-drop  values  for  a  number  of  different  caps  are  plotted 
in  Fig.  16-3  on  the  basis  of  the  groups  of  Eq.  (16-3)*  The  data  for,the 
large  caps  are  higher  than  for  the  small  caps,  and  the  slope  of  the  best 


FRACTIONATING  COLUMN  DESIGN 


407 


0.8 


TABIiB    16-1 

Cap  K 

6-in.-diameter  galvanized-iron  cap: 

Slots,  H  by  (0.5,  1.0,  1.5,  and  2.0)  in. 

Slots,  Ke  by,  (0.5,  1.0,  and  1.5)  in. 
6~in.-diameter  cast-iron  cap: 

Slots,  H  by  (0.5  and  1.0)  in 1.0 

4-in  .-diameter  copper  cap: 

Slots,  ^2  by  M  in 0.6 

4-in  .-diameter  cast-iron  cap: 

Slots,  Ke  by  (0.5  and  1.0)  in 0.7 

Tunnel  cap,  slots  KG  by  0.875  in 

4-in.  caps,  tangential  slots  }/%  by  0.75  in. 


0.55 


10 
08 
06 

04 

02 

rf 
*     01 

__ 

fr—^ 

-9  

*        j 

°*+ 

X 

' 

•^  */ 

fc 

,0 

\^ 

<  Q 

+ 

jx 

* 

' 

^ 

<" 

'&  * 

^008 
006 

0.04 
002 
0.0. 

./£ 

Legend 

; 

(14 
& 
(26 
(5, 
'  (5 
(5 
(St 
(5 
(5 

fICG 

.X"^ 

Tunnel  'cap-slots*/*  'w/da  XQBTS'high 

<.        Styles  /*%"*,& 
\        6*  galvanized  iron  cap  *sM$  fox  Q5* 
>       6'galvamzed  iron  cap-slots  '/8"x/'f5'anc/20 
7        6'galvanlzed  iron  cap  -s/ofs  Vie'xOS  ' 
3       6'casfironccp'Stofs%"X05'andJO* 
k       4"  copper  cap  -  slots  %/  X  05  * 
f       4'  casf  iron  cap  •  slots  fa  XQ5  and  /O  * 

\ 

^X^           c 

) 

« 

, 

/        *          \ 

/ 

^ 

> 

a) 
)    """ 

;  ._ 
; 
; 

X 

s 

t 

4 

s 

X 

s 

c 

I 
1 

) 
) 

1              02              04       0.6    OB  1.0               2                468    10               ZQ              40      6C 

FIG.  16-3.     Pressure  drop  through  bubble-cap  slots. 

line  through  the  data  for  any  one  cap  is  approximately  %  for  most 
cases.    The  line  drawn  for  all  the  caps  is 

h*  =  0.12  3-  +  0.058(/i*F,)0-7fi  (16-5) 

and,  including  the  square  root  of  the  density  terms,  Eq.  (16-5)  becomes 

(/ \0.75 
h*V,  J— e^-J  (16-6) 

. 


PL  —  Pv"/ 


408  FRACTIONAL  DISTILLATION 

It  is  believed  that  Eq,  (16-3)  with  a  value  of  K  by  Eq.  (16-4)  gives 
better  results  for  specific  cases,  but  that  Eq.  (16-6)  is  helpful  for  general 
cases. 

With  the  usual  plate  design  the  slots  become  completely  open  for 
vapor  flow  at  values  of  hs  less  than  A*  due  to  the  fact  that  the  average 
density  of  the  vapor-liquid  mixture  around  the  cap  is  only  ^  to  J^ 
that  of  the  liquid.  For  studies  of  single  caps  this  lowered-density  effect 
is  of  less  importance.  At  vapor  rates  higher  than  those  corresponding 
to  complete  opening  of  the  slots,  the  relation  between  hs  and  V8  should 
be  modified.  In  some  cases  the  bottoms  of  the  caps  are  raised  above 
the  plate  to  allow  excess  vapor  to  escape.  With  this  arrangement, 
high  capacities  can  be  obtained  without  excessive  pressure  drops, 
although  the  vapor-liquid  contact  for  such  operation  is  probably  of  low 
effectiveness.  In  other  cases,  the  bottoms  of  the  caps  are  sealed  to  the 
plates,  and  any  excess  vapor  is  forced  through  the  slots.  For  this  con- 
dition the  following  equation  is  suggested  for  h8  greater  than  h* : 

(16-7) 

where  7?  is  vapor  velocity  in  Eq.  (16-3)  for  h8  equal  to  h*  and  hf  is 
the  value  of  h8  for  slots  completely  open.  For  conventional  plate 
arrangements  it  is  suggested  that  hf  be  taken  equal  to  ^h*. 

Pressure  Drop  Due  to  Liquid  Head  above  Slots.  The  pressure  drop 
due  to  the  liquid  head  above  the  top  of  the  slot  is  customarily  taken  as 
equal  to  the  actual  liquid  depth  above  the  slots.  This  liquid  depth  is 
frequently  taken  as  weir  height,  hw,  plus  the  weir  crest,  hcr,  minus  the 
height  of  the  top  of  the  slot,  hap,  thus 

AL  =  hw  +  her  —  hap  (16-8) 

In  some  cases  a  correction  is  added  for  the  additional  liquid  head  at 
the  cap  in  question  owing  to  the  liquid  gradient.  This  method  of 
evaluating  hL  is  simple,  but  the  data  of  Seuren  (Ref.  29),  Ghormley 
(Ref.  11),  and  Kesler  (Ref.  19)  show  that  the  actual  liquid  head  in  the 
aerated  section  of  a  plate  can  be  considerably  less  than  that  at  the  weir. 
This  effect  is  the  result  of  liquid  flow  from  a  nonaerated  section  to  an 
aerated  section  and  back  to  a  nonaerated  section.  This  condition  will 
be  discussed  in  the  section  on  Hydraulic  Gradient.  The  use  of  Eq. 
(16-8)  with  a  correction  for  the  hydraulic  gradient  will  give  high  values 
for  the  liquid  head. 

The  over-all  pressure  drop  for  a  plate,  hp,  is  calculated  by 

hp  —  hc  +  ht  +  hi<  (16-9) 


FRACTIONATING  COLUMN  DESIGN 


409 


Liquid  Flow.  Weirs  and  Down  Pipes.  The  liquid  depth  on  a 
plate  is  controlled  by  exit  overflow  weirs.  The  action  of  the  liquid 
flowing  over  the  weir  is  complicated  by  the  action  of  the  caps  and  by 
the  restrictions  due  to  the  wall.  The  last  row  of  caps  may  blow  a  con- 


A  Cross  flow 
Single  downpfpes 


B  Cross  flow 
Multiple  downpipes 


C   Radial  flow 


0  Cross  flow 
Chord  weirs 


This  space 
blanked 


Partition 
Success 7i 


^  ^  plates  rotated 
•*   by  this  amount 


Circumferential  flow 


FIG.  16-4. 


F  Split  flow 
Chord  weirs 
Downflow  pipe  arrangements  on  bubble-cap  plates. 


siderable  quantity  of  liquid  over  the  weir  as  spray  and  in  surges.  The 
walls  may  be  so  close  to  the  downstream  side  of  the  weir  that  they 
interfere  with  the  liquid  flow. 

A  variety  of  different  weir  and  down-pipe  arrangements  are  employed 
(see  Fig.  16-4). 

In  small  columns,  the  overflow  from  plate  to  plate  is  usually  carried 
in  pipes,  the  upper  end  of  the  pipe  projecting  above  the  plate  surface  to 
form  an  overflow  weir  and  maintain  a  liquid  seal  on  the  plate.  The 
lower  end  extends  into  a  well  on  the  plate  below,  thereby  sealing  the 


410  FRACTIONAL  DISTILLATION 

pipe  so  that  the  vapor  may  not  pass  upward  through  it.  In  larger 
columns,  straight  overflow  weirs  placed  on  a  chord  across  the  tower 
are  often  used. 

Locke  (Ref  .  22)  from  a  study  of  circular  down  pipes  with  the  tops  act- 
ing as  weirs  concluded  that  at  least  three  types  of  liquid  flow  were 
possible  in  circular  down  pipes  with  liquid  seals  at  the  bottom.  At 
low  rates  of  liquid  flow,  the  top  of  the  pipe  acted  as  a  weir,  and  the 
liquid  flowed  down  in  a  film.  As  the  liquid  head  was  increased,  the 
pipe  became  full  and  sucked  vapor  bubbles  down  with  it;  at  still  higher 
liquid  rates,  the  pipe  ran  full  but  did  not  entrap  vapor.  The  first  type 
flow  occurred  for  liquid  head  less  than  one-sixth  to  one-fifth  of  the  pipe 
diameter,  and  this  type  of  flow  could  be  represented  by  the  familiar 
Francis  weir  formula: 

her  -  fc  m  (16-10) 

where  Q  =  cu.  ft.  of  liquid  per  sec. 

L  =  perimeter  of  inside  surface  of  pipe,  ft. 
her  =  head  of  liquid  above  top  of  pipe,  in. 
k  ~  constant,  increasing  from  3.9  to  4.3  as  pipe  size  was 

increased  from  0.87  to  2.07  in. 

Rowley  (Ref.  27)  recommends  the  following  equation  for  the  last  two 
types  of  flow: 


where  hf  =  head  of  liquid  necessary  to  overcome  down-pipe  friction 

and  entrance  and  exit  losses,  in. 
gc  =  proportionality  constant,  32.2 
D  =  inside  diameter  of  pipe,  ft. 
V0  *  linear  velocity  of  liquid  in  pipe,  f  .p.s. 
/  =  proportionality  constant  of  Fanning  friction  equation  (See 
Walker,  Lewis,  Me  Adams,  and  Gilliland,  "Principles  of 
Chemical  Engineering,"  3d  ed.,  p.  78.) 

For  large  straight  weirs  when  the  downspouts  are  not  running  full, 
the  Francis  weir  equation  may  be  used.  This  weir  formula  is  fre- 
quently employed  neglecting  the  approach  velocity,  and  in  some  cases 
this  may  not  be  a  satisfactory  approximation.  The  equation  for  zero 
approach  velocity  is 


Units  same  as  for  Eq.  (16-10). 


FRACTIONATING  COLUMN  DESIGN 


411 


To  allow  for  the  approach  velocity,  a  correction  factor  is  given  in 
Fig.  16-5,  as  a  function  of  Q/L  &nd  hor  +  hw,  where  hw  is  the  height  of 
the  weir.  This  correction  was  calculated  on  the  basis  that  the  liquid 
approaching  the  weir  had  a  velocity  corresponding  to  the  unaerated 
depth  of  hcr  +  hw.  The  actual  velocity  of  approach  is  probably  higher 
owing  to  the  caps  and  vapor.  The  correction  factor  is  multiplied  into 
the  right-hand  side  of  Eq.  (16-12)  to  calculate  hcr. 

When  the  head  on  the  weir  becomes  high,  the  liquid  carries  past  the 
dam  a  considerable  distance,  and  the  wall  may  interfere  with  the  flow. 
In  circular  pipes,  Locke  reported  that  the  streams  from  the  various 
sides  interfered  with  each  other  when  the  head  on  the  weir  at  the  top 
of  the  pipe  was  0.2  of  the  pipe  diameter.  This  would  indicate  an  over- 


FIG.  16-5.     Correction  factor  for  Eq.  (16-12). 

shoot  of  over  twice  the  head.  The  effect  for  straight  weirs  is  probably 
less,  and  Edminster  (Ref .  7)  has  suggested  neglecting  the  portion  of  the 
weir  that  is  closer  to  the  wall  than  the  value  of  hcr-  This  would  appear 
to  be  a  reasonable  assumption. 

For  large  downspouts  running  full,  it  is  recommended  that  Eq. 
(16-11)  be  used,  employing  for  D  four  times  the  hydraulic  radius,  which 
is  equal  to  the  cross-sectional  area  of  the  downspout  divided  by  the 
perimeter. 

When  the  liquid  head  over  the  weir  is  low,  the  discharge  from  one 
side  to  the  other  will  vary  if  the  top  edge  is  not  level.  This  variation 
can  be  reduced  by  using  a  V-notched  top  which  reduces  the  effective 
weir  length  at  low  rates  of  flow.  In  general,  it  is  desirable  to  keep  the 
head  over  the  weir  low  because  this  reduces  the  variation  of  liquid 
depth  on  the  plate  for  different  operating  rates.  Values  of  her  of  1.0 
in.  are  common,  but  they  seldom  exceed  3  in. 

In  some  cases,  under-  and  overweirs  are  used  at  the  outlet  to  handle 


412  FRACTIONAL  DISTILLATION 

two  liquid  layers  on  a  plate.  Such  an  arrangement  can  also  be  used  to 
hold  back  foam  from  the  down  pipe.  The  clearance  between  the 
underweir  and  the  plate  must  be  adequate  to  handle  the  liquid  load. 

The  bottoms  of  downflow  pipes  must  have  enough  clearance  to  allow 
the  liquid  to  flow  easily,  but  they  should  not  allow  vapor  to  by-pass  up 
through  them.  This  vapor  by-passing  is  usually  prevented  by  a  posi- 
tive seal  which  holds  liquid  above  the  bottom  of  the  down  pipe  (see 
Fig.  16-1).  In  large  columns,  this  seal  is  often  made  an  inlet  weir 
which  serves  to  distribute  the  liquid  as  well  as  seal  the  down  pipe. 

For  the  reversal  loss  at  the  bottom  of  the  down  pipe,  it  is  recom- 
mended that  a  loss  equal  to  one  kinetic  head  be  used  with  a  coefficient 
of  0.6. 


=  0.5F1, 

where  ho  =  loss  in  head  at  bottom  of  down  pipe,  in. 

VD  =  maximum  velocity  at  bottom  of  down  pipe,  f.p.s. 

Liquid  Gradient.  One  of  the  important  factors  that  must  be  con- 
sidered in  the  design  of  a  bubble  plate  is  the  liquid  gradient  across  the 
plate.  It  is  obvious  that  the  liquid  level  will  normally  be  higher  at  the 
liquid  inlet  than  at  the  outlet.  In  small  towers,  this  difference  in  level 
offers  no  serious  difficulties,  but  in  towers  of  moderate  and  large  diam- 
eters it  can  become  so  great  that  the  vapor  distribution  is  poor  and  the 
overflow  may  by-pass  the  plate  by  dumping  through  the  caps. 

The  gradient  is  due  to  the  resistance  to  liquid  flow  across  the  plate 
and  results  from  (1)  friction  with  the  caps  and  the  plate,  (2)  eddy  losses 
in  the  liquid  due  to  repeated  acceleration  and  deceleration,  and  (3) 
resistance  due  to  the  effects  of  vapor  flow. 

Experimental  data  on  this  gradient  have  been  published  by  several 
investigators  (Refs.  10,  11,  12,  13,  18,  19,29).  Gonzales  and  Roberts 
(Ref.  12)  studied  a  plate  with  4%~in.-diameter  caps  which  were  6J^  in. 
tall  using  air  and  water.  Their  column  was  rectangular  in  shape  with 
12  rows  of  caps  in  the  direction  of  liquid  flow.  Based  on  liquid  flow 
around  staggered  pipes,  they  developed  the  following  equation  for  the 
gradient  as  a  function  of  the  number  of  rows  of  caps: 

ft2.8  =  A  -  En  (16-14) 

where  h  ~  liquid  depth  on  plate 

A,B  SB  constants 
Their  data  were  taken  at  liquid  levels  less  than  the  top  of  the  caps 


FRACTIONATING  COLUMN  DESIGN  413 

and  agreed  well  with  the  equation  for  both  aerated  and  unaerated  con- 
ditions. A  and  B  were  functions  of  both  the  water  and  air  rates. 

Bijawat  (Ref.  2)  reviewed  the  data  obtained  by  Gonzales  and 
Roberts  and,  on  the  basis  of  orifice-type  flow  of  the  liquid  between 
the  caps,  suggested 

h*  =  A'  -  B'n  (16-15) 

This  equation  correlated  the  data  about  as  well  as  Eq,  (16-14). 
Seuren  (Ref.  29),  Ghormley  (Ref.  11),  and  Kesler  (Ref.  19)  obtained 
data  on  a  rectangular  plate  with  10  rows  of  4-in.  caps  in  the  direction 
of  liquid  flow.  Good,  Hutchinson,  and  Rousseau  (Ref.  13)  investi- 
gated the  liquid  gradient  on  a  rectangular  plate  having  12  rows  of  3-in. 
caps  for  a  number  of  operating  conditions. 

Klein  (Ref.  20<z)  has  investigated  the  factors  that  cause  the  loss  in 
head  of  the  liquid  flowing  across  the  plate.  His  data  indicate  that 
with  no  vapor  flow  there  is  essentially  no  hydraulic  gradient  even  at 
liquid  rates  considerably  greater  than  those  normally  employed.  It 
was  concluded  that  the  loss  in  head  was  due  to  high  f rictional  losses  for 
the  flow  of  the  vapor-liquid  mixture.  The  drag  per  unit  area  was 
determined  by  measuring  the  force  on  a  plate  suspended  in  the  aerated 
liquid,  and  values  ten  times  as  large  as  for  unaerated  liquid  flowing  at 
the  same  linear  velocity  were  found.  Abnormally  high  f  rictional 
losses  have  also  been  reported  for  the  flow  of  mixtures  of  vapor  and 
liquids  in  pipes. 

Klein  confirmed  the  observation  of  Ghormley  and  Kesler  that  the 
hydrostatic  head  of  the  liquid  in  the  aerated  section  was  frequently 
less  than  that  at  either  the  inlet  or  outlet  to  the  plate.  Klein,  Kesler, 
and  Bloecher  (Ref.  2a)  measured  the  potential  head  of  the  liquid  by 
determining  the  hydrostatic  head  as  a  function  of  depth  and  then 
integrating  up  from  the  bottom  of  the  plate.  The  potential  head  is 
equal  to  the  number  of  inches  above  the  tray  at  which  all  the  water 
present  would  give  the  same  total  potential  energy  as  the  aerated 
liquid.  For  a  nonaerated  liquid  the  potential  head,  according  to  this 
definition,  is  equal  to  one-half  the  liquid  depth.  It  was  shown  that 
the  potential  head  decreases  progressively  across  the  plate.  It  is  this 
head,  and  not  the  hydrostatic  head,  which  is  the  driving  force  for 
liquid  flow.  The  flow  across  a  bubble  plate  can  be  considered  in  terms 
of  five  zones.  (1)  There  is  the  inlet  unaerated  section  in  which  the 
potential  head  is  twice  the  hydrostatic  head.  (2)  There  is  the  transi- 
tion section  from  the  first  zone  to  the  aerated  region.  This  transition 
occurs  a  short  distance  upstream  from  the  first  row  of  caps  to  about 


414  FRACTIONAL  DISTILLATION 

the  second  row  of  caps.  In  this  region  (a)  the  apparent  depth  of  liquid 
rises  sharply  owing  to  the  aeration,  (b)  the  hydrostatic  pressure  drops 
abruptly,  and  (c)  the  potential  head  remains  almost  constant.  For  the 
potential  head  to  remain  constant  the  hydraulic  head  must  decrease 
because  some  of  the  liquid  is  at  a  higher  level.  Near  the  plate  the 
pressure  in  the  liquid  in  Zone  1  is  higher  than  in  the  transition  section 
and  liquid  flows  towards  the  caps,  but  at  higher  levels  the  reverse  is 
true  and  the  liquid  flows  back  (against  the  normal  flow).  (3)  The 
third  zone  is  the  main  aerated  section  in  which  the  hydrostatic  head 
remains  essentially  constant  and  there  is  a  progressive  decrease  in  the 
potential  head.  (4)  The  fourth  zone  is  the  outlet  transition  zone  which 
extends  from  the  last  two  rows  of  caps  to  the  calming  section  before  the 
outlet  weir.  The  phenomena  are  similar  to  those  for  the  inlet  transi- 
tion zone,  i.e.,  fa)  the  apparent  depth  of  the  liquid  drops  sharply,  (b) 
the  hydrostatic  head  increases  abruptly,  and  (c)  the  potential  head 
remains  almost  constant.  In  this  case  there  is  a  flow  of  liquid  back 
toward  the  caps  near  the  plate  and  toward  the  outlet  section  at  the 
higher  levels.  (5)  The  final  region  is  the  unaerated  calming  section 
before  the  outlet  weir.  These  effects  are  illustrated  in  Fig.  16-6.  At 
low  air  rates  there  is  a  decrease  in  hydrostatic  head  in  the  aerated 
zone,  but  a  clear  liquid  layer  on  the  plate  extends  from  the  inlet  to  the 
outlet.  At  higher  vapor  rates  the  liquid  becomes  "completely  aer- 
ated," and  no  apparent  clear  liquid  layer  remains.  This  condition  is 
shown  in  diagram  A  of  Fig.  16-6.  Data  on  the  hydrostatic  and  poten- 
tial head  for  a  completely  aerated  condition  are  shown  in  diagram  B. 
The  sharp  drop  and  rise  in  the  hydrostatic  head  at  the  two  ends  of  the 
plate  are  clearly  shown,  but  there  is  essentially  no  change  in  the  aerated 
section.  The  value  of  twice  the  potential  head  is  plotted  so  that  it 
will  be  numerically  equal  to  the  hydrostatic  head  in  the  two  nonaerated 
sections.  The  potential  head  decreases  regularly  across  the  plate. 

Referring  again  to  diagram  5,  Fig.  16-6,  the  hydrostatic  head  given 
is  the  value  at  the  bottom  of  the  plate.  If  values  are  determined  in 
planes  parallel  to,  but  above,  the  plate,  it  is  found  that  the  hydrostatic 
head  decreases  more  rapidly  with  height  in  the  nonaerated  sections 
than  in  the  bubbling  zone  owing  to  the  lower  density  of  the  vapor- 
liquid  mixture.  Eventually,  at  a  height  less  than  h<»  the  hydrostatic 
head  in  the  outlet  transition  zone  (Zone  4)  becomes  equal  to  that  in 
Zone  5,  and  at  higher  heights  above  the  plate  the  aerated  section  has 
the  higher  head.  Thus,  the  hydrostatic  head  is  a  complicated  func- 
tion of  the  position  and  the  distance  above  the  plate. 

Klein  found  that  for  a  given  liquid  rate  the  hydraulic  gradient 


FRACTIONATING  COLUMN  DESIGN 


415 


increased  with  increasing  vapor  rate  until  the  liquid-vapor  mixture  in 
the  aerated  section  had  an  apparent  density  approximately  one-third 
that  of  the  liquid  but  further  increases  in  vapor  rate  did  not  appreciably 
change  the  loss  in  head.  He  termed  the  condition  for  the  mixture 
density  equal  to  one-third  the  liquid  density  "complete  aeration." 


•§.§ 
H 


3 

I 


Liquid  vapor  mixture 

rn  n  n  n  n 


Zone  3 


Diagram  A 


Inches  of  wafer 

0  —  N>  01  ^  4s 

T" 

-+.-. 

*—  —  -. 

-fJ: 

>**. 

oote* 
~-+^ 

rf/at 

hea 

i 

A 

Hyd 

**o$h 

vtic  t 

heaa 

-~<*p«, 
f 

-~~  -, 

+ 

"•**•-*. 

7f 

5     ' 

f 

k  . 

,  P 

s 

Iming  sect 

Exi 
Wai- 
fool 
Supe 

t~  we 
er  re 

-Off 

fffc/, 

ir  *L 
ite  * 
>/a/<? 
&/  ai 

>" 

'2S$ 

widi 
>  vei 

ipm 
to 
octfy 

per 
>*2.8 

1  Inlet  cot 

ft./s 

ec. 

Outlet  ca 

23456789      10 
Number  of  rows  of  caps  in  direcfion  of  liquid  flow 

Diagram  B 
FIG.  16-6. 

Higher  air  rates  or  foaming  agents  increased  the  apparent  liquid  height, 
but  static  pressure  probes  indicated  that  for  these  cases  the  density 
was  approximately  one-third  that  of  the  liquid  for  the  main  region  of 
fluid  flow  just  above  the  plate  and  on  top  of  this  layer  there  was  a 
light  froth  containing  essentially  no  liquid.  Practically  all  the  liquid 
flow  was  accounted  for  in  the  layer  with  the  one-third  normal  density. 
The  light  froth  did  not  appear  to  have  a  significant  effect  on  the 
hydraulic  gradient. 


416  FRACTIONAL  DISTILLATION 

Klein  correlated  his  data  and  that  of  other  investigators  on  the  basis 
of  a  Fanning-type  friction  equation. 


F  =  (16-16) 

QcTh 

n  =  ^  (16-17) 

where  F  =  loss  in  head  from  inlet  to  exit  calming  sections,  ft. 
/'  =  friction  factor 
Vf  =  velocity  of  liquid  in  foam,  f  .p.s. 

-  Qw/(pfL0b) 

Qw  =  liquid  rate,  Ibs.  per  sec. 
pf  =  density  of  foam,  Ibs.  per  cu.  ft. 
Lo  =  foam  height,  ft. 
6  =  width  of  plate,  ft. 
B  =  length  of  bubbling  section,  ft. 
gc  =  conversion  constant  =  32.2 

Th  =  mean  hydraulic  radius,  ft.  =  ,    ,   °j 

The  values  of  /'  were  calculated  from  the  experimental  data  and  cor- 
related as  a  function  of  a  modified  Reynolds  number,  Ref  =  (r*7/p/)//*/, 
where  M/  =  viscosity  of  foam.  In  making  the  correlation  it  was 
assumed  (1)  that  the  average  foam  density  was  one-third  the  density 
of  the  normal  liquid,  (2)  that  the  viscosity  of  the  foam,  /*/,  was  one- 
third  the  true  viscosity  of  the  liquid,  and  (3)  that  L0  was  equal  to  two 
times  the  hydrostatic  head  in  the  outlet  calming  section.  Thus, 

#e>  =!^Z/^  (16-18) 

ML 
Vf  =  ^r  (16-19) 

PLLOO 
and 

Lo  =  2h0  (16-20) 

where  PL  =  normal  density  of  liquid,  Ibs.  per  cu.  ft. 
ML  =  normal  viscosity  of  liquid,  Ibs.  per  f  .p.s. 
h0  —  hydrostatic  head  in  outlet  calming  section,  ft. 
Klein  found  that  the  value  of/'  was  a  function  of  the  exit-weir  height 
relative  to  height  of  the  slot  and  used  an  empirical  relation  to  allow 

/     h       \« 

for  this  effect.     He  correlated  /'  (  ?  —  *2_  )     as  a  function  of  Re1, 

\h0  —  hsp/ 

where  h8p  «  height  of  top  of  slot  above  plate. 


FRACTIONATING  COLUMN  DESIGN 


417 


The  hydraulic  gradient  results  of  several  investigators  are  plotted 
in  Fig.  16-7.  In  view  of  the  number  of  factors  and  assumptions 
involved,  the  correlation  is  reasonably  good.  Most  of  the  results 


10 


4.0 


1.0 


a-  04 


0.1 


0.01 


\* 

4. 

s 

+ 

Q 

c 

\. 

4 

•P4! 

p 

4 

M 

n 

X 

Q 

n 

n 

*Q 

-*• 

T 

S 

i 

D 

c 

+\ 
i< 

£k 

c 

k 

^ 

> 

—  JXg- 

>A 

Klein  -  T'weir 
Klefn-3"\Mefr 
Klein  -5"  weir 
Kemp  and  Py/e 
Gooa,  Hutch/risi 
Ghorm/ey 
Kemp  and  Py/e 

\ 

S. 

^ 

0 
D 

X 

s 

(a/r-  water) 
>/7  Rousseau 

(air  -  Perch/oroefhy/en 

e) 

1,000  4,000  10,000  40,000 

Re-  Modified  Reynolds  number 
FIG.  16-7. 

correlated  were  for  essentially  standard  caps  and  plate  arrangements, 
and  the  relation  should  be  used  with  caution  for  plate  layouts  that 
differ  significantly. 

It  will  be  noted  for  Fig.  16-7  that  /'  is  essentially  inversely  propor- 
tional to  the  value  of  Re'.    Combining  this  relationship  with  the  fact 


418  FRACTIONAL  DISTILLATION 

that  for  large  plates  2L0  is  small  in  comparison  with  6  gives 


This  relation  would  indicate  that  the  hydraulic  gradient  is  directly 
proportional  to  the  liquid  flow  rate  per  foot  of  plate  width  and  to  the 
length  of  the  bubbling  section.  The  outlet  hydrostatic  head  is  prob- 
ably the  most  important  factor  determining  the  hydraulic  gradient, 
and  high  outlet  weirs  should  be  an  effective  method  of  reducing  the 
loss  in  head. 

Equation  (16-21)  indicates  that  the  gradient  is  approximately  pro- 
portional to  the  liquid  viscosity.  The  liquid  viscosities  studied  ranged 
from  that  of  water  at  room  temperature  to  glycerine.  Gardner  (Ref  . 
10)  studied  the  hydraulic  gradient  on  a  plate  with  tunnel  caps  using 
water  at  different  temperatures  and  concluded  that  the  liquid  viscosity 
was  not  a  factor.  The  plate  design  was  unusual  in  that  the  liquid  flow 
was  across  rather  than  along  the  tunnel  caps.  However,  until  addi- 
tional data  are  available  it  is  recommended  that  PL  be  taken  equal  to 
0.00067  pounds  per  f  .p.s.  for  all  liquids  of  lower  viscosity  and  equal  to 
the  actual  viscosity  for  those  having  higher  values. 

Liquid  Head.  As  shown  in  Fig.  16-6  the  hydrostatic  head  in  the 
bubbling  section  can  be  less  than  at  the  outlet  weir.  However,  the 
liquid  head  above  the  top  of  the  slots  may  be  greater  or  less  than  the 
difference  in  the  height  of  the  outlet  liquid  and  the  top  of  the  slots 
owing  to  the  increased  depth  of  the  vapor-liquid  mixture.  The  value 
of  the  liquid  head  calculated  by  Eq.  (16-8)  should  be  satisfactory  for 
most  cases,  but  the  actual  value  may  differ  somewhat  due  to  the  aera- 
tion effects. 

Plate  Stability.  The  term  "  plate  stability"  has  been  used  to 
describe  the  vapor  and  liquid  distribution  on  a  plate.  ,  A  stable  plate 
has  been  defined  as  one  in  which  all  the  caps  are  handling  vapor, 
although  the  quantity  of  vapor  per  cap  may  vary  widely  from  one  side 
of  the  plate  to  the  other.  The  plate  is  stable  in  the  sense  that  liquid  is 
flowing  across  the  plate  and  not  by-passing  by  "dumping"  through 
the  cap  risers,  but  the  distribution  of  vapor  may  be  quite  poor. 

The  term  "stable  plate"  does  not  differentiate  between  the  various 
types  of  plate  action  and  in  this  text  the  following  terms  will  be  used. 

Uniform  vapor  distribution  —  will  indicate  the  condition  when  each 
cap  on  the  plate  handles  the  same  amount  of  vapor  per  unit  time. 
Active  cap  —  will  indicate  that  vapor  is  passing  through  the  cap. 


FRACTIONATING  COLUMN  DESIGN  419 

Inactive  cap  —  will  indicate  that  vapor  is  not  passing  through  the  cap. 
Completely  active  plate  —  will  indicate  that  all  caps  are  active. 
Partly  active  plate  —  will  indicate  that  only  part  of  the  caps  are  active. 
Plate  dumping  —  will  indicate  that  liquid  is  flowing  to  the  plate  below 
through  some  of  the  cap  risers. 

The  distribution  of  the  vapor  among  the  various  caps  is  a  function  of 
the  pressure  drops  involved.  The  lateral  pressure  difference  in  the 
vapor  space  above  a  plate  is  usually  small  and,  when  this  condition  is 
true, 

hp  =  hc  +  h8  +  hL  =  constant 

Because  hL  varies  across  the  plate,  hc  +  h8  must  vary.  Substituting 
the  values  from  Eqs.  (16-2)  and  (16-3)  in  the  above  equation  gives 


(16-22) 

PL/  PL  \  \PL-~PvJ 

and,  for  a  given  plate,  this  can  be  condensed  to 

fciF2  +  kzV*  +  hL  =  hp  -  0.12  -£•  (16-23) 

The  right-hand  side  of  Eq.  (16-23)  would  be  constant  for  all  parts  of 
the  plate  on  the  basis  of  the  assumptions  made,  and  if  the  value  of  hL 
varies  across  the  plate,  then  the  vapor  flow  per  cap  must  vary.  This 
relation  indicates  that  a  given  cap  will  become  inactive  when  ht  at 
that  point  is  equal  to  hp  —  Q.l2(y/pL)  ;  i.e.,  when  the  liquid  head  over  a 
slot  becomes  equal  to  the  total  pressure  drop  minus  the  pressure  drop 
necessary  to  initiate  bubbling  against  the  surface  tension. 

Consider  the  case  for  the  caps  at  the  inlet  side  just  inactive.  By 
Eq.  (16-23), 


PL 

where  hi  is  the  head  of  liquid  above  the  slots  at  the  inlet  row  of  caps 
where  the  pressure  drop  is  just  sufficient  to  cause  bubbling. 

Assuming  that  the  plate  is  rectangular  and  that  the  two  terms  involv- 
ing velocity  can  be  combined, 

fc,72  +  ktV*  -  k*Vm    t  (16-25) 

where  V  ~  vapor  rate  per  row  of  caps  perpendicular  to  direction  of 

liquid  flow 
m  »  constant  between  %  and  2.0 


420  FRACTIONAL  DISTILLATION 

Then,  for  any  row  of  caps, 

k*Vm  «  hp  -  hL  -  0.12  -1  (16-26) 

PL 

Equation  (16-26)  could  be  used  to  evaluate  the  vapor  distribution  if 
the  variation  in  hi,  across  the  plate  were  known.  It  has  already  been 
shown  that  the  variation  of  the  hydrostatic  head  across  the  plate  where 
the  liquid  is  aerated  is  a  function  of  the  height  above  the  plate  at  which 
the  head  is  determined,  and  the  relation  for  the  plane  corresponding 
to  the  top  of  the  slots  could  vary  significantly  depending  on  the  height 
of  the  slots  relative  to  the  hydrostatic  heads  in  the  inlet  and  outlet 
calming  sections.  However,  as  a  first  approximation  it  will  be  assumed 
that 


where  h£,  h°L  =  liquid  head  above  top  of  slots  at  inlet  and  outlet  of 

plate 

N  =  rows  of  caps  on  plate 
n  =  number  of  rows  of  caps  from  outlet  weir 
When  the  inlet  row  of  caps  is  just  inactive,  the  condition  for  the  out- 
let row  of  caps  can  be  obtained  by  combining  Eqs.  (16-24)  and  (16-26), 
giving 

k*V?  =  hi  -  h°L  (16-28) 

where  V0  is  the  vapor  rate  at  the  outlet  row  of  caps.  Thus  the  pres- 
sure drop  due  to  vapor  flow  for  the  outlet  row  of  caps  is  equal  to  the 
hydraulic  gradient  when  the  inlet  row  of  caps  is  just  inactive.  Equa- 
tions (16-27)  and  (16-28)  can  be  combined  to  evaluate  the  pressure 
drop  and  the  vapor  distribution.  In  order  to  simplify  the  analysis,  it 
is  assumed  that  the  number  of  rows  of  caps  is  large  enough  that  con- 
ditions vary  approximately  continuously  across  the  plate  so  that  inte- 
gration instead  of  stepwise  summation  can  be  employed.  Thus,  for 
the  total  vapor  flow, 


dn 
KS     /      JN\        &/ 

N 


A*   -  h°\l/m   (°  t         n\lf 
V  dn  -   -  I     L          L  1         I     I  1   —  —  1 

Yd  \     k*     )      JN\l      N) 

(hj  -  h 

\        &3 


where  VT  equals  total  vapor  through  plate. 


FRACTIONATING  COLUMN  DESIGN  421 

Combining  this  relation  with  Eq.  (16-28)  gives 


If  there  were  uniform  vapor  distribution,  V0  would  equal  (Vr/N), 
and  the  above  relation  indicates  that  the  pressure  drop  is  increased  by 
the  liquid  gradient.  Where  the  inlet  row  of  caps  is  just  inactive,  the 
pressure  drop  due  to  velocity  is  greater  by  the  factor  [(m  +  l)/m]w, 
and  this  term  varies  only  from  1 .84  to  2.25  for  m  from  %  to  2.  A  value 
of  2.0  is  within  the  accuracy  of  the  assumption  made. 

The  plate  pressure  drop  when  the  inlet  caps  just  become  inactive, 
hp,  is 

hy  =  k,V?  +  hi  +  0.12  -^  (16-31) 

PL 

while,  for  a  plate  with  uniform  vapor  distribution,  it  would  be 

+  &S, +  0.12-2-  (16-32) 

PL 

and  using  the  factor  of  2.0  obtained  for  Eq.  (16-30)  gives 

hy  =  2/c3  (%H    +hl  +  0.12  -2. 
\M  /  PL 


In  cases  where  the  value  of  the  bracketed  term  is  small,  the  pressure 
drop  for  a  plate  with  caps  just  becoming  inactive  is  equal  to  approxi- 
mately twice  the  pressure  drop  for  the  same  plate  with  the  same  total 
vapor  load  uniformly  distributed.  If  the  last  terms  are  not  small,  the 
pressure  drop  will  be  increased  by  a  factor  less  than  2.0.  This  increased 
pressure  drop  is  one  of  the  objections  to  hydraulic  gradient. 

The  relations  given  in  Eqs.  (16-24)  to  (16-32)  were  based  on  the 
condition  that  the  inlet  row  of  caps  was  just  inactive.  Similar  analysis 
can  be  made  for  other  conditions.  Thus,  all  caps  on  a  plate  will  be 
active  if 


7 

Il~L   "t~   V.L& 
7 


hp  >  (hi  -  hi)  +  hi  +  0.12  -L  (16-33) 

PL 


For  an  active  plate,  with  k*V™>  hL  —  h°L,  it  is  frequently  desirable 


422  FRACTIONAL  DISTILLATION 

to  estimate  the  actual  pressure  drop  in  terms  of  that  for  a  plate  with 
uniform  vapor  distribution.  By  the  same  type  of  integration  employed 
for  Eq.  (16-29),  an  approximation  can  be  obtained  for  V0  and  the  value 
of  hp.  Thus, 

Va  . EOV^L (16_34) 


2  (*;-** -0.12  £) 

and,  by  Eq.  (16-23), 

!  +  W  +  h'L  +  0.12  £  (16-35) 


The  term  (h'9  -  h°L  -  0.12  %\  is  equal  to  ki(VT/N)*  +  kz 

and  will  be  termed  the  velocity  head  for  uniform  vapor  distribution, 
hy.  It  is  equal  to  the  pressure  drop  due  to  the  vapor  flow  in  the  riser, 
cap,  and  slots  for  a  plate  operating  with  a  uniform  velocity  distribution. 
For  a  plate  with  the  inlet  row  of  caps  just  inactive,  2hv  is  equal  to 
(Aj,  —  h°L)  and  Eq.  (16-34)  reduces  to  the  previous  criterion.  When 

(/4,  —  h°L)  is  small  in  comparison  to  2hv,  the  equation  gives  V0  =  -r£; 

i.e.,  the  vapor  distribution  is  uniform.  To  use  these  equations,  hfp  is 
calculated  using  V  equal  to  (Vr/N)  and  TIL  —  /&!,  and  F0  is  obtained 
from  Eq.  (16-34).  The  pressure  drop  for  the  plate  is  then  calculated 
by  Eq.  (16-35). 

The  distribution  of  the  vapor  among  the  caps  is  important,  and  an 
approximation  can  be  obtained  by  the  following  relation: 

Vn  _  i       (hj  -  kj)(n/N) 

To  ~  x  Wv 

where  F«  =  vapor  rate  through  inflow  cap. 

In  order  to  obtain  a  velocity  through  the  first  row  of  caps  equal  to 
one-half  that  through  the  last  row  would  require  that  2hv  be  twice  the 
hydraulic  gradient,  and  this  would  appear  to  be  about  the  minimum 
ratio  of  velocities  that  should  be  considered  for  design  purposes. 

Plate  Dumping.  This  condition  can  occur  in  extreme  cases.  When 
the  upstream  row  of  caps  becomes  inactive  owing  to  an  increase  in  the 
hydraulic  gradient,  the  liquid  level  in  the  caps  will  be  depressed  below 
the  top  of  the  slots  by  an  amount  equal  to  the  surface  tension  effect. 
If  the  gradient  is  increased  further,  the  liquid  level  under  the  inactive 
caps  rises  and  the  pressure  drop  for  the  same  rate  of  vapor  flow 


FRACTIONATING  COLUMN  DESIGN  423 

increases.  Eventually  a  condition  is  reached  where  the  liquid  level 
under  some  of  the  inactive  caps  reaches  the  top  of  the  riser  and  liquid 
spills  to  the  plate  below.  In  extreme  cases,  essentially  all  of  the  liquid 
flowing  to  a  plate  will  dump  down  through  ,the  risers  of  the  first  few 
rows  of  caps,  and  very  little  will  flow  across  the  plate.  Under  these 
conditions,  there  is  frequently  an  abrupt  drop  in  liquid  level  on  the 
plate  between  the  section  of  inactive  and  active  caps.  The  momentum 
of  the  vapor  issuing  from  the  active  caps  acts  as  dam  for  the  liquid 
giving  a  "Red  Sea "  effect.  Plate  dumping  is  undesirable  because  with 
the  usual  column,  with  liquid  flowing  in  opposite  directions  on  succes- 
sive plates,  the  liquid  that  spills  through  the  risers  by-passes  the  vapor 
on  two  plates. 

Equations  for  the  condition  of  plate  dumping,  similar  to  those  for 
inactive  caps,  can  be  derived  but,  because  it  is  an  undesirable  condition, 
it  is  better  to  control  conditions  such  that  all  the  caps  are  active.  This 
will  ensure  against  plate  dumping.  Plate  dumping  conditions  in  a 
large  commercial  tower  have  been  described  by  Harrington  et  al. 
(Ref.  15). 

General  Design  Considerations.  It  is  usually  desirable  to  have  a 
bubble  plate  that  gives  (1)  a  low  pressure  drop  and  (2)  a  reasonably 
uniform  vapor  distribution.  These  two  conditions  are  partly  incom- 
patible because  a  high  pressure  drop  usually  gives  more  uniform  vapor 
distribution. 

The  vapor  distribution  relation  of  Eq.  (16-36)  indicates  that  the  two 
main  factors  involved  in  obtaining  a  low  variation  in  (V/V0)  across  the 
plate  are  (1)  low  value  of  the  hydraulic  gradient  and  (2)  a  high  value  of 
hv.  Equation  (16-16)  shows  that  a  low  liquid  gradient  will  be  obtained 
for  a  given  liquid  flow  rate  by  increasing  h0,  or  decreasing  N. 

Increasing  the  liquid  depth  should  be  very  effective  in  lowering  the 
gradient  due  to  the  increase  in  rh  and  the  decrease  in  7/.  A  deep  liquid 
level  above  the  top  of  the  caps  is  not  desirable  because  it  allows  surges 
and  wave  action  to  occur.  Raising  the  caps  to  allow  liquid  flow  below 
the  skirts  is  probably  one  of  the  most  effective  ways  of  reducing  the 
hydraulic  gradient.  It  does  introduce  the  possibility  that  part  of  the 
liquid  will  cross  the  plate  without  intimate  contact  with  the  vapor. 
This  condition  would  be  most  serious  when  the  column  is  operating  at 
reduced  capacity.  This  method  has  the  advantage  that  it  reduces  the 
gradient  without  increasing  the  pressure  drop. 

The  space  above  the  caps  should  be  as  unobstructed  as  possible  to 
allow  free  passage  of  the  liquid.  Hold-down  bars  for  the  caps  or  other 
mechanical  devices  should  be  arranged  in  the  direction  of  liquid  flow. 


424  FRACTIONAL  DISTILLATION 

Decreasing  the  value  of  N  is  a  common  method  of  improving  vapor 
distribution,  and  this  result  is  obtained  by  the  use  of  split  plates  or 
multiple  downspouts.  This  shortened  path  may  be  accomplished  by 
having  the  inlet  downspout  on  one  plate  in  the  center  and  the  exit 
downspouts  placed  uniformly  around  the  circumference  (Fig.  16-4(7). 
Thus,  the  liquid  flows  out  radially  and  crosses  only  half  of  the  plate. 
On  the  next  plate  below,  the  liquid  would  flow  radially  into  the  center. 
Other  designs  bring  the  liquid  in  at  opposite  sides  and  flow  across  half 
the  tower  to  central  weirs  that  extend  across  the  tower  perpendicular 
to  the  direction  of  flow  (Fig.  16-4F).  On  the  plate  below,  the  liquid 
flow  is  outward.  This  latter  method  can  be  arranged  to  give  any  frac- 
tional distance  across  to  plate  desired,  such  as  ^,  J^,  ^,  or  ^. 

These  arrangements  tend  to  decrease  the  hydraulic  gradient,  but 
they  may  lower  the  plate  efficiency  by  reducing  the  cross  flow  effect, 
and  they  offer  difficulties  in  properly  proportioning  the  overflow 

between  the  different  sections. 

Another  arrangement  that  at- 
tempts to  obtain  complete  cross 
flow,  but  reduced  hydraulic  gradi- 
ent, is  the  staggered  plate  illus- 
trated in  Fig.  16-8.  In  this  case; 
the  plate  is  broken  up  into  narrow 
segments,  each  with  its  own  over- 
flow weir.  If  the  segments  are  nar- 
row and  the  weirs  are  all  adjusted 

FIG,  ie-8.    Cascade  plate.         to  tbe  proper  height,  the  level  at 

the  caps  can  be  kept  uniform.  The 
plate  also  has  the  advantage  of  complete  cross  flow.  The  main  diffi- 
culty is  the  constructional  complexity. 

Increasing  the  value  of  hv  by  increasing  the  pressure  drop  will  be 
effective  in  improving  vapor  distribution.  The  use  of  excessive  pres- 
sure drops  to  obtain  this  result  is  not  desirable,  although  it  may  be  the 
simplest  method  of  correcting  a  column  that  has  already  been  built. 
In  this  latter  case,  some  of  the  caps  can  be  removed,  or  constrictions  can 
be  placed  in  the  risers  to  increase  the  pressure  drop.  This  increase  in 
pressure  drop  increases  the  possibility  of  flooding  the  column  by  liquid 
backing  up  the  down  pipes.  Good  original  design  should  give  low 
pressure  drop  and  good  vapor  distribution. 

Entrainment.  Entrainment  is  the  carrying  of  the  liquid  from  one 
plate  to  the  plate  above  by  the  flow  of  the  vapor.  It  is  usually  defined 


FRACTIONATING  COLUMN  DESIGN  425 

as  the  weight  of  liquid  entrained  per  weight  of  vapor.  Entrainment  is 
undesirable,  since  it  reduces  plate  efficiency  by  tending  to  destroy  the 
countercurrent  action  of  the  tower,  and  it  also  may  affect  the  distillate 
adversely  from  the  standpoint  of  color  or  othjer  nonvolatile  impurities. 

The  entrainment  of  the  liquid  is  due  to  two  main  causes:  (1)  the 
carrying  of  liquid  droplets  due  to  the  mass  velocity  of  the  gas  and  (2) 
the  splashing  of  the  liquid  on  the  plate.  These  depend  on  the  slot- 
vapor  velocity,  the  superficial  column  velocity,  and  the  plate  spacing. 

Several  investigators  have  published  quantitative  data  on  the 
amount  of  entrainment  in  bubble-plate  columns.  Most  of  their  inves- 
tigations have  been  on  systems  involving  air  and  water. 

The  data  of  Volante  (Ref.  36)  are  given  in  Fig.  16-9  where  the 
entrainment,  expressed  as  pounds  of  liquid  per  pound  of  vapor,  is 
plotted  as  a  function  of  the  superficial  velocity,  Ve,  in  feet  per  second, 
multiplied  by  the  square  root  of  the  vapor  density  in  pounds  per  cubic 
foot.  The  entrainment  increases  rapidly  with  the  vapor  throughput 
and  with  a  decrease  in  plate  spacing.  An  entrainment  of  0.01  Ib  of 
liquid  per  pound  of  vapor  does  not  seriously  lower  the  plate  efficiency 
(see  page  454),  although  it  may  give  contamination.  The  superficial 
vapor  velocity  for  these  data  is  lower  than  commercial  practice,  and 
values  of  the  abscissa  of  0.3  and  0.7  would  be  more  comparable.  Other 
data  are  given  in  Fig.  16-10.  The  curves  labeled  A  are  based  on  the 
data  of  Peavy  and  Baker  (Ref.  26)  for  the  entrainment  in  an  18-in.- 
diameter  column  with  ten  3-in.-diameter  caps  per  plate.  They  investi- 
gated the  entrainment  when  distilling  an  alcohol-water  mixture  for 
plate  spacings  of  12  and  18  in. 

Curves  B  arc  based  on  the  air-water  results  of  Sherwood  and  Jenny 
(Ref.  31).  The  tower  contained  two  plates  and  was  18  in.  in  diameter. 
Four-inch  caps  were  employed  having  33  notched-type  slots.  The 
slots  were  ^  in.  high  and  tapered  from  %Q  in.  at  the  bottom  to  %  in. 
at  the  top. 

Holbrook  and  Baker  (Ref.  16)  studied  entrainment  in  an  8-in. 
bubble-plate  column  using  steam  and  water.  Curves  C  are  based  on 
a  portion  of  their  data.  They  conclude  that  the  plate  spacing  and 
vapor  velocity  were  the  main  factors  in  determining  the  amount  of 
entrainment  and  that  the  amount  of  liquid  flow  and  slot-vapor  velocity 
were  of  less  importance. 

Curve  E  is  based  on  the  data  of  Ashraf,  Cubbage,  and  Huntington 
(Ref.  1)  for  the  entrainment  in  a  7-  by  30-ft.  commercial  absorber. 
The  tower  contained  10  trays,  22  in.  apart.  The  tower  was  operating 


426 


FRACTIONAL  DISTILLATION 


on  a  gas  oil-natural  gas  system  at  45 1  p.s.i.a.  The  investigators 
obtained  a  maximum  entrapment  of  0.0017  at  a  mass  velocity  of  23.4 
Ib.  per  min.  per  sq.  ft. 


0,1 
fc  0.01 

^late  spacing  ~ 

/-*' 

I 

ff 

/ 

A 

, 

/« 

r 

-f- 

s 

s. 

T3 

1- 

JO 

C 

1 
1 

"c  0.001 

hJ 

0.0001 

,  ^ 

j- 

/ 

J 

/ 

/ 

y 

t 

2 

°/ 

/ 

~r  — 

—  A 

1— 

/ 

f* 

/ 

J 

f 

0  01  0.2 

vc/v 

Fio.  16-9.     Entrainment  of  water  by  air.  % 

Entrainment  separators  and  baffles  have  been  suggested,  and  tests 
of  various  arrangements  have  shown  that  they  are  effective  in  reducing 
the  amount  of  liquid  carried  by  the  vapor,  but  they  have  not  been  used 
to  any  extent  in  industrial  rectifying  towers. 

Plate  Spacing.  Rectifying  columns  are  built  with  the  plates  spaced 
as  close  as  6  in.  to  as  much  as  4  to  6  ft*  There  are  a  numbersr  of  facto 


FRACTIONATING  COLUMN  DESIGN 


427 


that  influence  this  spacing,  such  as  the  proper  flow  of  vapor  and  liquid 

in  the  column  and  the  necessity  of  a  man  working  between  the  plates. 

To  obtain  the  proper  flow  of  liquid  down  the  column,  it  is  essential 


1.0 

0.8 

0.6 
04 

0.2 

|  0.08 
£    0.06 

V. 

I  004 
I- 

JO 

-g    0.02 
I 

C 

s 

c    0.01 
**  0.008 

0.006 
0.004 

0.002 

0.00  1 
0 

>4-/% 
B-M 
C-Ht 
0-Sf 
E-As 

T 

ffVffffJ 

erwoo 
tibrook 
vang- 
hrafti 
laiur 

tfBakt 
d  ana 
\anclt 
air-* 
Cubba 
a  /go* 

r-a/co 
'Jenrq 
b*er- 
wfer 
yean 
t-ya* 

ho/'M 
y  -air- 
sham 

c/Mjn 
-oil 

rfar 

/A         t* 

late 
•xncing 

wafer 
-wafa 

tingh. 

f 

p 

\ 

* 

f 

} 

/ 

f 

/c 

-t 

f- 

•/dm 
plah 
spacl 

ch  — 

r 

/ 

// 

"9 

f 

_A- 

/ 

B/i 

r 

/ 

K 

A/ 

'/] 

/ 

/ 

/, 

-4 

-A 

/ 

/ 

X 

-f- 

-y 

f/ 

•^^ 

/ 

/ 

fS 

r 

/r^ 

B/ 

X 

/*^ 

4 

^x 

^- 

^ 

<^22inchf. 
spacing 

i/ate 

2            04             0.6            0.8             1.0             12             1.4 

FIG.  16-10.     Entrainment  in  bubble-cap  plate  columns. 

that  there  be  sufficient  liquid  head  in  the  down  pipes.  Figure  16-11 
shows  schematically  the  liquid  head  in  the  down  pipes  and  its  causes. 
The  plate  spacing  must  be  great  enough  to  allow  for  the  sum  of  these 
heads  plus  some  extra  height  to  handle  short  periods  of  excess  flow. 
Values  of  hp  normally  range  from  2  to  4  in.  of  fluid;  the  various  liquid 
heads  on  the  plate  (hw  +  hcr  +  ho)  amount  to  3  to  4  in.  A  plate 


428 


FRACTIONAL  DISTILLATION 


spacing  of  6  in.  leaves  little  margin  of  safety  and  requires  the  use  of 
low  liquid  heads  on  the  plate  and  low  vapor  velocities  to  give  low 
pressure  drop  and  reasonable  entrainment.  A  plate  spacing  of  12  to 
24  in.  would  appear  to  be  more  feasible. 

In  addition  to  the  items  included  in  Fig.  16-11,  there  are  the  factors 
of  foaming  and  cross  flow  of  the  vapor.    Foaming  of  the  liquid  may  be 


hw  «•  height  of  exit  weir 

her  =  height  of  crest  on  weir 

ho  *»  liquid  gradient 

ho  **  reversal  loss  in  down  pipe 

hf  =  friction  in  down  pipe 

Ac  =*  pressure  drop  through  riser  and  inside  cap 

ha  «=  pressure  drop  through  slot 

hi  **  liquid  head  above  slot 

hp  *»  pressure  drop  through  plate 

FIG.  16-11.     Schematic  drawing  of  the  liquid  heads  in  a  bubble-plate  column. 

so  great  that  it  extends  from  one  plate  to  the  next  and  results  in  high 
pressure  drop  and  entrainment,  but  it  is  usually  not  that  serious. 
However,  a  relatively  small  amount  of  foam  may  cause  difficulties  by 
filling  the  upper  part  of  the  down  pipe  and  hindering  the  liquid  flow. 
In  many  cases,  foam  blocking  the  down  pipes  is  the  limiting  factor  in 
column  capacity.  The  foam  is  produced  largely  by  the  vapor  on  the 
plate,  and  its  plugging  effect  in  the  down  pipes  can  be  reduced  by 
including  a  short  calming  section  before  the  down  pipes  or  by  baffles 


FRACTIONATING  COLUMN  DESIGN  429 

that  will  hold  the  foam  back  on  the  plate  and  prevent  it  from  flowing 
into  the  down  pipes.  If  relatively  stable  foams  are  produced,  it  may 
be  necessary  to  add  some  foam-breaking  agent  to  the  column.  Experi- 
mental data  indicate  that  the  average  depsity  of  the  liquid  and 
entrained  vapor  mixture  is  about  one-half  that  of  the  liquid  itself. 
For  design  purposes,  it  is  therefore  desirable  to  have  a  downspout 
height  equal  to  approximately  twice  the  value  of  the  calculated  liquid 
head. 

Vapor  flow  across  the  column  in  the  vapor  space  results  from  the 
nonuniform  vapor  distribution  and  from  the  usual  reverse  direction  of 
liquid  flow  on  succeeding  plates.  On  a  given  plate  %  to  %  of  the  total 
vapor  may  flow  up  the  downstream  half  of  the  plate.  This  necessi- 
tates from  J^  to  }^  of  the  vapor  flowing  across  the  center  of  the  plate  to 
enter  the  other  half  of  the  plate  above.  Using  the  factor  of  J^,  for 
purposes  of  illustration, 

VcpHsD  =  ^~  (16-37) 

where  VCF  =  cross-flow  velocity 
Vc  =  superficial  velocity 
Hs  —  free  clearance  between  plates 
D  —  column  diameter 

VcF  *  5  W~  V* 
o  /z  s 

In  small-diameter  columns,  D/HS  is  usually  so  small  that  the  effect 
of  cross  flow  is  negligible,  but  in  large  columns  it  may  become  so  great 
that  the  kinetic  head  equivalent  to  VCF  may  be  significant  in  terms  of 
liquid  head.  Under  these  conditions,  the  pressure  in  the  vapor  space 
is  not  constant  across  the  cross  section,  and  its  variation  is  such  that  it 
forces  an  increase  in  the  hydraulic  gradient.  In  order  to  make  Ha  as 
large  as  possible,  any  beams  or  projection  on  the  bottom  of  a  plate 
should  be  positioned  to  aid  the  vapor  cross  flow.  Good  vapor  distribu- 
tion on  the  plate  will  eliminate  the  effect  of  vapor  cross  flow. 

An  effect  equivalent  to  the  action  of  vapor  cross  flow  is  frequently 
obtained  when  the  vapor  is  introduced  through  the  side  of  the  frac- 
tionating column.  For  example,  the  vapor  from  the  reboiler  is  nor- 
mally introduced  into  the  side  of  the  column  near  the  bottom.  If  this 
vapor  pipe  terminates  at  the  column  wall,  the  kinetic  energy  of  the 
vapor  may  be  so  great  that  it  will  cause  a  high  impact  pressure  on  the 
opposite  side  of  the  column.  This  type  of  action  can  result  in  very 


430 


FRACTIONAL  DISTILLATION 


poor  vapor  distribution  for  plates  up  the  column,  and  some  type  of  dis- 
tributor should  be  employed.  These  frequently  take  the  form  of  baf- 
fles which  serve  to  direct  the  vapor  over  the  whole  cross  section. 

Allowable  Vapor  Velocity.  A  factor  closely  related  to  plate  spacing 
is  the  allowable  superficial  vapor  velocity  that  can  be  employed.  The 
limiting  factor  can  be  either  the  liquid-handling  capacity  of  the  down 
pipes  or  the  loss  of  rectification  efficiency  due  to  entrainment.  The 
limiting  capacity  in  the  first  case  is  calculated  on  the  basis  of  the  pres- 


0.3 

i 

& 

§    0.2 

I- 
1 

—•* 
^ 
X 

£    O.f 

o 

**» 

•»*• 

^\ 

o" 

*~*\ 

+** 

S3" 

^ 

t** 

^ 

V** 

^ 

+** 

'^* 

-*• 

t 

Q 

l^ 

^ 

^ 

^ 

<^+ 
^ 

^ 

-y» 
r*- 



-—  ' 

£ 

^ 

s 

O^J 

^ 

^ 

^ 

*>* 

^^ 

5 

'\ 

.0    ti 

^ 

s* 

^ 

s 

/ 

/ 

^ 

/ 

S 

W 

y/ 

/ 

/ 

S 

'^ 

^ 

/ 

/ 

\/ 

/ 

/ 

/, 

/s 

/ 

^ 

'  / 

/ 

'/ 

^ 

/ 

/ 

tL^ 

/ 

/ 

'  / 

V 

/ 

V 

/ 

0                      10                     20                    30                    40                    5C 

Fm.  16-12. 


Plate  spacing,  Inches 
Allowable  superficial  velocity. 


sure  drops  and  the  liquid  heads  as  outlined  in  the  preceding  sections. 
For  the  capacity  limited  by  entrainment,  correlations  have  beeti  pre- 
sented by  Souders  and  Brown  (Ref.  33)  and  Peavy  and  Baker  (Ref .  26). 
Their  results  plus  other  data  have  been  used  to  construct  Fig.  16-12. 
The  superficial  vapor  velocity,  Fc,  in  feet  per  second  multiplied  by  the 
square  root  of  the  density  ratio,  is  plotted  as  a  function  of  the  plate 
spacing,  taken  as  the  distance  from  the  liquid  surface  to  the  plate 
above.  Lines  for  different  values  of  liquid  seals  above  the  top  of  the 
slots  are  given.  High  liquid  levels  above  the  slots  give  increased 
splashing  and  entrainment.  The  values  given  by  this  figure  are  reason- 
able design  factors  for  most  cases. 


FRACTIONATING  COLUMN  DESIGN  431 

Plate  Layout.  A  large  number  of  types  of  bubble  caps  have  been 
employed,  but  in  most  cases  circular  caps  4  to  6  in.  in  diameter  are 
used.  The  slots  are  usually  rectangular,  although  triangular  and 
trapezoidal  shapes  are  also  employed.  In  some  cases,  the  slots  have 
been  cut  through  the  caps  shell  tangentially,  but  in  the  majority  of 
cases  they  are  cut  radially.  Tunnel  caps  are  employed  in  some  cases, 
but  long  caps  of  this  type  are  sensitive  to  hydraulic  gradient  because 
the  corrective  action  of  the  pressure  drop  through  the  risers  is  less 
effective.  Hexagonal  caps  have  been  used  but  do  not  appear  to  have 
any  advantage  over  circular  ones. 

Caps  are  usually  arranged  on  triangular  centers,  and  they  are  most 
commonly  placed  so  that  the  liquid  flow  is  at  right  angles  to  the  rows 
of  caps.  This  requires  the  liquid  to  follow  a  staggered  path  across  the 
plate,  and  it  has  been  assumed  that  this  gave  good  contact  of  the  liquid 
with  the  vapor. 

Small-diameter  caps  can  give  more  slot  area  and  more  free  space  for 
liquid  flow  than  large  caps.  Caps  larger  than  6  in.  in  diameter  are 
seldom  employed  and,  except  in  small  laboratory  columns,  caps  less 
than  3  in.  are  not  used  because  of  the  mechanical  problem  of  handling 
the  large  number  of  them  needed  even  for  plates  of  moderate  diameter. 
With  the  usual  cap  design  and  plate  layout,  the  ratio  of  total  slot  area 
to  superficial  area  generally  falls  between  0.1  to  0.2. 

The  caps  should  not  be  placed  too  near  the  weirs  or  walls.  A  clear- 
ance of  2  to  3  in.  between  the  caps  and  the  weirs  and  1  to  2  in.  at  the 
walls  is  usually  adequate.  Clearances  of  1  to  2  in.  between  caps  are 
employed;  i.e.,  4-in.-diameter  caps  might  be  placed  with  centers  at  the 
corners  of  equilateral  triangles  with  sides  5%  to  6  in. 

OTHER  TYPES  OF  PLATES 

Bubble-cap  plates  are  the  most  common  plate  type  of  contacting 
device,  but  other  arrangements  are  used.  Perforated  or  sieve  plates 
are  effective  vapor-liquid  contacting  devices  and  are  frequently  used 
for  liquids  containing  suspended  solids,  such  as  the  beer  mashes  in 
alcohol  production.  At  their  rated  capacity,  their  efficiency  is  equal 
to  that  of  bubble-cap  plates,  but  they  have  two  disadvantages  that 
have  limited  their  utility.  (1)  Because  the  pressure  drop  through  the 
openings  of  the  perforated  plate  is  proportional  to  the  square  of  the 
vapor  velocity,  it  varies  .more  rapidly  with  changes  of  vapor  rate  than 
a  bubble-cap  plate,  and  this  reduces  its  flexibility.  (2)  The  liquid  on 
the  plate  can  dump  if  the  vapor  rate  is  momentarily  stopped.  Various 
arrangements  have  been  proposed  to  make  the  opening  of  the  perfora- 


432  FRACTIONAL  DISTILLATION 

tions  vary  with  the  vapor  in  a  manner  similar  to  the  slot  action  of  a 
bubble  cap,  but  their  use  has  been  limited. 

The  design  problems  of  the  weirs  and  downflow  pipes  for  perforated 
plates  are  the  same  as  those  for  bubble-cap  plates.  However,  the 
pressure  drop  due  to  vapor  flow  through  the  perforation  will  be  con- 
siderably different  than  for  a  bubble  cap.  The  small  holes  behave  like 
a  submerged  orifice,  but  their  cross  section  will  not  vary  with  the  vapor 
rate  in  a  manner  similar  to  the  slot  action  of  a  cap.  For  the  pressure 
drop  through  these  perforations,  the  following  equation  is  suggested, 

h,  =  0.04  X  +  er  7,»  (16-39) 

dpL       PL 

where  h*  =  pressure  drop  through  perforation,  in.  of  liquid 
7  «=  surface  tension,  dynes  per  pm. 
pv  =  vapor  density,  Ib.  per  cu.  ft. 
pi,  =  liquid  density,  Ib.  per  cu.  ft. 
V«  =  velocity  based  on  total  perforation  area,  f.p.s. 
d  =  diameter  of  perforation,  in. 

The  total  pressure  drop  across  the  plate  will  be  equal  to  hv  plus  the 
liquid  head  on  the  plate.  The  problem  of  calculating  the  liquid  head 
is  similar  to  that  for  a  bubble-cap  plate,  and  the  actual  depth  at  the  exit 
weir  is  a  good  approximation  in  most  cases.  The  hydraulic  gradient 
would  increase  the  liquid  head,  but  the  aeration  effect  would  reduce  it. 

The  allowable  vapor  velocities  for  perforated  plates  should  be  com- 
parable to  those  for  bubble-plate  trays,  and  Fig.  16-12  should  give 
suitable  design  values.  Perforations  from  0.1  to  0.5  in.  are  employed, 
and  they  are  usually  placed  on  triangular  centers  spaced  such  that  the 
total  area  of  the  openings  is  0.10  to  0.15  times  the  tower  cross-sectional 
area.  The  larger  openings  have  more  of  a  tendency  to  by-pass  liquid 
and  give  plate  dumping  but  are  less  likely  to  become  plugged.  Except 
for  the  reduction  of  liquid  head  due  to  aeration,  dumping  should  occur 
when  hr  for  the  downstream  row  of  perforations  is  equal  to  the  hydraulic 
gradient.  Actually  the  dumping  will  occur  before  the  gradient  is  this 
great,  owing  to  aeration  effects  and  surges. 

Spray-type  plates  have  been  described  for  low-pressure  operations. 
In  order  to  obtain  high  rates  of  mass  transfer,  the  liquid  is  collected 
periodically  and  resprayed.  Kraft  (Ref.  21)  has  described  shower-type 
trays  in  which  the  liquid  is  allowed  to  rain  down  from  one  plate  to  the 
next  through  small  perforations,  and  the  vapor  flows  across  the  shower 
but  does  not  bubble  through  the  liquid.  These  plates  are  reported  to 
give  pressure  drops  of  0.75  mm.  Hg  per  plate  at  practical  operating 


FRACTIONATING  COLUMN  DESIGN  433 

rates.  This  is  about  one-half  the  pressure  drop  of  the  best  bubble-cap 
plates.  Disk  and  doughnut-type  plates  have  also  been  used,  and 
these  also  pass  the  vapor  through  a  liquid  shower. 

PLATE  DESIGN  EXAMPLE 

As  an  example  of  the  use  of  the  methods  outlined  in  this  chapter,  consider  the 
design  of  a  bubble  plate  for  a  column  to  handle  a  benzene-toluene  mixture.  For  the 
separation  20  actual  plates  will  be  used,  and  the  plate  design  will  be  based  on  the 
bottom  plate.  This  plate  will  handle  essentially  pure  toluene,  and  the  liquid  and 
vapor  rates  will  be  480  and  400  Ib.  mols  per  hr.,  respectively. 

Data  and  Notes. 

The  pressure  at  the  bottom  plate  will  be  2.5  p.s.i.g. 

Use  chord  overflow  weir. 

Cap,  4  in.  O.D.  by  3><j  I.D.  by  4Ji6  in.  high  cast  iron. 

Caps  placed  on  equilateral  triangle  centers  of  6  in.,  3-in.  minimum  clearance  at 

weirs  and  1-in.  minimum  clearance  at  walls. 
Slots,  33  per  cap,  %  by  1^  in. 
Minimum  seal  on  slots  at  no  flow,  J4  in* 
Minimum  liquid  depth  at  no  flow,  2  in. 
Minimum  down-pipe  seal,  1J4  in. 

Plate  will  have  cross  flow  with  only  one  down  pipe  per  plate. 
Plate  spacing,  24  in. 

Solution.  The  seal  on  the  caps  at  no  flow  is  K  in->  and  the  head  on  the  weir  at 
operating  conditions  will  be  taken  as  1.0  in.,  making  a  total  liquid  seal  at  the  outlet 
of  \y±  in.  A  hydraulic  gradient  of  0.1  in.  is  assumed,  making  the  inlet  seal  1.35  in. 
An  average  seal  of  1.3  in.  will  be  used  with  Fig.  16-12,  to  determine  the  allowable 
superficial  velocity.  The  plates  are  spaced  24  in.  apart,  but  this  will  be  reduced  to 
20.5  in.  to  allow  for  liquid  level  and  plate  thickness. 

From  Fig.  16-12, 

jz—Y  -  °-164 

i  -  PV) 

At  2.5  p.s.i.g.  the  boiling  point  of  toluene  is  694°R.,  and  pL  «  52.8  Ib.  per  cu.  ft. 
Assuming  the  perfect  gas  laws  apply  to  toluene  vapor, 

92 


pv      359(<^92)(14.7/17.2) 
«*  0.212  Ib.  per  cu.  ft. 

Allowable  vapor  velocity, 


=  2.6  f.p.s. 

48'2 


Net  area  of  plate  -  ^  -  18.6  sq  ft. 


434  FRACTIONAL  DISTILLATION 

Allow  2  aq.  ft.  for  the  two  down-pipe  areas 

~20lT 


D  - 


Weir  Length.    By  Eq.  (16-12), 


1.0  -  5.4 


5.2  ft. 


f    480(92)    "IH 


L53(3,600)LJ 
and,  using  a  factor  of  0.96  from  Fig.  16-5, 

480(92) 


L  -  [5.4(0.96)P-6 
«  2.73  ft. 


53(3,600) 


In  the  text  it  was  suggested  that  the  weir  within  a  weir  head  of  the  wall  should 
be  neglected,  and  the  actual  weir  length  will  be  longer  than  just  calculated.  The 
following  sketch  shows  these  conditions: 


+  26.8  -  912 
A  -  25.8  in. 
C  -  5.4  in. 


A*  +  B* 


~(¥x12)2 

Ba  -  974  -  644 
B  -  18.2  in. 


974 


Length  of  weir  -  2  f ~~  J  -  3.03  ft. 
Let  </>  **  angle  subtended  by  weir  chord 

sinl^^  »o.582 

<f>    sac   71  2° 

A         ,  j                      »  /*  OM  /71.2\       3.03(5.2/2)  cos  (0/2) 
Area  of  downcomer  «•  |  (5.2)a  (  355  ) " — ^ ^ZL- • 

-  1.0  sq.  ft. 

Total  area  of  downcomers  on  each  side  of  plate  equals  2  sq.  ft,,  which  is  the 
assumed  value. 


FRACTIONATING  COLUMN  DESIGN 


435 


A  schematic  plate  layout  is  shown  in  Fig,  16-13.  The  center  line  of  the  row  of 
caps  nearest  the  weir  is  25.8  —  5  in.  from  the  center  with  a  length  of  46.6  in.  The 
number  of  caps  in  this  row  will  be  46.6/6,  or  7  caps.  The  distance  between  the 
inlet  and  outlet  rows  of  caps  is  41.6  in.,  and  this  will  allow  9  rowsof  caps  with  1,6  in. 
over  which  will  be  used  for  the  inlet  weir.  The  total  number  of  caps  is  77. 


Weep  holes  —  - - 


/Down flow  pfpe  to 
X>SSs^  Pfaf*  befow 


O  O  O  O  O  O  ON 

oooooooo 
ooooooooo 

OOOOOOOOOOl 

ooooooooo 

OOOOOOOOOOi 

ooooooooo 

oooooooo 

ooooooo 


FIG.  16-13. 


*-//7/e/  weir 

*^  ~  Down  f low  pipe 
f romp/ate  above 
Layout  of  bubble-cap  plate. 


Hydraulic  Gradient*  The  section  with  the  bubble  caps  approximates  a  rec- 
tangular section,  and  Eq.  (16-16)  will  be  employed  for  the  calculation  of  the 
hydraulic  gradient. 

The  average  foam  velocity  by  Eq.  (16-19)  is 


Vt 


3(480)  (92) 
3600(53)  (0.5)  (4.6) 
0.30  f  .p.s. 


The  value  of  6  *»  4.6  feet  was  taken  as  the  average  of  the  diameter  of  the  column 
md  the  width  at  the  first  row  of  caps. 
By  Eq,  (16-17), 


ind  from  Eq.  (16-18),  using  the  viscosity  of  the  liquid  equal  to  0,00067  (see  p,  418), 


Re' 


0.41  (0.8)  (53) 
0.00067 


9730 


436  FRACTIONAL  DISTILLATION 

Using  this  value  with  Fig.  16-7  gives 

f  (     *•>     V*  «  0.3 

\h0  —  hap) 


The  loss  in  head  is  calculated  from  Eq.  (16-16) 

(0.24)  (0.30)  2(3.7) 
*  32.2(0.41) 

«  0.0061  ft.  of  toluene 
=  0.073  in.  of  toluene 
*  0.062  in.  of  water 

The  gradient  for  this  plate  is  somewhat  smaller  than  the  assumed  value  of  0.1  in. 
of  toluene.     However,  the  difference  is  not  large  and  no  corrections  will  be  made. 

Pressure  Drop  through  Caps. 

1.  Assume  uniform  vapor  distribution  with  no  gradient, 

48  2 
,Cu.  ft.  vapor  per  second  per  cap  =  ~-  ^  0.626 

a.  Pressure  drop  through  risers  and  inside  cap 

V    =     °-626(144) 
R       2.352(0.7854) 

By  Eq.  (16-2), 


=  1.9  in.  of  toluene 
6.  Pressure  drop  through  slots 

At  694°R.,  T  for  toluene  =  18  dynes /cm. 

_  0.626(144) 
v.  -        6  2        -  14.0 

By  Eq.  (16-4), 


This  compares  with  a  value  of  0.7  for  a  slightly  different  4-in.  cap  given  on 
page  407.    To  be  conservative  use  0.7. 
From  Eq.  (16-3), 

h.  -  0.12(i%3)  +  0.7  [  1.5(14.5 

-  0.04  +  0.86 

»  0.90  in.  of  toluene 

c.  Liquid  head  above  slots  «•  1.25  in.  of  toluene. 


FRACTIONATING  COLUMN  DESIGN  437 

d.  Pressure  drop  for  ideal  plate  with  no  liquid  gradient, 

hp  «  1.9  -f  0.90  +  1.25 
»  4.05  in.  of  toluene 

2.  Pressure  drop  for  actual  plate,  , 

hv  -  4.05  -  0.04  -  1.25 
»  2.76  in.  of  toluene 

The  value  of  the  gradient  is  so  small  relative  to  hy  that  Eq.  (16-34)  indicates 
essentially  uniform  vapor  distribution,  and  the  pressure  drop  will  be  the  same  as 
calculated  for  an  ideal  plate. 

Height  of  Liquid  in  Down  Pipe. 

1.  Loss  in  head  at  bottom  of  downpipe.     Assume  the  chord  down  pipe  comes 
within  1.0  in.  of  plate.     For  Eq.  (16-13), 

092f.p.s. 

hD  -  0.5(0  92) 2 

=  0.42  in.  of  toluene 

2.  Friction  loss  in  down  pipe.     Because  the  velocity  in  the  down  pipe  is  only 
0.23  f.p.s.,  the  friction  will  be  negligible. 

3.  Head  of  liquid  in  down  pipe  above  plate  level  (see  Fig.  16-11). 

JiDp  «  2.0  +  1.0  +  0.07  -f  0.42  +  0  +  4.05 
—  7. 54  In.  of  toluene 

The  plate  spacing  of  24  in.  should  be  adequate  unless  foaming  is  excessive. 

General  Consideration.  The  general  arrangement  of  the  plate  is  shown  in  Fig 
16-13.  Weep  holes  have  been  added  to  allow  the  column  to  drain  when  shut  down. 
Three  %-in.  weep  holes  were  placed  in  the  plate  before  the  exit  weir.  These  should 
allow  the  column  to  drain  in  about  5  hr.  At  rated  load,  the  weep  holes  handle 
about  2.5  per  cent  of  the  total  liquid,  but  even  this  amount  does  not  short-circuit 
because  it  flows  down  on  the  proper  side.  These  weep  holes  can  be  placed  in  the 
inlet  and  outlet  weirs  instead  of  the  plate. 

PACKED  TOWERS 

Pressure  Drop.  Although  the  pressure  drop  through  packed  towers 
is  usually  small  at  atmospheric  pressure,  it  may  become  a  limiting  fac- 
tor in  vacuum  distillations. 

Chilton  and  Colburn  (Ref .  4)  have  published  a  method  for  predicting 
such  pressure  drop  for  solid  packings,  based  on  the  Fanning  equation 
for  friction  in  pipes.  They  modify  the  friction  equation  to 


438 


FRACTIONAL  DISTILLATION 


where  Ap  =  pressure  drop  in  height  h 

Aw  =  correction  factor  for  wall  effect 

A  i  «  correction  factor  for  wetting  of  the  packing,  by  the  liquid 
pv  =  density  of  the  vapor 
g  =  conversion  factor  for  consistent  units 
u  «  superficial  gas  velocity,  i.e.,  linear  gas  velocity  based  on 

the  total  cross  section  of  the  tower 
d*  «  size  of  packing,  nominal 
ji  »  viscosity  of  vapor 

/  =  function  of  (d*upfii)  as  given  in  Eqs.  (16-41)  and  (16-42) 
Chilton  and  Colburn  gave  a  plot  of  /  as  a  function  of  the  Reynolds 
number  (d*up//i).    The  data  of  this  plot  may  be  approximated  by  the 
following  equations: 

For  (d*up/n)  less  than  40,  use 


For  (d*up/fji)  greater  than  40, 


850 

d*Ufi/IJL 


38 


(16-41) 


(16-42) 


All  these  equations  are  dimensionally  sound,  and  any  consistent  set 
of  units  may  be  used.  The  following  table  contains  values  of  the 
factor  Aw: 


Aw 

Packing  diameter 

Tower  diameter 

(*?)<« 

(*?)>• 

0 

1.0 

1  0 

0.1 

0.83 

0  72 

0.2 

0.74 

0.65 

0.3 

0  71 

0.57 

The  correction  for  the  wetting  of  the  packing,  AL,  is  1.0  for  dry  pack- 
ing and  is  greater  than  1.0  for  wet  packing  because  the  liquid  decreases 
the  free  volume.  The  factor  is  greater  than  1,0  for  wet  packing  even 
without  liquid  flow,  and  it  increases  with  the  liquid  rate.  As  an 
approximation,  the  following  relation  is  suggested  for  AL  for  wet 


FRACTIONATING  COLUMN  DESIGN  439 

packing, 

AL  =  1.4  +  0.0005F  (16-43) 

where  F  ==  liquid  rate,  Ib.  per  hr.  per  sq.  ft. 

Liquid  rates  above  5,000  to  10,000  Ib.  per  h».  per  sq.  ft.  give  values  of 

AL  so  large  that  the  tower  may  flood. 

The  pressure  drop  with  hollow  packings  is  less  than  given  by  the 
equation  for  solid  packings.  The  data  on  this  effect  are  not  very  con- 
clusive as  to  absolute  magnitude  but  do  indicate  that,  for  hollow  pack- 
ing, the  smaller  the  packing  size  the  larger  the  pressure  drop.  For  a 
detailed  discussion  of  this  factor,  the  reader  is  referred  to  Sherwood's 
summary  (Ref,  30). 

Allowable  Gas  and  Liquor  Velocities.  The  capacity  of  packed 
towers  is  limited  by  the  tower's  becoming  flooded  with  liquid.  The 
flooding  can  be  caused  by  increasing  either  the  liquid  or  the  gas  flow. 
This  flooding  is  a  result  of  the  pressure  drop  through  the  tower  exceed- 
ing the  gravity  head  of  the  liquid  flowing  down. 

These  pressure  drops  per  foot  of  height  are  given  by  the  modified 
Fanning  equation: 


<~    2gm 

where  /  =  proportionality  factor,  a  function  of 

Vo  =  actual  linear  gas  velocity  =  uG/FAo,  f.p.s. 
VL  =  actual  linear  liquid  velocity  =  UL/FAL,  f.p.s. 
p  =  density,  Ib.  per  cu.  ft. 

g  =  conversion  factor—  32.2  (Ib.  force)  (ft.)/(lb.  mass)(sec.2) 
n  =  viscosity,  poises 

F  =  fraction  of  tower  that  consists  of  voids1 
S  =  surface  of  packing,1  sq.  ft.  per  cu.  ft.  of  packing 
m  =  hydraulic  radius  =  free  volume  /contact  area  =  F/S,  ft. 
A0i  AL  =  fraction  of  free  cross  section  occupied  by  gas  and  liquid, 

respectively 

u  =  superficial  velocity,  f.p.s. 

The  flooding  occurs  when  Ap*  =  PL  —  Apa  is  a  small  fraction  of  Apo, 
since  under  such  conditions  a  slight  increase  in  the  rate  of  flow  of  either 
stream  or  an  uneven  surge  in  the  tower  will  increase  AL  and  decrease 
Ao,  because  of  increased  liquid  holdup.  This  decrease  in  Ao  will 
increase  VG  and  thereby  Ap<?.  If  Api,  is  large  relative  to  Ap*,  this 

1  For  data  on  the  values  of  S  and  F  for  various  packings,  the  reader  should  con- 
sult Ref  s.  9  and  30  at  the  end  of  the  chapter. 


440 


FRACTIONAL  DISTILLATION 


X-fe- 


^ 


*3" 


& 


0, 

e 


4 


FRACTIONATING  COLUMN  DESIGN 


441 


increase  in  Ape  will  affect  the  liquid  flow  only  slightly;  however,  if 
is  small  compared  to  Ap<?,  a  small  increase  in  Ap<?  will  make  a  large  per- 
centage decrease  in  ApL,  causing  the  holdup  to  increase  and  the  tower 
to  flood. 
Combining  the  pressure-drop  equations, 


Setting  the  ratio  (PL  —  Apc?)/Ap(?  equal  to  b  and  noting  that  at  flooding 
Ap<?  becomes  approximately  equal  to  PL  modifies  the  previous  equation 
to  give 

'Vu?-,  (16-44) 


At  low  values  of  (UG/UL^PV/PL,  the  denominator  of  the  right-hand  side 
becomes  1,  and  the  limiting  gas  velocity  is  a  function  of  (UQ/UL)(PV/PL), 
the  tower  dimensions,  and/L,  the  latter  term  being  chiefly  a  function  of 
PL.  It  is  interesting  to  note  that  in  this  region  the  gas  viscosity  is  not 
a  factor;  but  when  the  last  term  of  the  denominator  is  not  negligible, 
the  viscosity  of  the  gas  becomes  a  factor  in  the  limiting  gas  velocity. 
At  very  high  values  of  this  group,  the  right-hand  side  reduces 

TABLE  16-2 


Packing 


No.  19  aluminum  jack  chain 
1-in.  Raschig  rings  .     . . 
1-in.  Berl  saddles. . 
8-mm.  Raschig  rings . . . 
J^-in.  Raschig  rings. 

J^j-in.  Raschig  rings       

J^-in.  Raschig  rings  . .   . . 

J«2-in.  Berl  saddles 

3^-in.  Raschig  rings  

J^-in.  Rachig  rings      

J^-in.  Raschig  rings 

}£-in.  Raschig  rings  

H2-itt-  carding  teeth . 
M-in.  bent  carding  teeth 
0.23-  by  0.27-in.  glass  rings. . 

0.18-in.  glass  rings 

0.47-in.  glass  rings 


System 


Heptane-methyl  cyclohexane 
Air-water 
Air-water 
Air-water 
H2-water 
Air- water 
COrwater 
Air-water 
Air-methanol 

Air- (50%  H2O  -f  50%  CH3OH) 
Air-glycerin 

Water  -f-  butyric  acid-air 
Benzene-carbon  tetrachloride 
Benzene-carbon  tetrachloride 
Benzene-carbon  tetrachloride 
Ethanol-water 

Quinoline  (distillation  at  10  mm. 
Hg  abs.) 


Reference 


8 

23 
23 
23 
32 
32 
32 
32 
32 
32 
32 
32 
9 
9 
9 
17 

17 


442  FRACTIONAL  DISTILLATION 

This  would  be  the  case  when  a  very  low  liquid  rate  was  employed  and 
the  gas  occupied  essentially  the  whole  free  cross  section  of  the  tower. 
For  convenience  in  plotting,  the  right-hand  side  will  be  taken  as  a  func- 
tion of  (UQ/UL)  (PV/PL)H/I**L.  The  data  of  a  number  of  investigators  are 
correlated  in  this  way  in  Fig.  16-14  using  a  value  of  a  =  0.21.  The 
systems  involved  are  given  in  Table  16-2. 

The  data  are  seen  to  correlate  well  except  at  high  values  of  the 
abscissa.  This  deviation  may  be  due  to  the  fact  that  the  f0  factor  in 
the  denominator  is  neglected  in  the  method  of  plotting  of  this  figure. 

Nomenclature 

AL  **  correction  factor  for  liquid  on  packing  in  packed  tower 

A«  —  area  under  skirt  of  caps  for  liquid  flow 

Aw  **  correction  factor  for  wall  effect  in  packed  towers 

B  —  length  of  bubbling  section,  ft. 

6  *•  width  of  plate,  ft. 

C  «  pressure  recovery  factor  for  orifice 

D  «  inside  diameter,  ft.  ' 

d  «•  diameter,  in. 
d*  «  size  of  packing,  nominal 

F  »  loss  in  head  from  inlet  to  exit  calming  sections,  ft. 

/  «  friction  factor  for  fluid  flow 

/'  «•  friction  factor 
g,gc  «•  conversion  constant  -  32.2  [(ft.)(lb.force)]/[(sec.2)  (Ib.  mass)] 

h  «•  liquid  depth  above  surface  of  plate 
he  «•  pressure  drop  through  riser  and  undercap,  in.  of  liquid 
her  **  height  of  liquid  above  top  of  weir,  in. 

ho  ••  pressure  drop  for  reversal  of  flow  at  bottom  of  down  pipe,  in.  of  liquid 
hop  »  liquid  level  in  down  pipes  above  surface  of  plate,  in. 
ha  «.  kinetic  head,  in.  of  liquid 
1  hi  «  liquid  depth  above  surface  of  plate  at  inlet  weir 
hi  «*  liquid  depth  above  top  of  slots,  in. 
h*L  **  value  of  KL  at  inlet  row  of  caps 
hi  **  value  of  hi  at  outlet  row  of  caps 
h0  «•  liquid  depth  above  surface  of  plate  at  outlet  weir,  in. 
h*  »  slot  height,  in. 

h*  «•  value  of  h,  for  slots  completely  open 
hj  »  over-all  pressure  drop  for  plate,  in.  of  liquid 
h^  »  over-all  pressure  drop  for  uniform  vapor  distribution,  in.  of  liquid 
hp  —  over-all  pressure  drop  when  inlet  row  of  caps  is  just  inactive,  in,  of  liquid 
Ht  «*  free  clearance  between  plates^  ft. 
h,  **  pressure  drop  through  slots,  in.  of  liquid 
htp  —  height  of  top  of  slot  above  surface  of  plate 
hy  •»  velocity  head  for  uniform  vapor  distribution,  in.  of  liquid 
hw  ««  height  of  weir  above  surface  of  plate 
h*  •»  pressure  drop  through  perforations*  in.  of  liquid 
K,k  —constants 


FRACTIONATING  COLUMN  DESIGN  443 

L  *  perimeter  of  weir,  ft. 
Lo  «  foam  height,  ft. 
m  «*  exponent 

N  «  number  of  rows  of  caps  from  outlet  to  inlet  weir 
n  «•  number  of  rows  of  caps,  counting  from  the  overflow  weir 
Q  »  liquid  flow  rate,  cu.  ft.  per  sec. 
Qw  =  liquid  rate,  Ibs.  per  sec. 
q  =  liquid  flow  rate,  g.p.s. 

> 

'•'" ' 

/*/ 

° 


mean  hydraulic  radius,  ft. 


u  =*  superficial  velocity  in  packed  tower,  f ,p.s. 
UQ  «•  superficial  velocity  of  gas 
UL  «  superficial  velocity  of  liquid 

V  —  vapor  flow  rate 
Ve  =*  superficial  vapor  velocity,  f .p.s. 
VCF  =  cross  flow  velocity,  f.p.s. 

VD  •»  maximum  velocity  at  bottom  of  down  pipe,  f  .p.s. 
V/  =  velocity  of  liquid  in  foam,  f.p.s.  «=  Qw(p/L0b) 
Vo  —  linear  gas  velocity  in  packed  tower,  f.p.s. 
VL  *•  linear  liquid  velocity  in  packed  tower,  f  ,p.s. 
V0  *»  liquid  velocity  in  down  pipe,  f  .p.s. 
VR  —  maximum  velocity  under  cap,  f.p.s. 
V9  «•  slot  velocity  based  on  total  slot  area,  f.p.s. 
V%  =•  slot  velocity  for  ha  equal  to  h* 
VT  «*  total  vapor  flow  rate  for  plate 
Vjf  «•  velocity  based  on  total  perforation  area 

7  ss  surface  tension,  dynes  per  cm. 
Pf  ~  density  of  foam,  Ibs.  per  cu.  ft. 
pv  —  vapor  density 
PL  =*  liquid  density 

0  »  viscosity 
/i/  =  viscosity  of  foam 
ML  »  normal  viscosity  of  liquid,  Ibs.  per  f.p.s. 

References 

1.  ASHBAF,  CUBBAGE,  and  HXJNTINGTON,  Ind.  Eng.  Chem.,  26, 1068  (1934). 

2.  BIJAWAT,  S.M.  thesis  in  chemical  engineering,  M.I.T.,  1945. 
2a.  BLOECHER,  S.B.,  thesis  in  chemical  engineering,  M.I.T.,  1949. 

3.  CAREY,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1929. 

4.  CHILTON  and  COLBURN,  Trans.  Am.  Inst.  Chem.  Engrs.,  26,  178  (1931). 

5.  DAUPHINE,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1939. 

6.  DAVIES,  Ind.  Eng.  Chem.,  39,  774  (1947). 

7.  EDMINSTER,   "Hydrocarbon  Absorption  and  Fractionation  Process  Design 
Methods,"  reprinted  from  The  Petroleum  Engineer. 

8.  FENSKE,  LAWROSKI,  and  TONGBERG,  Ind.  Eng.  Chem.,  30, 297  (1938). 

9.  FENSKB,  TONGBERG,  and  QTTIGGLB,  Ind.  Eng.  Chem.,  26, 1169  (1934). 


444  FRACTIONAL  DISTILLATION 

10.  GARDNER,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1946. 

11.  GHORMLBY,  S.M.  thesis  in  chemical  engineering,  M.I.T.,  1947. 

12.  GONZALES  and  ROBERTS,  S.M.  thesis  in  chemical  engineering,  1943. 

13.  GOOD,  HUTCHINSON,  and  ROUSSEAU,  Ind.  Eng.  Chem.,  34, 1445  (1942). 

14.  GRISWOLD,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1931. 

15.  HARRINGTON,  BRAGG,  and  RHYS,  Petroleum  Refiner,  24,  502  (1945). 

16.  HOLBROOK  and  BAKER,  Ind.  Eng.  Chem.,  26,  1063  (1934). 

17.  JANTZEN,  Dechema  Mon.,  6,  No.  48  (1932). 

18.  KEMP  and  PYLE,  Chem.  Eng.  Progress,  45,  435  (1949). 

19.  KESLER,  S.M.,  thesis  in  chemical  engineering,  M.I.T.,  1949. 

20.  KIRKBRIDE,  Petroleum  Refiner,  23,  321  (1944). 

20a.  KLEIN,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1950. 

21.  KRAFT,  Ind.  Eng.  Chem.,  40,  807  (1948). 

22.  LOCKE,  S.M.  thesis  in  chemical  engineering,  M.I.T.,  1937. 

23.  MACH,  Forschungsheft,  375,  9  (1935). 

24.  MAYER,  S.M.  thesis  in  chemical  engineering,  M.I.T.,  1938. 

25.  10.45  Notes  M.I.T.,  1945. 

26.  PEAVY  and  BAKER,  Ind.  Eng.  Chem.,  29,  1056  (1937). 
26a.  ROGERS  and  THIELE,  Ind.  Eng.  Chem.,  26,  524  (1934). 

27.  ROWLEY,  S.B.,  thesis  in  chemical  engineering,  M.I.T.,  1938. 

28.  SCHNEIDER,  S.M.  thesis  in  chemical  engineering,  M.I.T.,  1938. 

29.  SEUREN,  S.M.  thesis  in  chemical  engineering,  M.I.T.,  1947. 

30.  SHERWOOD,  "Absorption  and  Extraction,"  p.  141,  McGraw-Hill  Book  Com- 
pany, Inc.,  New  York,  1937. 

31.  SHERWOOD  and  JENNY,  Ind.  Eng.  Chem.,  27,  265  (1935). 

32.  SHIPLEY,  S.M.  thesis  in  chemical  engineering,  M.I.T.,  1937. 

33.  SOUDE;RS  and  BROWN,  Ind.  Eng.  Chem.,  26,  98  (1934). 

34.  SOUDERS,  HUNTINGTON,  CoRNEiL,  and  EMERT,  Ind.  Eng.  Chem.,  30,  86  (1938). 

35.  STRANG,  Trans.  Inst.  Chem.  Engrs.,  12,  169  (1934). 

36.  VOLANTE,  S.B.  thesis  in  chemical  engineering,  M.I.T.,  1929. 


CHAPTER  17 
FRACTIONATING  COLUMN  PERFORMANCE 

The  design  calculations  considered  in  the  preceding  chapters  were 
based  on  theoretical  plates.  In  order  to  complete  the  design,  it  is 
necessary  to  have  the  relationship  between  these  idealized  values  and 
the  actual  performance  of  the  contacting  device.  The  vapor  and 
liquid  brought  into  contact  with  each  other  in  the  tower  are  not  at 
equilibrium,  and  the  rate  of  mass  transfer  determines  the  effectiveness 
of  the  unit.  This  chapter  will  consider  the  methods  of  predicting  the 
effectiveness  of  the  vapor-liquid  contact  for  the  various  types  of  units. 

PLATE-TYPE  COLUMNS 

Plate  Efficiency  Definitions  and  Relations.  Over-all  Column 
Efficiency.  The  relation  between  the  performance  of  actual  and  theo- 
retical plates  is  expressed  as  plate  efficiencies.  A  number  of  different 
plate  efficiencies  have  been  proposed,  but  the  two  most  commonly  used 
are  the  "over-all  column  efficiency"  which  was  proposed  by  Lewis 
(Ref.  20)  and  "plate"  or  "point"  efficiencies  suggested  by  Murphree 
(Ref.  24).  The  over-all  column  efficiency,  E°,  is  the  number  of  theo- 
retical plates  necessary  for  a  given  separation  divided  by  the  number  of 
actual  plates  required  to  perform  the  same  separation;  i.e.,  it  is  the 
factor  by  which  the  number  of  theoretical  plates  is  divided  to  give  the 
actual  number  of  plates.  This  efficiency  has  no  fundamental  mass- 
transfer  basis,  but  it  serves  as  an  easily  applied  and  valuable  design 
factor  and  is  therefore  widely  used.  v , 

Murphree  Efficiencies.  The  Mur- 
phree efficiencies  are  based  on  more 
fundamental  concepts  than  E°y  but 
even  in  this  case  the  basic  relations 
employed  are  more  qualitative  than 
quantitative.  Murphree  developed 
two  cases:  one  employing  vapor-phase  relations  and  the  other  liquid- 
phase  conditions.  For  the  vapor-phase  derivation  he  assumed  that 
a  bubble  of  vapor  in  rising  through  the  liquid  on  a  plate  was  in  con- 
tact with  a  liquid  of  constant  composition,  and  that  the  composition 

445 


446  FRACTIONAL  DISTILLATION 

in  the  bubble  changed  continuously  by  mass  transfer.  Consider  the 
bubbles  shown  in  the  simplified  schematic  diagram  of  Fig.  17-1.  These 
bubbles  enter  with  a  composition  y*  pass  up  through  a  liquid  of  com- 
position x,  and  leave  with  a  composition  yf0.  Using  a  simplified  expres- 
sion for  the  instantaneous  mass  transfer  to  one  of  the  bubbles, 

-(?  dy'  -  PK0a(yf  -  y.)  de  (17-1) 

where  G  =  mols  of  gas  in  bubble 

yf  =  mol  fraction  of  component  in  bubble 
2/«  =  vapor  in  equilibrium  with  liquid 
P  =  total  pressure 
KQ  ~  over-all  mass-transfer  coefficient,  (mols)/(unit  time)  (unit 

pressure  difference)  (unit  interfacial  surface) 
a  =  interfacial  area  of  bubble 
0  =  contact  time  of  bubble  with  liquid 

Assuming  (?,  P,  ye,  and  K&a  are  constant,  this  equation  can  be 
integrated  to  give 

PK0ae 
G 

Murphree  applied  the  relation  to  the  whole  plate,  assuming  that  con- 
ditions were  the  same  at  all  points.     In  this  case  (G/0)  is  replaced  by 
the  vapor  rate,  and  a  becomes  the  average  total  interfacial  area  of  all 
the  bubbles  on  a  plate  at  any  instant. 
This  equation  can  be  rearranged  to 


where  m 


EMV  -  **  __  f  °  _  1  -  e~m  (17-3) 

PKGaO 


G 

In  many  cases  there  are  considerable  differences  in  the  compositions 
of  the  liquids  at  various  points  on  a  plate,  and  the  conditions  assumed 
in  the  derivation  are  not  satisfied  for  the  whole  plate.  However,  for 
convenience,  the  value  of  the  Murphree  plate  efficiency  is  defined  as 

(17-4) 

^ 

where  y»-,  y*  *  average  composition  of  vapor  entering  •  and  leaving 
plate,  respectively  '  ' 

y*  =*  composition  of  vapor  in  equilibrium  witji  liquid  flowing 
to  plate  below 


FRACTIONATING  COLUMN  PERFORMANCE  447 

Equations  (17-3)  and  (17-4)  appear  similar,  but  the  former  uses  the 
values  for  a  small  region  of  the  plate,  while  the  latter  uses  average 
values  for  the  various  streams  entering  and  leaving  a  plate.  U£F  is 
not  equal  to  1  —  e~m. 

The  derivation  of  Eq.  (17-3)  should  apply  to  a  limited  region  of  the 
plate,  and  this  has  been  termed  the  Murphree  point  efficiency,  E%v. 
In  this  case  " 

(17-5) 

where  y',  y0  =  vapor  compositions  entering  and  leaving  the  local 

region 
y<  =  vapor  composition  in  equilibrium  with  the  liquid  in  the 

local  region 

"the  Murphree  plate  efficiency  is  the  integrated  effect  of  all  the 
Murphree  point  efficiencies  on  the  plate. 

The  derivation  of  the  Murphree  equation  is  based  on  a  very  qualita- 
tive picture  of  the  mass  transfer.  At  low  vapor  rates  individual  bub- 
bles are  obtained,  but  a  study  of  mass  transfer  for  such  systems  indi- 
cates that  it  cannot  be  expressed  as  a  simple  rate  equation  involving  a 
constant  multiplied  by  a  driving  force  in  mol  fraction  units.  The 
experimental  data  indicate  that  mass  transfer  is  very  rapid  while  the 
bubble  is  being  formed  and  then  is  relatively  slow  while  the  bubble  is 
rising  through  the  liquid.  At  higher  vapor  rates  channels  are  blown 
through  the  liquid,  and  a  large  quantity  of  .spray  is  thrown  up  into  the 
vapor  space  giving  additional  mass  transfer.  The  simple  rate  Eq. 
(17-1)  cannot  be  any  more  than  a  crude  expression  of  the  phenomena 
involved,  and  K <?a  must  be  a  complicated  function  of  a  large  number  of 
variables  including  by  and  0.  y 

The  value  of  E°f  E°MV,  and  E %v  are  the  most  commonly  used  design 
factors  for  plate  efficiencies.  The  over-all  tower  plate  efficiency  is 
simpler  to  apply  than  the  Murphree  efficiencies  because  only  terminal 
conditions  are  required;  whereas  in  the  calculation  of  the  Murphree 
plate  efficiency,  plate-to-plate  compositions  are  required,  and  for  the 
Murphree  point  efficiency,  complete  liquid-  and  vapor-composition 
traverses  are  required  on  each  plate.  However,  the  Murphree  effi- 
ciencies are  probably  on  a  more  fundamental  basis  than  the  over-all 
efficiencies. 

By  definition,  a  theoretical  plate  is  one  on  which  the  average  compo- 
sition of  the  vapor  leaving  the  plate  is  the  equilibrium  value  for  the 
liquid  leaving  the  plate.  If  the  vapor  and  liquid  upon  a  plate  were 


448  FRACTIONAL  DISTILLATION 

completely  mixed,  it  would  be  impossible  to  obtain  better  separation, 
than  that  given  by  a  theoretical  plate.  However,  when  there  is  a  con- 
centration gradient  in  the  liquid  across  the  plate,  the  average  concen- 
tration of  the  more  volatile  component  in  the  liquid  on  the  plate  may 
be  appreciably  greater  than  the  concentration  of  the  liquid  leaving  the 
plate;  as  a  result  of  this  greater  concentration,  the  vapor  actually  leav- 
ing the  plate  may  exceed  the  concentration  of  the  vapor  in  equilibrium 
with  the  liquid  leaving.  It  is  thus  possible  for  the  concentration-gradi- 
ent effect  to  give  over-all  and  Murphree  plate  efficiencies  greater  than 
100  per  cent;  but  since  such  gradients  do  not  apply  to  the  Murphree 
point  efficiency,  this  latter  efficiency  should  never  exceed  100  per  cent. 
The  theoretical  effect  of  the  concentration  gradient  has  been  studied 
by  a  number  of  investigators  (Refs.  15,  19,  21).  Three  cases  were  con- 
sidered by  W.  K.  Lewis,  Jr. :  Case  I,  vapor  completely  mixed,  liquid 
unmixed;  Case  II,  vapors  do  not  mix,  and  the  overflows  are  arranged 
such  that  the  liquid  flows  in  the  same  direction  on  all  plates;  Case  III, 
the  vapor  rises  from  plate  to  plate  without  mixing,  and  the  liquid  flows 
in  the  opposite  direction  on  successive  plates. 

Lewis  assumed  that  (1)  E^v  was  constant  over  all  of  the  plate,  (2) 
the  equilibrium  curve  is  a  straight  line  over  the  concentration  range 
involved,  ye  —  Kx  +  b,  and  (3)  the  liquid  flows  across  the  plate  with- 
out mixing. 

The  results  of  this  analysis  for  the  three  cases  are  given  in  Fig.  17-2 
in  which  the  ratio  (E°MV/E^V)  is  plotted  as  a  function  of  E^v  and  the 
ratio  of  the  slope  of  equilibrium  curve,  K,  to  the  slope  of  the  operating 
line,  (0/F).  The  slope  of  the  equilibrium  curve,  K,  should  be  the 
average  slope,  dye/dx,  over  the  concentration  region  involved.  These 
calculated  values  indicate  that  it  should  be  possible  to  obtain  high 
plate  efficiencies  by  preventing  the  liquid  from  mixing.  The  usual 
bubble-cap  plates  probably  fall  between  Cases  I  and  III  as  far  as  the 
vapor  is  concerned,  but  they  give  considerable  liquid  mixing  which 
would  lower  the  value  of  (E°UV/E^V)  as  compared  to  the  values  given 
by  the  plot.  Case  II  has  the  possibility  of  giving  higher  plate  effi- 
ciencies than  the  other  two  cases,  but  in  practice  it  is  difficult  to  arrange 
the  downflow  pipes  such  that  the  liquid  flows  in  the  same  direction  on 
all  plates.  The  circumferential  flow  plate,  Fig.  16-4$,  gives  essen- 
tially this  type  of  flow,  but  it  is  not  a  desirable  construction  in  most 
cases. 

The  relationship  between  the  over-all  column  efficiency  and  E°MY  can 
be  derived  in  a  similar  manner  and  assuming 

1.  Constant  0/7 


FRACTIONATING  COLUMN  PERFORMANCE 


449 


FIG.  17-2.     Relation  between  EMV° 


450 


FRACTIONAL  DISTILLATION 


2.  Constant  slope  of  the  equilibrium  curve,  i.e.,  dy,/dx  <=  K 

3.  E°ur  same  for  all  plates  considered. 
Lewie  (Ref.  21)  obtained 

pa_\nll+E°uvlK/(0/V)-l]} 
ln(K/(0/V)] 


(17-6) 


This  equation  is  plotted  in  Fig.  17-3.    It  will  be  noted  that  in  gen- 
eral E°/E0MV  is  close  to  1.0.     In  cases  where  the  rectifying  system 


0.3 


FIG.  17-3.     Relation  between  E°  and 


operates  from  low  to  high  concentrations,  the  value  of  K/(0/V)  will 

average  out  to  around  1.0,  making  E°  approximately  equal  to  E°uv. 

For  cases  where  K/(0/V)  is  widely  different  from  1.0  and  E°uv  is 

low,  the  ratio  of  E°  can  be  either  much  larger  or  smaller  than  Efo 


FRACTIONATING  COLUMN  PERFORMANCE 


451 


Thus  Fig.  17-4  illustrates  a  case  where  K/(0/V)  is  very  small,  and  one 
theoretical  plate  would  go  from  yn  to  a  composition  of  almost  1.0.  If 
the  over-all  column  efficiency  for  this  region  were  0.5,  two  actual 
plates  would  give  the  same  increase,  but  the  diagram  indicates  that 
about  five  plates  with  E°MV  ~  0.5  would  b6  required.  Actually  the 
efficiencies  employed  were  incompatible,  and  if  E°MY  =  0.5  then  E° 


FIG.  17-4. 

would  be  small  for  the  low  value  of  K/(0/V).  The  first  step  with 
E°MV  =  0.5  does  make  a  change  in  the  vapor  composition  equal  to 
one-half  of  that  for  a  theoretical  plate  but,  because  of  the  convergence 
of  the  equilibrium  curve  and  the  operating  line,  the  available  potential 
decreases  and  the  succeeding  plates  do  not  make  so  large  a  change  in 
the  vapor  composition. 

Murphree  also  gave  a  derivation  for  a  plate  efficiency  based  on 
liquid-phase  compositions.     The  basic  differential  equation  was 


-L'  dx  «  KUQ,(X  - 


d6 


(17-7) 


where  x  **  mol  fractiovn  i»  liquid 

x*  s*  liquid  in  equilibrium  with  Vapor  leaving 
Ki.  =  mass-transfer  coefficient 
V  «  liquid  in  slug  Under  consideration 
The  equation  was  integrated  with  Kt,  a,  L',  and  x*  constant  to  give 


E 


ML 


(17*8) 


452 


FRACTIONAL  DISTILLATION 


where  x09  Xi  «  liquid  composition  to  and  leaving  plate,  respectively 
EML  =  liquid-phase  plate  efficiency 
Z//0  =  liquid  rate 

It  is  difficult  to  picture  any  mass-transfer  process  on  a  bubble-type 
plate  that  corresponds  to  the  derivation  of  Eq.  (17-8).  For  example, 
consider  the  application  of  these  equations  to  a  local  section  of  the 
plate.  The  liquid  flowing  across  this  section  contacts  the  vapor  rising 
through  it,  but  the  vapor  composition  varies  with  the  liquid  depth 
while  the  integration  assumed  that  x*  was  constant.  The  mol  fraction 
ratio  of  Eq.  (17-8)  will  be  used  as  the  definition  of  EML,  but  the  mass- 
transfer  portion  of  the  equation  does  not  apply  to  bubble-cap  plate 
conditions. 

Equations  (17-3)  and  (17-8)  can  be  related  by  the  operating  line  and 
the  equilibrium  curve.  Assuming  that  the  equilibrium  curve  is  such 
that  Ke  =  y0/x*  =  y*/x0  (Ke  is  equal  to  K  if  the  equilibrium  curve  is 
a  straight  line  through  the  origin)  gives 


Ke          E°ML(l  - 


0/V      E°MV(l  ~ 


(17-9) 


The  assumptions  made  in  the  derivation  of  Eq.  (17-3)  appear  to  be 
on  a  somewhat  sounder  mass-transfer  basis  than  Eq.  (17-8).  If  this 
is  true,  E^v  depends  on  the  operating  conditions  only  to  the  extent 
that  they  affect  the  mass-transfer  conditions  while  EML  in  addition  is  a 
function  of  the  ratio  Ke/(0/V). 

TABLE  17-1 


Location  in  column 


Bottom 

Just  below 
feed  plate 

Just  above 
feed  plate 

Top 

o/v 

1.2 

1.2 

0.7 

0.7 

»K 

2.5 

1.0 

1.0 

0.4 

K/(om 

2.08 

0  883 

1.43 

0  57 

E°MV/E*aiV  " 

»  1.9 

1.38 

1.67 

1.18  (Fig.  17-2) 

EMV 

1  14 

0.83 

1.0 

0.71 

E°/E°MV 

0.96 

0.98 

1.0 

0  91  (Fig.  17-3) 

E° 

1  09 

0.81 

1.0 

0.65 

K. 

2.5 

1.5 

1.5 

1.0 

EML 

1.06 

0.86 

1.0 

0.78 

The  vapor  efficiency  E°MV  is  much  more  commonly  used  than  E°ML. 
This  preference  is  justified  in  view  of  the  derivation  of  the  latter. 


FRACTIONATING  COLUMN  PERFORMANCE 


453 


As  an  illustration  of  these  relationships,  consider  the  separation  of  a 
benzene-toluene  mixture  with  (0/F)n  =  0.7  and  (0/F)m  =  1.2  for 
Eft?  =  0.6.  The  flow  conditions  correspond  to  Case  III. 

The  feed-plate  region  was  arbitrarily  chosen  to  make  K  =  1.0. 
This  corresponds  to  a  liquid  composition  of  about  40  per  cent  benzene. 

The  values  given  for  E°MV  would  be  higher  than  actually  obtained 
because  of  the  liquid  mixing  on  the  plate.  It  is  interesting  to  note  that 
E°/E°MV  is  near  to  1.0  at  all  positions,  and  the  use  of  E°  =  E°MV  would 
be  a  reasonable  approximation.  The  values  of  Ke  are  larger  than  K 
except  at  the  bottom  of  the  column.  For  usual  cases,  Ke/(0/V)  is 
greater  than  1.0,  and  E°ML  will  be  between  E°MV  and  1.0. 


FIG.  17-5.     Application  of  Murphree  plate  efficiencies. 

From  the  calculational  viewpoint,  E°  is  the  easiest  to  apply,  but 
E°MV  and  E°ML  can  be  used  without  much  additional  effort  either  in 
algebraic  or  graphical  design  calculations.  Thus  in  algebraic  calcula- 
tions, the  usual  theoretical  plate  calculations  will  give  y*  for  a  known 
2/n_i,  and  yn  =  t/n-i  +  E°MV(y%  —  2/n_i).  Thus  the  values  of  yn  can  be 
calculated,  if  E°MV  is  known  and  the  calculation  is  repeated  for  the  next 
plate.  The  efficiencies  can  also  be  applied  to  the  graphical  calculation 
for  binary  or  multicomponent  mixtures.  This  is  illustrated  in  Fig. 
17-5.  The  use  of  E°MV  is  illustrated  on  the  upper  operating  line :  yn-i  is 
the  actual  vapor  composition  entering  plate  n  and  by  material  balance 
the  composition  xn  is  fixed  on  the  operating  line;  the  vapor  in  equilib- 


454 


FRACTIONAL  DISTILLATION 


rium  with  xn  is  y*  and  the  actual  increase  in  vapor  composition  is 
I/ft  —  y^i,  which  is  obtained  by  a  vertical  step  of  a  height  equal  to 
E*Mv(y%  -~  #n~i);  i.e.,  a  fraction  of  the  theoretical  plate  increase  equal 
to  El[V  is  taken.  The  use  of  E°M L  for  plate  m  is  shown  on  the  lower 
operating  line.  In  this  case  a  fractional  horizontal  step  equal  to  E°ML 
times  the  theoretical  plate  change  is  used. 

Effect  of  Entrainment  on  Efficiency.  Entrainment  can  lower  the 
apparent  plate  efficiency  because  the  vapor-liquid  mixture  carried  to 
the  plate  above  will  have  a  lower  average  concentration  of  the  more 
volatile  components  than  the  vapor  alone.  Colburn  (Ref .  6)  has  given 
an  equation  relating  the  measured  efficiency  with  entrainment  to  that 
for  a  plate  giving  the  same  change  in  vapor  composition  but  without 
entrainment. 


Jpo      

&UT    — 


E 


1  + 


eE 

oTv 


(17-10) 


where  E°MT 
E 


apparent  efficiency  with  entrainment 
efficiency  for  same  change  in  vapor  composition 
e  =  mols  of  entrained  liquid  per  mol  of  vapor 
For  most  cases  the  effect  of  entrainment  does  not  become  serious 
until  e  is  0.1  or  greater.    Referring  to  Fig.  16-10,  it  will  be  noted  that 
e  «  0.1  corresponds  to  VcpQ*  of  about  0.9  or  for  atmospheric  pressure 
a  velocity  of  2  to  4  f  .p.s.  with  12-in.  plate  spacing  and  to  several  fold 
higher  velocities  for  24-in.  spacing.    Equation  (17-10)  is  normally  used 
with  E  taken  as  the  expected  value  of  E°MVJ  but  this  is  in  error  due  to 

TABLE  17-2 


Bottom 

Just  below 
feed  plate 

Just  above 
feed  plate 

Top 

*fcr 

1.04 

0.77 

0.87 

*v 

the  mass  transfer  that  takes  place  between  the  vapor  and  liquid  drop- 
lets above  the  main  liquid  body  on  the  plate.  The  composition  of  the 
liquid  employed  was  that  of  overflow  to  the  plate  below  while  the 
actual  composition  will  be  different  because  of  mass  transfefc^nd  con- 
centration gradients.  This  equation  will  give  a  too  high  value  for  E^ T 
when  Eltr  is  used  f  or  E.  For  a  value  of  e  »  0.1  and  using  E°ur  *  E; 
these  values  of  E«MT  corresponding  to  Table  17-1  are  given  in  Table  17-2. 
In  this  case  the  loss  in  efficiency  for  e  «  0.1  is  relatively  small,  and 


FRACTIONATING  COLUMN  PERFORMANCE 


455 


larger  values  of  e  are  not  commonly  found  in  commercial  practice 
because  the  operation  of  the  column  becomes  unstable  before  the  cor- 
responding velocities  are  obtained. 

Experimental  Data  on  Plate  Efficiencies.  A  considerable  number 
of  investigations  of  plate  efficiencies  have  been  reported,  but  it  is  diffi- 
cult to  obtain  a  coherent  picture  from  the  data  because  of  the  number 
of  unknown  factors  usually  involved.  Thus  some  of  the  results  are 
reported  on  an  over-all  column  basis  while  others  are  given  on  a  plate 
basis.  They  involve  unknown  amounts  of  entrainment,  unknown 
amounts  of  liquid  mixing  in  the  plate,  unknown  hydraulic  gradients, 
unknown  vapor  distribution,  unknown  interface  temperatures,  and  in 
some  cases  unknown  degrees  of  liquid  by-passing  or  dumping.  As  a 
result,  most  of  these  data  are  not  suitable  for  correlation  purposes  but 
are  useful  for  giving  a  general  picture  of  the  results  obtained.  The 
following  discussion  reviews  some  of  these  data  for  orientation  pur- 
poses, but  it  is  not  intended  to  be  a  comprehensive  survey. 

TABLE  17-3 


System 

Average  Murphree  plate  efficiency, 

E°MV 

Average  superficial  vapor  velocity,  f.p.s  .  .    . 

I 

2 

3 

4 

5 

Methanol-water.  .  . 

99 

96 

90 

82 

73 

n-Propanol-water  . 

83 

85 

88 

88 

80 

Isobutanol-water  . 

98 

95 

90 

84 

75 

3V&8thanol-n-propanol  . 

90 

88 

87 

87 

87 

Methanol-isobutanol  

75 

71 

75 

76 

73 

Benzene-carbon  tetrachloride 

82 

88 

89 

84 

74 

Gadwa  (Ref.  12)  has  studied  the  plate  efficiency  in  the  fractionation 
of  mixtures  of  (1)  benzene-carbon  tetrachloride,  (2)  methanol-iso- 
butanol,  (3)  methanol-ra-propanol,  (4)  isobutanol-water,  (5)  n-pro- 
panol-water,  and  (6)  inethanol-water.  A  small  four-plate  column 
containing  one  bubble  cap  per  plate  was  employed.  The  bubble  caps 
were  3J^  in.  in  diameter  and  2  in.  high  containing  38  slots  %  in.  wide 
by  %  in.  high  per  cap.  A  vapor  space  of  5  by  5  in.  was  partitioned  off 
from  the  overflow  pipes,  giving  a  ratio  of  slot  area  to  superficial  area 
of  0.12.  The  plates  were  spaced  11  in.  apart,  and  overflow  weirs  were 
employed.  Plate  samples  were  taken  so  that  the  Murphree  plate  effi- 
ciencies could  be  Calculated.  Some  of  these  results  are  given  in  Table 
17-3.  The  efficiencies  i&  this  table  were  calculated  for  the  vapor  phase. 


456  FRACTIONAL  DISTILLATION 

Gadwa  concluded  that,  for  the  mixtures  he  studied,  the  Murphree 
plate  efficiency  was  substantially  independent  of  the  concentration  and 
of  the  vapor  velocity  so  long  as  foaming  and  entrainment  were  not 
appreciable  but  that,  when  foaming  and  entrainment  did  occur,  the 
efficiency  decreased  with  increasing  velocity. 

Brown  et  al.  (Ref .  4)  and  Gunness  (Ref .  14)  both  report  Murphree 
plate  efficiencies  of  100  per  cent  or  greater  for  large  commercial  gasoline 
stabilizers.  The  tower  studied  by  Gunness,  operated  at  250  p.s.i.g., 
was  4  ft.  8^i  in.  in  diameter,  and  contained  28  plates  each  having  27 
cast-iron  bubble  caps.  The  bubble  caps  were  6^[  in.  in  diameter  and 
contained  32  1-  by  -Hrin.  rectangular  slots  per  cap.  The  plate  spacing 
was  18  in.  In  these  columns,  there  were  a  number  of  bubble  caps  per 
plate,  and  the  liquid  flowed  in  opposite  directions  on  successive  plates. 
Gunness  analyzed  his  data  by  Lewis's  cross-flow  enrichment  method 
(page  448)  and  concluded  that  the  Murphree  point  efficiency  was 
between  70  and  80  per  cent. 

Lewis  and  Smoley  (Ref.  22)  studied  the  plate  efficiency  in  the  rectifi- 
cation of  mixtures  of  (1)  benzene-toluene,  (2)  benzene-toluene-xylene, 
and  (3)  naphtha  and  mixtures  of  pinene  and  aniline  in  naphtha.  An 
experimental  column  8  in.  in  diameter  with  10  plates  spaced  16  in. 
apart  was  used.  The  bubble  cap  was  rectangular,  being  2  in.  high  and 
2  in.  wide,  and  extended  across  the  column.  There  were  24  slots  % 
by  ^{6  in.  on  each  side  of  the  cap,  giving  a  ratio  of  slot  area  to  super- 
ficial area  of  about  0.16.  The  investigators  found  average  plate  effi- 
ciencies of  60  per  cent  for  the  benzene-toluene  mixture,  75  per  cent  for 
the  ternary  mixture,  and  80  to  95  per  cent  for  the  naphtha  mixtures. 

In  the  same  tower,  Carey,  Griswold,  Lewis,  and  McAdams  (Ref.  5) 
found  an  average  Murphree  efficiency  of  70  per  cent  when  fractionating 
benzene-toluene.  They  found  the  efficiency  substantially  constant  for 
superficial  velocities  from  0.2  to  4.5  f.p.s.  and  independent  of  liquid 
composition.  The  same  investigators  report  efficiencies  of  50  to 
99.75  per  cent  for  the  fractionation  of  an  ethanol-water  mixture  in  a 
6-in.-diameter  tower  containing  one  plate.  The  logarithm  of  100 
minus  the  plate  efficiency  was  found  to  be  a  linear  function  of  the 
depth  of  submergence  of  the  slots.  A  benzene-toluene  mixture  in  the 
same  one-plate  tower  gave  an  average  Murphree  efficiency  of  58  per 
cent.  A  distillation  of  an  aniline-water  mixture  in  the  10-plate  tower 
gave  an  average  plate  efficiency  of  58  per  cent  at  a  vapor  velocity  of 
2.77  f.p.s. 

Lewis  and  Wilde  (Ref.  23)  found  an  average  plate  efficiency  of  65  per 
cent  at  a  vapor  velocity  of  2.8  f.p.s.  for  the  rectification  of  naphtha  in  a 


FRACTIONATING  COLUMN  PERFORMANCE  457 

10-plate  column  9  ft.  in  diameter.  There  were  115  bubble  caps  per 
plate  containing  slots  Y±  by  1  in.  The  ratio  of  slot  area  to  superficial 
area  was  0.10,  and  the  plate  spacing  was  2  ft. 

Brown  (Ref.  3)  reports  efficiencies  as  high  as  120  per  cent  for  a  com- 
mercial beer  column  using  perforated  plates.  The  same  efficiency  was 
reported  for  the  rectification  of  an  ethanol-water  mixture  in  a  special 
laboratory  column.  The  same  investigator  reports  efficiencies  of  about 
20  per  cent  for  naphtha-absorption  towers. 

Atkins  and  Franklin  (Ref.  1)  found  an  over-all  column  efficiency  of 
18  per  cent  for  a  natural  gasoline  absorber  using  gas  oil  as  the  absorbing 
liquid.  Walter  (Ref.  30)  obtained  Murphree  vapor  plate  efficiencies 
from  80  to  95  per  cent  in  a  2-in.  laboratory  column  for  air  humidifica- 
tion.  Data  taken  in  the  same  unit  on  the  absorption  of  propylene 
and  isobutylene  in  gas  oil,  heavy  naphtha,  and  mixtures  of  gas  and  lube 
oil,  gave  plate  efficiencies  on  the  vapor  basis  of  5  to  36  per  cent. 

Horton  (Ref.  16)  studied  the  absorption  of  carbon  dioxide  and 
ammonia  in  water  on  a  single  18-in. -diameter  plate  and  reported  values 
of  E°MV  about  3  per  cent  for  CC>2,  and  70  per  cent  for  ammonia.  Fair- 
brother  (Ref.  9)  studied  the  absorption  of  carbon  dioxide  in  aqueous 
solutions  of  glycerine  and  obtained  values  of  the  Murphree  plate  effi- 
ciency of  0.65  to  4  per  cent. 

Peavy  and  Baker  (Ref.  27)  investigated  the  rectification  of  ethanol- 
water  mixtures  in  an  18-in.-diameter  column,  and  obtained  plate  vapor 
efficiency  from  80  to  120  per  cent  for  superficial  vapor  velocities 
between  1  and  3  f.p.s.  For  other  data  on  plate  efficiency,  see  Refs.  7, 
8,  13,  18,  25,  26,  28. 

The  values  of  the  plate  vapor  efficiency  vary  from  less  than  1  to  over 
100  per  cent.  The  absorption  systems  with  gases  of  low  solubility  and 
liquids  of  high  viscosity  have  low  efficiencies;  while  most  of  the  distilla- 
tion systems  give  values  of  60  to  100  per  cent.  In  most  cases  the 
efficiency  is  relatively  independent  of  vapor  rate  until  appreciable 
foaming  and  entrainment  are  encountered.  For  the  distillation  sys- 
tems there  does  not  appear  to  be  any  significant  effect  of  liquid 
composition. 

Plate  Efficiency  Correlations.  It  has  been  pointed  out  that  the 
correlation  of  the  plate  efficiency  data  is  difficult  because  of  the  large 
number  of  unknown  conditions  involved  in  most  cases.  However, 
some  correlations  have  been  developed  that  are  helpful  in  estimating 
the  plate  efficiency. 

Gunness  (Ref.  14)  analyzed  the  data  for  several  columns  and  con- 
cluded that  the  liquid-film  resistance  was  a  major  factor,  and  he  sug- 


458 


FRACTIONAL  DISTILLATION 


gfcsted  that  the  plate  efficiency  be  correlated  with  the  viscosity  of  the 
liquid,  because  this  characteristic  is  a  major  factor  in  liquid-phase  mass 
transfer.  He  plotted  the  efficiency  as  a  function  of  the  operating  pres- 
sure, on  the  basis  of  the  fact  that  the  viscosities  of  liquids  are  approxi- 
mately the  same  at  a  given  vapor  pressure.  This  correlation  on  a  vis- 
cosity basis  is  given  in  Fig.  17-6. 


to 

6 
6 

4 
2 

1.0 
08 

_  

yn  

1  {  1  II  —  le^u  — 

-4  —  — 

11—  

u 

A  -Watte 

r  and  Sherwood         Q 
r  and  Sherwood     /Q(ft 

S,  I 

J\ 

ft  .  U/ftfJ-a 

/i 

N  \ 

/*  .  fir/nna 

V 

v 

Ss  >L 

D  ~  Or/earner  and  Bradford     •  — 

f 

^ 

^ 

X 

s, 

<' 

E  '  O'Connell  (absorber)         10,000 
F'Q'ConrnN  (absorber)            0 
G  'Q'Cortne/tffrQcffonofor)     ft  *2.Q  — 

0.6 

UJ 

1    0.2 
0.! 

i  — 

s 

N 

--. 

•'Sb_ 

•«• 

s^ 

^ 

>& 

"^ 

•*•-. 

*>, 

^( 

^ 

x 

IN- 

K 

X 

">s 

v^ 

s 

s      X. 

X 

^ 

^ 

s-^^ 

^ 

^x 

X 

s 

\ 

\ 

X 

c 

X 

008 
0.06 

).04 
101 

s 

\-\ 

D. 

_^_ 

Jisi^ 

s^ 

V1  "' 

s^ 

s 

X 

s 

x  , 

s 

N 

sf 

s 

\ 

x 

\ 

s 

\ 

s 

V 

\ 

0.01     Q02      Q04Q06    01       0.2       04  060810        2        4     6  8  10       20 

Liquid  viscosiiy-centipoises 
FIG.  17-6.     Comparison  of  plate  efficiency  correlations. 


40  60 


Walter  and  Sherwood  (Ref.  31)  gave  a  correlation  for  the  plate  effi- 
ciency based  on  the  derivation  of  Eq.  (17-3).  Using  the  two-film 
absorption  concept  (Ref.  33),  they  separated  the  over-all  mass-trans- 
fer resistance  into  a  liquid-  and  vapor-film  resistance. 


wiiere  KQ&  «  same  as  for  Eq.  (17-1) 

koa  *  gas-film  transfer  coefficient 
kid  »  liquid-filin  transfer  coefficient 
H  «  Henry's  law  constant 


(17-11. 


FRACTIONATING  COLUMN  PERFORMANCE  459 

The  basic  assumptions  of  the  two-film  theory  are  not  satisfied  by  the 
action  of  a  bubble-cap  plate,  but  Eq.  (17-11)  is  probably  a  reasonable 
approximation  for  the  division  of  the  resistance  between  the  two 
phases,  indicating  that  solubility  is  a  major  factor  in  the  relative 
resistances.  * 

Walter  and  Sherwood  assumed  (1)  that  KG  was  proportional  to 
G/0,  (2)  that  the  total  interfacial  area  of  all  the  bubbles,  a,  was  pro- 
portional to  the  liquid  depth  from  the  center  of  the  slots  to  the  top  of 
the  overflow  weir,  (3)  on  the  basis  of  the  data  of  Carey,  Griswold, 
Lewis,  and  Me  Adams  (Ref.  5),  that  KQCL  was  proportional  to  the  cube 
root  of  the  slot  width,  and  (4)  that  both  k0a  and  kLa  were  proportional 
to  the  0.68  power  of  the  liquid  viscosity.  Their  equation  for  the  value 
of  m  in  Eq.  (17-3)  is 


HIS) 


„(). 68,^0. 33 


where  H  =  Henry's  law  constant,  Ib.  mols  per  cu.  ft.  per  atm. 
P  =  pressure,  atm. 
p  =  viscosity  of  liquid,  centipoises 
w  =  slot  width,  in. 

h  =  height  from  center  of  slots  to  top  of  weir,  in, 
El v  -  1  -  er* 

Equation  (17-12)  can  be  made  more  suitable  for  distillation  calcula- 
tions by  replacing  Henry's  law  constant  by  the  equilibrium  constant 
Ko,  («  y/x),  giving 


where  Ke  =  equilibrium  constant  =  yjx 

M  =  molecular  weight  of  the  liquid 
d  *  specific  gravity  of  liquid  relative  to  water 
tie  terms  2.5  and  O.Q0591£«Af /d  are  the  relative  resistances  of  the 
vajzfor  tod  liquid  phases,  respectively. 

/Ke  and  M  are  properties  of  the  system  and,  for  a  given  mixture,  it 
is  difficult  to  make  any  major  changes.  In  distillation,  the  values 
of  Ke  for  the  key  components  usually  range  from  1.0  to  a  maximum 
of  approximately  5,  but  for  the  absorption  of  relatively  insoluble 
gases,  Kg  may  be  as  large  as  1,000.  The  latter  systems  give  small 
Values  of  m  and  low  EMV.  By  increasing  the  total  pressure,  the  values 


460  FRACTIONAL  DISTILLATION 

of  K e  for  the  absorption  cases  can  be  reduced,  but  the  values  for 
distillation  cases  remain  about  the  same. 

The  viscosity  term  plays  a  major  part  in  determining  the  value  of 
w,  and  it  is  subject  to  some  control.  In  the  case  of  distillation  systems, 
the  liquids  are  essentially  at  their  boiling  points  under  the  pressure 
involved.  Under  such  conditions  most  common  liquids  have  viscosi- 
ties of  the  same  order  of  magnitude.  However,  raising  the  operating 
pressure  increases  the  temperature  and  lowers  the  viscosity.  Thus, 
high-pressure  towers  tend  to  give  high  plate  efficiencies,  but  the  gain  is 
usually  not  great  enough  to  justify  such  operation  for  this  purpose 
only.  In  the  case  of  absorption  towers,  the  operating  temperature  can 
often  be  varied  independently  of  the  pressure,  and  at  a  given  operating 
pressure  a  high  temperature  gives  a  larger  value  of  Ke  and  a  lower 
value  of  ju.  These  two  counterbalancing  effects  usually  work  to  pro- 
duce an  optimum  operating  temperature. 

In  most  distillation  systems,  the  value  of  the  viscosity  is  from  0.15  to 
0.5  centipoise,  and  with  slots  J^$  to  J^  in.  wide  they  give  point  effi- 
ciencies of  60  to  95  per  cent.  However,  in  the  case  of  absorption,  the 
liquid  viscosities  may  be  as  high  as  20  centipoises,  and  the  point  effi- 
ciencies may  be  10  per  cent  or  lower. 

For  a  binary  mixture  it  can  be  shown  that  the  value  of  EMV  must  be 
the  same  for  both  components  in  order  to  make  the  mol  fractions  in 
the  liquid  and  the  vapor  add  up  to  unity.  In  the  case  of  multicom- 
ponent  mixtures,  each  of  the  components  can  have  a  different  value  of 
EMV  for  a  given  plate.  On  the  basis  of  Eq.  (17-13)  it  might  be  assumed 
that  the  heavier  components  with  lower  values  of  Ke  would  have  the 
highest  efficiency.  There  are  no  data  which  prove  conclusively  the 
relative  efficiencies  in  a  multicomponent  mixture  on  a  given  plate. 
The  inaccuracies  of  the  measurements  are  such  that  they  leave  the 
trend  in  doubt.  However,  they  do  indicate  that  for  the  mixtures  so 
far  tested  the  difference  in  the  value  of  E°MV  between  components  is  not 
large.  On  the  basis  of  diffusion  theory,  it  would  be  expected  that  the 
h^atjlfef;ij^mponents  would  approach  equilibrium  more  slowly  and 
therefore  have  lower  values  of  EMV. 

Tfe^&lues  calculated  from  these  equations  are  given  in  Fig.  17-6, 
for  h  «  l.o,  w  =  0.25,  and  various  values  of  KM/d. 

Drickamer  and  Bradford  (Ref .  7)  analyzed  the  test  data  for  a  num- 
ber of  rectifying  towers  separating  hydrocarbons  and  gave  a  plot  of 
E*  as  a  function  of  the  molal  average  liquid  viscosity.  The  data  pre- 
sented by  these  investigators  correlated  reasonably  well  over  a  con- 
siderable variation  in  liquid  viscosity.  They  did  not  include  systems 


FRACTIONATING  COLUMN  PERFORMANCE  461 

that  would  have  large  values  of  Ke  =  y/x,  and  the  correlation  would 
probably  not  be  suitable  for  such  cases.  Their  relation  is  given  in  Fig. 
17-6. 

A  correlation  similar  to  that  of  Drickamer  and  Bradford  was  devel- 
oped by  O'Connell  (Ref.  26),  but  in  this  caste  the  plate  efficiency  was 
plotted  as  a  function  of  ap  (a  =  relative  volatility)  for  fractionating 
towers  and  of  p,/HP  or  K6M^/d  for  absorption  towers.  These  correla- 
tions include  both  of  the  main  factors,  solubility  and  liquid  viscosity, 
found  to  be  important  by  Walter  and  Sherwood.  The  inclusion  of  the 
solubility  factor  directly  with  the  viscosity  is  probably  not  so  sound  as 
the  type  of  grouping  used  in  Eq.  (17-13).  It  would  be  expected  that 
O'ConnelFs  correlation  would  break  down  for  extreme  or  unusual 
variations  in  the  solubility  factor. 

Geddes  (Ref.  13)  has  presented  a  semitheoretical  method  for  the 
estimation  of  plate  efficiency.  The  method  is  complicated  to  apply, 
and  several  of  the  assumptions  made  in  the  derivation  are  question- 
able. He  obtained  good  results  by  the  method  in  cases  involving 
widely  differing  conditions.  For  the  present  state  of  the  art,  it  is 
believed  that  the  Walter  and  Sherwood  equation  gives  as  satisfactory 
results  and  is  easier  to  apply. 

A  more  detailed  analysis  of  the  mechanism  of  the  mass  transfer 
between  the  two  phases  of  a  bubble-cap  plate  has  been  made  by 
Etherington  (Ref.  8).  He  presented  correlations  for  both  kGa  and  kLa. 
These  values  would  then  be  combined  with  Eq.  (17-11)  to  give  KGa 
which  would  be  used  to  calculate  E^v.  The  method  has  been  tested 
on  only  a  few  mixtures,  and  a  more  detailed  evaluation  is  needed  to 
determine  whether  the  added  complications  are  justified. 

The  correlations  proposed  by  Gunness,  Walter  and  Sherwood, 
Drickamer  and  Bradford,  and  O'Connell  are  compared  in  Fig.  17-6. 
The  ordinate  is  —  In  (1  —  E),  and  the  abscissa  is  the  viscosity  of  the 
liquid  in  centipoises.  The  correlations  are  not  all  comparable. 
Walter  and  Sherwood's  relation  was  based  on  small  laboratory  units 
and  probably  corresponds  to  point  conditions,  while  the  other  rela- 
tions were  based  on  plate  or  over-all  column  efficiencies.  For  Eq. 
(17-13)  h  was  taken  =  1.0,  w  =  0.25,  and  two  curves  are  given  with 
KeM/d  =  0  being  used  to  approximate  fractionating  conditions  and 
KM  Id  =  10,000  to  correspond  to  absorber  conditions.  Three  curves 
are  given  for  O'ConnelPs  relations;  curves  E  and  F  are  for  the  absorber 
correlation  with  KeM/d  =  0  and  10,000,  respectively;  and  curve 
G  is  for  the  fractionator  correlation  with  a  =  2.0.  The  curves  for 
Gunness  and  Drickamer  and  Bradford  agree  with  the  K6M/d  =  0 


462  FRACTIONAL  DISTILLATION 

curve  of  Walter  and  Sherwood  at  low  viscosities  and  approach  the 
KM  I A  **  10,000  curve  at  high  viscosities.  This  is  because  the  data 
used  in  these  correlations  for  the  high  efficiencies  were  for  fractionating 
columns  having  low  values  of  liquid  viscosities  and  of  KeM/d,  while 
the  low  efficiency  points  were  for  absorbers  with  high  values  of  liquid 
viscosity  and  KeM/d.  While  in  many  cases  the  solubility  and  vis- 
cosity factors  tend  to  parallel  each  other,  it  is  possible  to  vary  the 
viscosity  widely  for  the  same  value  of  the  solubility.  Etherington's 
data  indicate  that  the  viscosity  and  solubility  factors  should  be  sepa- 
rated. The  curves  based  on  O'ConnelPs  correlation  for  absorbers 
with  KeM/d  =  10,000  are  in  reasonable  agreement  with  curve  B,  but 
the  values  for  high  solubilities  (KeM/d  near  0)  do  not  appear  to  repre- 
sent the  data.  At  higher  values  of  KeM/dj  the  curves  agree  approxi- 
mately with  those  of  Walter  and  Sherwood. 

From  a  review  of  the  available  data,  it  is  recommended  that  Walter 
and  Sherwood's  equation  be  used  for  values  of  E^v.  In  using  this 
equation,  it  is  suggested  that  for  fractionating  towers  the  relative  vola- 
tility of  the  light  key  to  heavy  key  components  be  used  instead  of  Ke  to 
calculate  the  Murphree  point  vapor  efficiency  for  the  light  key  com- 
ponent and  that  this  value  of  the  efficiency  be  used  for  all  components 
in  the  mixture.  The  experimental  data  for  fractionation  systems  do 
not  vary  with  compositions  to  the  extent  that  would  be  indicated  by 
the  use  of  Ke  values.  The  viscosity  term  can  be  taken  as  molal  aver- 
age viscosity  of  the  liquid.  In  the  case  of  absorbers,  it  is  suggested 
that  K  =  dy/dx  be  used  instead  of  Ke  =  y/x,  since  the  former  is  more 
consistent  with  the  derivation  of  Eq.  (17-11).  The  efficiency  should 
be  calculated  for  the  key  component,  and  it  is  suggested  that  the  same 
value  be  employed  for  the  other  components.  The  value  of  h  should 
be  a  function  of  tfye  liquid  depth  instead  of  the  "  fixed  "  value  given  for 
the  correlation.  It  is  suggested  that  this  term  be  replaced  by  (&')°'5> 
where  hf  is  the  distance  from  the  center  of  the  slot  opening  to  the  top 
of  the  liquid  level  over  the  weir.  With  these  changes,  the  relations 
become 


where  m'  -  -j ^ r (17-14) 

'        ,  0.00594OA    068   083 

2,5  .] 1  ^0.68^0.83 

h'  =  distance  from  top  of  liquid  level  at  weir  to  center  of  slot 

opening 
K  «•  use  relative  volatility  of  key  components  for  fractionating 


FRACTIONATING  COLUMN  PERFORMANCE  463 

columns,   and  K  =  (dy/dx)   for  equilibrium  curve  for 
absorbers 

jj  =  viscosity,  centipoises 
w  =  slot  width,  in. 

For  values  of  the  Murphree  plate  vapor*  efficiency  in  commercial 
size  towers,  the  following  equation  is  recommended: 


- 

where  m"  =  7  -     *\  -  (17-16) 


7  \ 

/0  K   ,  0.005JOf  \    0  68   o  33 

/  35   _|  --  -  1  ^0.88^0.33 
\  d  / 


The  units  are  the  same  as  those  of  Eq.  (17-14). 

Plate  Efficiency  Example.  As  an  example  of  the  use  of  these  equations,  the 
plate  efficiency  for  the  plate  design  example  of  Chap.  16,  page  433  will  be  calcu- 
lated. The  values  of  the  various  terms  from  this  previous  example  are  summarized 
below: 

K  «  relative  volatility  «  2.4 
M  -92 

d  -  52.8/62.4  -  0.86 
w  «*  M  in. 
0/V  -  1.2 

At  6WfR.  viscosity  of  toluene  «  0.24  centipoise. 
The  liquid  depth  at  the  outlet  weir  is  hw  -f-  hcr  —  2.0  +  1.0  —  3.0  in. 
The  top  of  the  slots  is  1.75  in.  above  the  plate  and  the  slot  opening,  h9  «*  1.13, 

1  13 
giving  the  center  of  the  slot  opening  at  1.75  --  ^—'  »  1.2  in.  above  the  plate. 

The  value  of  V  is  3,0  -  1.2  -  1.8  in. 
By  Eq.  (17-14). 

(1.8)0'5 


-  1.75 

E^y     -     1     - 

=  0.83 
By  Eq.  (17-16), 


m' 


,// d-8)" 


1.47 


«-  1.02 


464  FRACTIONAL  DISTILLATION 

EFFICIENCY  OF  PACKED  TOWERS 

The  efficiency  of  packed  towers  is  generally  expressed  as  the  height 
equivalent  to  a  theoretical  plate.  Most  of  the  reported  values  of 
H.E.T.P.'s  are  for  small  laboratory  columns,  since  this  is  one  of  the 
largest  uses  of  packed  columns.  H.E.T.P.  is  a  function  of  the  packing 
dimension  and  construction,  tower  size,  vapor  velocity,  and  system 
being  rectified.  The  efficiency  of  packed  towers  may  be  seriously 
impaired  by  the  liquid's  tending  to  pass  down  one  side  while  the  vapor 
flows  up  the  other.  This  channeling  of  vapor  and  liquid  prevents  effec- 
tive interaction  between  the  vapor  and  liquid. 

Baker,  Chilton,  and  Vernon  (Ref .  2)  report  the  results  of  tests  on  the 
distribution  of  water  over  various  packing  materials  with  air  flowing  up 
through  the  packing.  The  water  rate  was  500  Ib.  per  hr.  per  sq.  ft.  in 
all  tests.  They  found  that  a  ratio  of  tower  diameter  to  packing  size 
greater  than  8  to  1  gave  a  fairly  uniform  liquid  distribution.  At  values 
of  the  ratio  less  than  8,  the  liquid  tended  to  run  down  the  tower  walls 
and  leave  the  center  of  the  column  nearly  dry.  A  multiple-point 
liquid  distributor  at  the  top  improved  the  liquid  distribution  at  the  top 
portion  of  the  tower.  The  results  of  these  investigators  indicate  the 
desirability  of  having  the  tower  diameter  over  eight  times  the  size  of 
the  packing  material  and  of  using  multiple-point  liquid  distributors; 
this  latter  is  most  important  in  short  towers. 

Fenske,  Tongberg,  and  Quiggle  (Ref.  11)  give  the  results  of  a  large 
number  of  tests  on  packed  laboratory  towers.  A  comparison  of  some 
of  their  results  for  the  distillation  of  a  carbon  tetrachloride-benzene 
mixture  is  given  in  Table  17-4. 

The  investigators  conclude  that  (1)  the  best  packings  are  one-turn 
and  two-turn  wire  or  glass  helices,  carding  teeth,  and  No.  19  jack 
chain,  (2)  the  efficiency  of  the  packing  decreases  when  the  tower 
diameter  is  increased  or  when  the  height  of  the  packed  section  is 
increased,  and  (3)  different  hydrocarbon  mixtures  give  approximately 
the  same  value  for  H.E.T.P. 

Weimann  (Ref.  32)  has  published  results  on  the  fractionation  of 
ethanol-water  mixtures  in  packed  towers.  Using  a  superficial  velocity 
of  1  f.p.S.  at  0/D  =  1.0  with  8-  by  8-mm.  porcelain  Raschig  rings  in  a 
0.11-m.  (43^-in.)  diameter  tower,  H.E.T.P.  values  of  6  to  8.5  in.  were 
obtained  for  packing  heights  of  3.5  to  13  ft.  A  larger  tower,  approxi- 
mately 1  ft.  in  diameter,  using  a  7-ft.  depth  of  the  same  packing  gave 
an  H.E.T.P.  of  8.5  in.  at  a  superficial  vapor  velocity  of  \Y±  f.p.s. 

Jantzen  (Ref.   17)  has  presented  the  results  of  fractionating  an 


FRACTIONATING  COLUMN  PERFORMANCE 


465 


ethanol-water  mixture  in  a  13.5-cm.  (1.37-in.)  tower  packed  to  a  depth 
of  1  m.  with  either  1-  or  0.46-cm.  Raschig  rings.  The  values  of 
H.E.T.P.  calculated  from  his  data  range  from  about  3  to  6  in.,  for 
superficial  vapor  velocities  ranging  from  0.15  to  2  f  .p.s.  The  H.E.T.P. 
values  increased  as  the  0.2  power  of  the  vapor  velocity  and  were  about 
50  per  cent  larger  for  the  large  than  for  the  small  rings.  These  values 
were  found  to  be  independent  of  the  liquid  concentration  and  of  the 
reflux  ratio  (experimental  values  of  0/D  ranged  from  4  to  10). 

TABLE  17-4 


Packing 

Tower  dimensions, 
in. 

H.E.T.P., 
in. 

Straight  %2-i11*  carding  teeth.  . 
Straight  JHj2-in.  carding  teeth. 
Bent  H-in«  carding  teeth  ... 
Miscellaneous  carding  teeth                            .    .  . 
Double-cross  wire  form  .... 
Hollow-square  wire  form 
No.  20  single-link  iron  jack  chain 
No.  2  cut  tacks      ....                             ... 

0  76  by  27 
0  76  by  27 
0.76  by  27 
0  76  by  27 
0.76  by  27 
0.80  by  55 
2.0  by  53 
0  8  by  55 

1.5 
1.7 
1.7 
2.2 
2.1 
5.4 
5.2 
2.4 

6-turn  No.  24  Lucero  wire  helix 
No.  18  single-link  iron  jack  chain.  . 
Glass  tubes  
No.  16  single-link  iron  jack  chain.  ...          ... 

0.8  by  66 
2  0  by  53 
0  78  by  27 
0.76  by  27 

8.0 
6.5 
5  5 
4.2 

Fenske  and  coworkers  (Ref .  10)  have  also  made  an  extensive  study 
of  the  efficiency  of  packings  when  used  for  the  separation  of  a  n-hep- 
tane-methyl  cyclohexane  mixture  in  a  2-in,-diameter  glass  tower  at 
total  reflux.  The  tower  was  114  in.  high  and  was  operated  at  atmos- 
pheric pressure.  A  few  of  their  results  are  summarized  in  Table  17-5. 

These  data  indicate  that  H.E.T.P.  values  as  low  as  1.5  in.  have  been 
obtained,  making  it  possible  to  obtain  the  equivalent  of  a  large  number 
of  theoretical  plates  in  a  relatively  short  height.  Because  of  their  effi- 
ciency and  simplicity,  packed  towers  are  widely  used  for  laboratory 
columns.  However,  when  larger  packed  towers  are  used,  the  efficiency 
in  general  decreases,  and  H.E.T.P.  values  of  a  few  feet  are  more  com- 
mon for  columns  of  commercial  size. 

The  cause  of  the  poor  results  in  large-diameter  columns  is  apparently 
poor  liquid  distribution.  The  use  of  liquid  redistributor  plates  every 
few  feet  can  increase  the  effectiveness  of  the  units,  but  they  increase  the 
pressure  drop.  With  a  number  of  such  liquid  distributors,  the  values 


466  FRACTIONAL  DISTILLATION 

of  H.E.T.P.  can  be  made  reasonably  low  and  reproducible,  but  the 
tower  has  become  almost  equivalent  to  a  perforated  plate  column. 

The  values  of  H.E.T.P.  for  large  columns  are  so  random  that  any 
correlation  gives  only  a  qualitative  idea  of  what  might  be  expected. 
The  following  equation  is  presented  as  a  rough  guide  only,  and  it  would 
not  be  surprising  if  actual  values  in  some  cases  were  several  fold 
different. 

G 


H.E.T.P. 


+ 


0  5 


(17-17) 


where  H.E.T.P.  =  feet,  for  nominal  packing  size  less  than  one-tenth 

tower  diameter 
dt  =  tower  diameter,  ft. 

MQ  =  average  molecular  weight  of  vapor  phase 
0  =  superficial  mass  velocity  of  vapor,  Ib.  per  hr.  per 

sq.  ft. 

P.  =8  absolute  pressure,  atm. 

D  =  diffusivity  of  solute  gas  in  liquid,  sq.  cm.  per  sec. 
L  =  liquid  rate,  Ib.  per  hr.  per  sq.  ft.  tower  cross  section 
p'  as  liquid  viscosity,  poises 
p  an  liquid  density,  g.  per  cu.  cm. 
H  =  Henry's  law  constant  (Ib.  mols/cu.  ft.)  per  atm. 
HP  =  0.016  (KMi/p),  where  K  -  y/x,  ML  =  molecular 
weight  of  liquid 

The  individual  film  coefficients  are  based  mainly  on  the  results  of 
Sherwood  and  Holloway  (Ref .  29),  but  they  have  been  combined  with 
other  factors  and  the  equation  should  be  considered  purely  empirical- 
If  the  values  obtained  from  Eq.  (17-17)  are  greater  than  the  distance 
between  distributor  plates,  it  is  suggested  that  this  latter  difference  be 
employed. 

Packed  Tower  Example.  Equation  (17-17)  will  be  used  to  estimate  the 
H.E.T.P.  values  for  the  atmospheric  distillation  of  a  benzene-toluene  mixture  in  a 
packed  tower  5  ft.  in  diameter.  The  liquid  and  vapor  rates  are  480  and  400  Ib. 
mols  per  hr.,  respectively.  The  calculation  will  be  made  for  the  section  near  the 
bottom  of  the  column  where  the  liquid  and  vapor  are  essentially  toluene. 

Solution 

Area  of  column  -  5»(0.7854)  «  19.6 
0  -  ffl)(^2)  -  1,880  Ib.  per  hr.  per  sq.  ft. 


FRACTIONATING  COLUMN  PERFORMANCE  467 

L  -  -^p  -  2,250  Ib.  per  hr.  per  sq.  ft. 

MG  -92 
dt  -  5.0 

HP  -  0.016  (MA  -  0.016  [2Q(g52-1  (see  page  463  for  values) 
-  4.15 

The  diffusivity  for  the  system  was  estimated  to  be  1.0  X  10~"5  sq.  cm./sec.  at 
t°C.,  and  it  was  corrected  to  the  higher  temperature  assuming  that  the  diffusivity 
paried  inversely  as  the  1.5  power  of  the  liquid  viscosity.1  This  calculation  gave 
>  «  3.5  X  10~5  sq,  cm.  per  sec. 

ia»V_  3o)2(X) 


fjL\*''       f         °-OQ24         1°'B 
\PD)      **  [o.85(3.5  X  10~6)  J 

By  Eq.  (17-1?), 
H.E.T.P.  -  A  [l2(i,880)o  •  +  (3.5  x  ^JS)  (30,200)  (9)]  ~  6'5  ft' 

The  pressure  drop  for  this  tower  can  be  estimated  from  Eq.  (j#-40).  It  will  be 
ssumed  that  the  packing  size  is  3.0  in. 

AP 
h 

The  Reynolds  number,  »^  -  00024(242)  "  81° 
By  Eq.  (16-42), 

/  «     38 

*       810°-18 

From  the  table  on  page  438,  Aw  =  0.86. 

By  Eq.  (16-43),  ^-14+  0.0005(2,250)  -  2.5 

The  superficial  gas  velocity,  assuming  that  absolute  pressure  is  1.0  atm., 

400(359)  (385) 
U  "  19.6(3,600)(273) 

-  2.87  f  .p.s. 

AP       2(14)  (0.86)  (2.5)  (0.21 )  (2.87)* 
h    *  32.2(H) 

-  12.9  p.s.f .  per  ft. 

»  2,9  in.  of  liquid  toluene  per  ft.  of  length 

This  pressure  drop  is  high  indicating  that  the  tower  is  operating  near  the  flooding 
ondition.  Figure  16-4  would  give  a  flooding  velocity  of  about  4  f .p.s.  for  these 
onditions. 

1  See  ARN6UD,  Sc,D.  thesis  in  chemical  engineering,  M.I.T.,  1931. 


468 


FRACTIONAL  DISTILLATION 
TABLE  17-5 


Packing 

Vapor 
velocity,  f  .p.s. 

H.E.T.P., 
in. 

Open  tower  

0  25  to  1  7 

25  5  to  29  2 

H-  by  J^-in.  carbon  Raschig  rings  ... 
H-  by  H-in-  stoneware  Raschig  rings  . 
%.  by  JHj-in.  stoneware  Raschig  rings  .  .             ... 
}4r  by  Ji-in.  carbon  Raschig  rings  .  . 
H-  by  J£-in.  glass  Raschig  rings  .   . 
No.  19  aluminum  jack  chain  
J£-  in.  clay  Berl  saddles  
^-in.  aluminum  Berl  saddles.  . 

0.25  to  1.45 
0  8    to  1.55 
03    to  1  .  1 
0  15  to  0  5 
04    to  0  9 
0.1    tol  65 
0.05  to  1.6 
03    to  1  75 

6  0  to  11  3 
5  0  to    85 
3.8to    7.3 
4  7  to    60 
4  3  to    6.8 
4  2  to    89 
5  8to    7.0 
4  1  to    70 

6-mesh  carborundum  

01    to  0.35 

1  6  to    51 

J4-in.  aluminum  single-turn  helices 
JKe-in.  aluminum  single-turn  helices 
JHj2-in-  single-turn  stainless-steel  helices  , 
y^i-\VL.  single-turn  nickel  helices.          .    .  ^v 
J^-in.  single-turn  nickel  helices  ... 
%2-fa-  single-turn  stainless-steel  helices 
%-  by  %2-in-  carding  teeth     . 

0.7    to2.1 
0.1    to  1.8 
0  3    to  1.25 
0  55  to  1.65 
01    to  1.0 
0  15  to  0.95 
0.4    to  1  35 

5.0  to  10  7 
3  7  to    62 
4  Oto    5.4 
2  9  to    55 
2.9  to    5  5 
1.5  to    2  3 
2.  Oto    4  2 

Nomenclature 

a  «•  interfacial  area  of  bubble  or  total  contact  area  per  plate 
D  «  diffusivity,  cm.2/sec. 
d  «*  specific  gravity,  relative  to  water 
dp  *  diameter  of  packing,  ft. 
dt  «•  tower  diameter,  ft. 
E  «•  plate  efficiency 
E°  *•  over-all  plate  efficiency 
E°MV  *•  Murphree  vapor-phase  plate  efficiency 
E*MV  **  Murphree  point  efficiency 
E°ML  s«  Murphree  liquid-phase  plate  efficiency 
E°MT  «  apparent  plate  efficiency  with  entrainment 
e  *»  mols  of  liquid  entrained  per  mol  of  vapor 

O  «•  mols  of  gas  in  bubble  or  superficial  mass  velocity  of  vapor,  lb./(hr.)  (sq.  ft.) 
H  »  Henry's  law  constant,  (Ib.  mols)/(cu.  ft.)  (atm.) 
h  *  height  from  center  of  slots  to  top  of  weir,  in. 
hf  »  distance  from  top  of  liquid  level  to  center  of  slot  opening 
K  «  slope  of  equilibrium  curve  «  dy/dx 
K.  -  y/x 

KG  •»  over-all  mass-transfer  coefficient 
koa  «•  gas-film  mass-transfer  coefficient 
KL  ••  over-all  mass-transfer  coefficient 
kua  «  liquid-film  mass-transfer  coefficient 
27  **  liquid  in  slug 


FRACTIONATING  COLUMN  PERFORMANCE  469 

L  «•  liquid  rate,  lb./(hr.)(sq.  ft.  of  tower  cross  section) 
M  »•  molecular  weight  of  liquid 
I o  »  average  molecular  weight  of  vapor  phase 
P  »  total  pressure,  atm. 
w  —  slot  width,  in. 
a;  =•  mol  fraction  in  liquid 
x*  **  equilibrium  with  vapor  leaving 
#0  =*  liquid  leaving  plate 
#*  =  liquid  entering  plate 
y  =  mol  fraction  in  vapor 
yr  =  mol  fraction  in  bubble 
ye  —  equilibrium  with  liquid 
2/»  «  entering  plate 
y0  ««  leaving  plate 

2/J  »  equilibrium  with  liquid  leaving  plate 
M  =  viscosity  of  liquid,  centipoise 
M'  =  viscosity  of  liquid,  poise 
p  «  liquid  density,  g./cu.  cm. 
0  =  contact  time  of  bubble 

References 

L.  ATKINS  and  FRANKLIN,  Petroleum  Refiner,  16,  No.  1,  30  (1936). 
!.  BAKER,  CHILTON,  and  VERNON,  Trans.  Am.  Inst.  Chem.  Engrs.,  31,  296  (1935). 
8.  BROWN,  World  Power  Conference,  1936,  Trans.  Chem.  Eng.  Congress,  2,  330. 
L  BROWN,  SOUDERS,  NYLAND,  and  HESLER,  Ind.  Eng.  Chem.,  27,  383  (1935). 

1.  CAREY,  GRISWOLD,  LEWIS,  and  McADAMs,  Trans.  Am.  Inst.  Chem.  Engrs., 
30,  504  (1934). 

>.  COLBURN,  Ind.  Eng.  Chem.,  28,  526  (1936). 

T.  DRICKAMER  and  BRADFORD,  Trans.  Am.  Inst.  Chem.  Engrs.,  39,  319  (1943). 

5.  ETHERINGTON,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1949. 

).  FAIRBROTHER,  S.M.  thesis  in  chemical  engineering,  M.I.T.,  1939. 

).  FENSKE,  LAWROSKI,  and  TONGBERG,  Ind.  Eng.  Chem.,  30,  297  (1938). 

L.  FENSKE,  TONGBERG,  and  QUIGGLE,  Ind.  Eng.  Chem.,  26,  1169  (1934). 

5.  GADWA,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1936. 

J.  GEDDES,  Trans.  Am.  Inst.  Chem.  Engrs.,  42,  79  (1946). 

L  GTTNNESS,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1936, 

>.  HAUSEN,  Forschungsheftt  7,  177  (1936). 

5.  HORTON,  S.M.  thesis  in  chemical  engineering,  M.I.T.,  1937. 

r.  JANTZEN,  Dechema  Mon.,  6,  No.  48  (1932). 

*.  KEYES  and  BYMAN,  Univ.  Illinois  Eng.  Expt.  Sta.,  Butt.  329  (1941). 

J.  KIRSCHBAUM,  Forschungsheft,  8,  63  (1937). 

).  LEWIS,  Ind.  Eng.  Chem.,  14,  492  (1922). 

L.  LEWIS,  Ind.  Eng.  Chem.,  28,  399  (1936). 

2.  LEWIS  and  SMOLEY,  Bull.  Am.  Petroleum  InsLt  11,  Sec.  3,  No.  1,  73  (1930). 
I  LEWIS  and  WILDE,  Trans.  Am.  Inst.  Chem.  Engrs.,  21,  99  (1928). 

L  MURPHREE,  Ind.  Eng.  Chem.,  17,  747  (1925). 

5.  NORD,  Trans.  Am.  Inst.  Chem.  Engrs.,  42f  863  (1946). 

5.  O'CoNNELL,  Trans.  Am.  Inst.  Chem.  Engrs.,  42,  741  (1946). 


470  FRACTIONAL  DISTILLATION 

27.  PEAVY  and  BAKER,  Ind.  Eng.  Chem.,  29,  1056  (1937). 

28.  RAOATZ  and  RICHARDSON,  Calif.  Oil  World,  October,  1946, 

29.  SHERWOOD  and  HOLLO  WAY,  Trans.  Am.  Inst.  Chem.  Engrs.,  36,  39  (1940). 

30.  WALTER,  Sc.D.  thesis  in  chemical  engineering,  M.I.T.,  1940. 

31.  WALTER  and  SHERWOOD,  Ind.  Eng.  Chem.,  33,  493  (1941). 

32.  WEIMANN,  Chem.  Fabrik,  6,  411  (1933);  Z.  Ver.  deut.  Chem.,  No.  6  (1933). 

33.  WHITMAN,  Chem.  Met.  Eng.,  29,  No.  4  (1923). 


CHAPTER  18 
FRACTIONATING  COLUMN  AUXILIARIES 

In  addition  to  the  fractionating  column,  there  are  certain  auxiliaries 
necessary  for  its  operation.  These  include  the  condenser,  reboiler, 
pumps,  instruments,  valves,  and  receivers. 

Condensers.  The  condensers  are  tubular  heat  exchangers  used  to 
liquefy  the  overhead  vapor  to  produce  the  necessary  reflux.  For  small 
columns,  the  condensers  are  mounted  above  the  column,  and  the  con- 
densate  returns  to  the  system  by  gravity.  For  larger  installations,  the 
condensers  are  generally  mounted  below  the  top  of  the  column,  and  the 
reflux  is  pumped  back  to  the  system.  Placing  the  condensers  at  the 
lower  level  reduces  the  size  of  the  supporting  structure  and  makes  them 
more  accessible  for  cleaning.  The  design  of  condensers  is  largely  a 
problem  of  heat  transfer,  and  the  reader  is  referred  to  "Heat  Trans- 
mission" by  Me  Adams  (Ref.  1).  In  addition  to  the  heat  removal 
problems,  the  various  pressure  drops  should  be  evaluated  to  determine 
whether  the  vapor  and  condensate  will  flow  properly. 

Reboilers.  In  small  units  the  reboiler  is  frequently  mounted 
directly  below  the  column  and  consists  of  a  large  kettle  containing  a 
steam  coil.  In  large  installations  it  is  customary  to  build  the  reboiler 
as  a  separate  unit.  This  allows  easier  servicing  of  the  heat-transfer 
surface.  Where  fouling  of  the  heating  tubes  is  rapid,  two  reboilers  are 
often  included  to  allow  a  spare  that  can  be  used  while  the  other  is  being 
cleaned.  The  design  of  a  reboiler  is  also  mainly  a  heat-transfer  prob- 
lem, but  pressure  drops  and  adequate  settling  space  above  the  liquid 
should  be  considered.  Where  separate  reboilers  are  employed,  the 
vapor  is  piped  to  the  column,  and  the  pressure  drop  may  be  such  that  it 
hinders  the  liquid  flow  to  the  reboiler.  Considerable  splashing  of  the 
liquid  in  the  still  is  often  encountered  due  to  the  deep  liquid  level,  and 
a  settling  space  of  at  least  2  ft.  should  be  employed  with  a  cross-sec- 
tional area  two  to  three  times  that  of  the  column.  In  order  to  obtain 
a  large  cross-sectional  area,  horizontal  drums  are  frequently  used  foi 
reboilers,  and  these  units  can  be  arranged  to  give  longitudinal  flow  of 
the  liquid.  This  can  give  additional  separation  by  a  Rayleigh-type 
distillation.  The  reboiler  should  be  designed  to  obtain  the  desired 

471 


472  FRACTIONAL  DISTILLATION 

vapor  production,  and  any  extra  rectification  in  the  still  should  be  a 
secondary  consideration. 

When  live  steam  is  employed,  it  can  be  added  directly  under  the  bot- 
tom plate  of  the  column,  or  a  still  can  be  used  and  the  steam  intro- 
duced under  the  liquid  through  some  type  of  distributor.  Steam  is 
the  most  common  medium  for  the  stills,  but  in  some  cases  it  is  not 
desirable  because  (1)  the  still  temperature  is  too  high,  (2)  the  leakage 
of  water  into  the  system  may  be  undesirable  and  hazardous,  or  (3) 
the  still  temperature  may  be  so  low  that  freezing  may  be  encountered. 
In  some  cases,  organic  liquids  or  vapors,  such  as  Dowtherm,  are  used 
as  the  heating  medium,  or  the  liquid  at  the  bottom  of  the  column  may 
be  pumped  through  a  direct-fired  heater  and  partly  vaporized.  In 
this  last  case,  the  liquid-vapor  mixture  from  the  heater  is  passed 
through  a  separator,  with  the  vapor  going  to  the  column  and  the 
liquid  being  removed  as  bottoms,  although  it  may  be  recycled  through 
the  heater  to  obtain  additional  vaporization. 

Fractionating  Column  Control.  In  order  to  obtain  satisfactory 
performance  from  a  fractional  distillation  unit,  it  is  necessary  to  con- 
trol the  operating  conditions.  In  the  case  of  a  continuous  rectification 
system,  it  is  desirable,  and  almost  essential,  to  introduce  a  feed  of  rela- 
tively constant  composition  and  temperature  at  a  reasonably  constant 
rate.  In  general  the  system  is  to  make  a  specified  separation  with  the 
given  column  and  a  given  feed  plate,  and  this  leaves  two  main  variables 
to  be  controlled:  (1)  the  reflux  ratio  and  (2)  the  heat  supply  to  the  still. 
For  example,  consider  the  separation  of  a  binary  mixture  and  assume 
that  the  desired  separation  is  not  being  obtained.  The  reflux  ratio 
can  be  increased  which  should  increase  the  degree  of  separation,  but 
if  the  same  heat  supply  to  the  still  is  maintained,  the  overhead  product 
rate  will  decrease.  To  obtain  the  desired  percentage  of  the  feed  as 
overhead  product,  it  may  be  necessary  either  to  decrease  the  feed  rate 
or  to  increase  the  heat  supply  if  the  column  will  handle  the  added  load. 
The  problem  of  fractionating  tower  control  becomes  one  of  maintaining 
the  proper  balance  of  the  feed  rate,  reflux  ratio,  and  heat  supply  to  the 
still.  Manual  control  can  be  used,  but  automatic  control  instruments 
generally  do  a  better  job  because  operators  tend  to  overcorrect.  The 
problem  of  overcorrecting  is  especially  serious  in  rectification  because 
the  large  liquid  holdup  introduces  long  time  lags.  In  columns  with  a 
large  number  of  plates,  the  mols  of  liquid  held  up  in  the  system  may  be 
the  same  order  of  magnitude  as  the  vapor  or  liquid  throughput  per 
hour  and  may  correspond  to  several  times  the  feed  rate  per  hour, 
This  large  quantity  of  liquid  acts  like  a  flywheel  and,  smooths  out  the 


FRACTIONATING  COLUMN  AUXILIARIES 


473 


operation,  but  it  makes  the  response  to  corrective  action  very  slow  and 
leads  to  overcorrection. 

Instrumentation.  One  method  of  instrumentation  of  a  fractionating 
column  is  shown  in  Fig.  18-1.  The  feed  rate  to  the  column  is  held  con- 
stant by  the  feed-flow  controller.  The  overhead  vapor  is  condensed 


Feed  flow 
controller? 


Feed 

Reboiler 
vapor 
controller  i 


Reflux 
drum 


Vent 

Liquid  level 
controlhr^ 


„„ 

Reflux 


IX, 


product 


Reflux  flow 
controller 


Liquid  level 
controller-* 


Steam 


^ Steam  trap  Bottom  product 

FIG.  18-1.     Diagram  of  instrumentation  for  continuous  rectification  system. 

and  runs  to  a  reflux  drum  accumulator,  and  a  portion  of  the  condensate 
is  pumped  back  to  the  column  for  reflux.  The  excess  condensate  is  the 
overhead  product  and  is  removed  from  the  drum  by  the  liquid  level 
controller.  The  reflux  rate  is  regulated  by  the  reflux-flow  controller 
which  attempts  to  hold  the  temperature  near  the  top  of  the  column  at 
a  fixed  value.  The  steam  rate  to  the  still  is  controlled  by  a  reboiler 
vapor  controller  operating  on  a  temperature  indicator,  and  the  unva- 
porized  liquid  is  removed  from  the  still  by  a  liquid-level  controller. 
In  the  operation  of  the  system  illustrated  in  Fig.  18-1,  the  feed-flow, 


474  FRACTIONAL  DISTILLATION 

reflux-flow,  and  reboiler  vapor  controllers  would  be  set  at  the  desired 
values.  If  the  temperature  at  the  top  control  point  became  lower 
than  the  desired  value,  the  reflux-flow  controller  would  reduce  the 
reflux  rate  which  would  increase  the  product  withdrawal  rate  and  raise 
the  temperature.  If  the  temperature  were  too  high,  the  controller 
would  take  the  opposite  action.  If  the  temperature  at  the  reboiler 
control  point  becomes  lower  than  desired,  the  controller  will  increase 
the  steam  supply  to  the  still,  resulting  in  a  higher  vapor  rate.  This  will 
give  better  stripping  of  the  light  components  and  will  raise  the  tem- 
perature at  the  control  point.  This  action  of  the  reboiler  vapor  con- 
troller will  lower  the  reflux  ratio  by  increasing  the  overhead  product 
rate  and  will  probably  require  the  reflux-flow  controller  to  increase  the 
reflux  rate  to  hold  the  top  temperature  down.  In  this  system  the 
reboiler  vapor  controller  operates  to  give  the  desired  temperature  at 
the  bottom  control  point  which  presumably  gives  the  specified  bottom 
product.  The  reflux-flow  controller  operates  to  give  the  desired  tem- 
perature difference  between  the  two  control  points.  The  system  is 
operable  because  for  a  given  number  of  plates  the  degree  of  separation 
can  be  varied  by  changing  the  reflux  ratio.  The  feed  rate  and  the  tem- 
perature values  that  are  fixed  must  obviously  be  within  the  fractionat- 
ing capabilities  of  the  system. 

A  number  of  alternate  control  systems  can  be  employed.  Thus  a 
fixed  reflux  ratio  could  be  used  with  the  top  control  point  regulating 
the  feed  rate  to  the  column  with  the  bottom  control  operating  the  same 
as  before.  In  case  the  overhead  product  is  difficult  to  liquefy,  a  partial 
condenser  may  be  employed,  and  this  requires  a  different  method  of 
control.  A  typical  case  of  this  type  is  gasoline  stabilization  where  a 
small  amount  of  €2,  Ca,  and  C4  hydrocarbons  are  removed  to  obtain 
the  desired  volatility  of  the  bottoms  product.  For  this  case,  the  sys- 
tem of  Fig.  18-1  could  be  modified  as  shown  in  Fig.  18-2  by  removing 
the  overhead  product  through  the  vent  line.  This  vent  line  would  be 
equipped  with  a  pressure  controller  which  would  adjust  the  vapor 
removal  rate  to  maintain  the  desired  operating  pressure.  The  liquid- 
level  controller  would  be  used  to  adjust  the  cooling  water  rate  to  the 
condenser  instead  of  the  product  withdrawal  rate.  The  reflux-flow 
controller  would  operate  to  maintain  the  top  temperature  as  before, 
and  the  liquid-level  controller  would  decrease  the  cooling  water  rate  if 
the  level  became  too  high,  thereby  decreasing  the  rate  of  condensate 
production  and  increasing  the  overhead  vapor  rate. 

Control  Variables.  One  of  the  difficult  problems  in  the  automatic 
control  of  fractionating  towers  is  finding  an  easily  measured  charac- 


FRACTIONATING  COLUMN  AUXILIARIES 


475 


teristic  that  will  ensure  the  desired  separation.  Temperature  is  the 
most  commonly  used  factor,  but  it  is  not  always  a  satisfactory  criterion. 
Thus,  if  the  product  is  of  high  purity  and  contains  only  a  small  amount 
of  other  constituents,  these  can  vary  several  fold  without  significantly 
changing  the  equilibrium  temperature.  This  is  particularly  serious 
in  the  separation  of  close  boiling  constituents.  In  the  case  of  multi- 
component  mixtures,  temperature  is  not  a  good  criterion  of  composi- 
tion, but  it  can  be  a  satisfactory  indication  of  volatility.  The  problem 

Overhead 
product 


^Pressure 
controller 


Liquid  level 
controller 


Reflux  flow 
controller 


FIG.  18-2.     Diagram  for  partial  condenser  operation. 

is  particularly  difficult  in  extractive  distillation  systems  where  the 
presence  of  the  large  quantity  of  solvent  masks  the  effect  of  composi- 
tion of  the  key  components  on  the  temperature.  Temperature  control 
is  also  sensitive  to  the  column  pressure. 

In  some  of  these  cases,  the  effectiveness  of  temperature  can  be 
improved  by  proper  location  of  the  temperature  control  point.  For 
example,  in  the  case  of  high  product  purity,  the  temperature  control 
point  can  be  placed  several  plates  f  rom^the  end  of  the  column  at  a  point 
where  the  minor  constituents  have  higher  concentrations.  This  results 
in  a  larger  temperature  variation  which  makes  the  instrumentation 
easier,  but  it  removes  the  direct  control  on  the  product. 

Other  control  factors  besides  temperature  have  been  used,  such  as 
density  of  the  liquid,  refractive  index  of  the  liquid,  infrared  or  other 
spectrographic  types  of  analysis,  and  freezing  point.  The  factor  should 
be  one  which  gives  an  indication  of  product  composition  and  which  can 
be  easily  and  rapidly  determined  by  a  relatively  inexpensive,  stable 


476 


FRACTIONAL  DISTILLATION 


instrument  capable  of  operating  electronic  or  pneumatic  equipment. 
The  spectographic  type  of  instruments  should  be  very  useful  because 
they  can  frequently  be  adjusted  to  indicate  the  amount  of  an  impurity 
present  in  extremely  small  amounts.  Unfortunately  they  are  some- 
what delicate  and  need  frequent  adjustment,  but  these  defects  can 
undoubtedly  be  eliminated. 


0.1 


20      30 


nh 

FIG.  18-3. 


Rate  of  Approach  to  Equilibrium.  Corrective  action  or  other 
changes  in  the  operating  conditions  introduce  a  transient  condition  into 
the  system.  Consider  a  section  of  a  column  operating  at  steady-state 
conditions.  Assume  that  a  change  is  made  on  one  of  the  plates  (such 
as  a  change  in  the  feed  composition  to  the  feed  plate),  and  the  problem 
is  to  determine  how  rapidly  other  plates  will  approach  their  new 
equilibrium  condition.  The  exact  mathematical  solution  of  the  gen- 
eral case  is  very  complex,  but  solutions  based  on  certain  approxima- 
tions can  be  obtained.  The  curves  given  in  Fig.  18-3,  based  on  a 
number  of  stepwise  integrations,  should  be  helpful  in  obtaining  approxi- 
mate results.  In  this  figure  the  ordinate,  F,  is  the  fractional  approach 


FRACTIONATING  COLUMN  AUXILIARIES  477 

to  the  new  equilibrium  conditions  of  a  plate  which  is  n  plates  from  the 
one  on  which  the  change  in  composition  was  made.  The  value  of  F  is 
defined  as, 


(yn)*  -  y°n  * 

where  yn  =  vapor  leaving  plate  at  time  0 

(2/n)oo  =  new  equilibrium  value  of  yn,  i.e.,  value  of  yn  at  0  =  °o 
y°n  =  equilibrium  vapor  composition  before  change  in  condi- 

tions were  made 

The  abscissa  is  the  dimensionless  ratio  of  liquid  flow  through  the 
section  to  liquid  holdup  in  the  section.  This  group  is 

00 
nh 

where  0  =  overflow  rate,  mols  per  unit  time 
0  =  time 
n  =  number  of  plates  in  section  between  where  conditions  were 

modified  at  6  =  0  and  plate  n 
h  =  holdup  per  plate,  mols 

The  value  of  F  decreases  rapidly  as  (00/nh)  increases,  indicating  that 
the  section  approaches  the  new  equilibrium  after  a  change,  more  rap- 
idly for  large  values  of  the  overflow  rate,  small  values  of  holdup  per 
plate,  and  a  small  number  of  plates.  Lines  are  given  on  Fig.  18-3  for 
different  relative  volatilities  between  the  key  components,  and  for  mul- 
ticomponent  mixtures  the  value  of  F  should  be  applied  to  the  light  key 
component.  For  a  change  in  condition  that  affects  all  plates  almost 
simultaneously,  such  as  a  change  in  reflux  ratio,  it  is  suggested  that  the 
correlation  of  Fig.  18-3  be  employed  with  n  =  2.0. 

To  bring  a  section  of  a  column  to  a  reasonable  approach  to  the  new 
equilibrium  would  require,  according  to  Fig.'  18-3,  a  liquid  throughput 
of  two  to  ten  times  the  liquid  holdup  in  the  section.  For  usual  design 
conditions,  a  tower  operating  at  atmospheric  pressure  will  have  a  liquid 
throughput  per  minute  equal  to  about  the  holdup  per  plate.  For  high- 
pressure  towers  the  ratio  of  liquid  holdup  to  liquid  flow  per  plate  may 
be  as  low  as  0.1  ;  for  vacuum  towers  it  may  be  as  high  as  5.0.  Thus  an 
atmospheric  pressure  column  with  20  plates  above  the  feed  plate  may 
require  from  0.5  to  3  hr.  to  steady  down  after  a  change  in  feed  composi- 
tion, while  a  high-pressure  tower  would  adjust  itself  much  more  rapidly. 

Figure  18-3  indicates  that  a  change  in  the  vapor  composition  entering 
a  section  will  make  a  difference  in  the  plate  just  above  much  sooner 


478  FRACTIONAL  DISTILLATION 

than  on  plates  farther  removed,  while  a  change  in  the  reflux  ratio  would 
change  (y)«>  for  all  the  plates  almost  simultaneously.  Thus  changes 
in  feed  composition  will  be  noticeable  much  sooner  at  plates  just  above 
and  below  the  feed  plate,  and  instruments  located  in  these  positions 
would  be  able  to  correct  for  such  changes  in  composition  much  more 
rapidly.  However,  in  nmlticornponent  mixtures  it  may  be  very  diffi- 
cult to  relate  conditions  at  these  locations  to  the  desired  separation. 
While  the  instruments  can  be  set  to  control  the  temperature  at  the  feed 
plate  region,  this  may  not  give  control  of  the  product  compositions. 
Control  points  near  the  ends  of  the  columns  are  less  affected  by  changes 
in  feed  compositions  but  do  not  anticipate  variations  so  rapidly  and  are 
generally  less  sensitive.  Changes  due  to  reflux  ratio  or  vapor  rate  are 
most  apparent  where  the  change  in  composition  of  the  key  components 
per  plate  is  greatest.  This  usually  occurs  in  the  intermediate  section 
somewhat  above  and  below  the  feed  plate. 

In  most  cases  with  temperature  regulation,  the  control  point  for  the 
top  should  be  down  enough  plates  from  the  top  to  gain  the  amplified 
temperature  difference  and  the  anticipatory  effect  of  being  closer  to  the 
feed  plate,  but  it  should  not  be  so  close  to  the  feed  plate  that  it  is 
appreciably  affected  by  the  components  heavier  than  the  heavy  key 
component.  Similar  considerations  apply  to  the  section  below  the 
feed  plate. 

Reference 

1.  McADAMS,  "Heat  Transmission,"  2d  ed.,  McGraw-Hill  Book  Company,  Inc., 
New  York,  1942. 


APPENDIX 

TABLE  OF  ENTHALPIES,  on  LATENT  HEATS  OF  VAPORIZATION 


Substance 

Boiling  point, 
°C. 

Molecular 
Weight 

Latent  heat  in 
calories  per  gram 
at  boiling  point 

Acetal  . 

104.0 

118.1 

66  2 

Acetaldehyde 

20  8 

44.0 

134.6 

Acetic  acid 

118  7 

60.0 

89.8 

Acetic  anhydride 

136  4 

102.1 

66.1 

Acetone  .  . 

56.6 

58.1 

125.3 

Acetyl  chloride 

55  6 

78.5 

78.9 

Ammonia  .  . 

-  34  7 

17.0 

341  0 

Aniline  .... 

183.9 

93,1 

109  6 

Benzaldehyde 

178  3 

106.1 

86.6 

Benzene  .  . 

80.2 

78.1 

93.5 

Benzyl  alcohol 

205  0 

108.1 

98  5 

Brombenzene 

155.5 

157.0 

57.9 

Butyl  alcohol  (n) 

117.6 

74.1 

143  3 

Butyl  alcohol  (iso) 

107  9 

74.1 

138  9 

Butyric  acid  (iso) 

162.2 

88.1 

114  0 

Carbon  tetrachloride 

76.8 

153.8 

46  4 

Chlorbenzene  . 

131.8 

112.5 

75.9 

Chloroform  . 

61.2 

119.4 

58  9 

Cresol  (m)   , 

200  5 

108.1 

100  5 

Ethyl  bromide  . 

38.2 

109.0 

60  4 

Ethyl  iodide  — 

72.3 

155.9 

47  6 

Formic  acid  .  . 

100.8 

46.0 

120  4 

Glycol  

197.1 

62.1 

190  9 

Heptane  .    . 

98.4 

100.2 

74.0 

Hexane     

69.0 

86.1 

79  2 

Iso  amyl  alcohol  .  . 

130.1 

88.1 

125  1 

Iso  propyl  alcohol  . 

82.9 

60.1 

161.1 

Methyl  alcohol  .  . 

64.7 

32.0 

261.7 

Methyl  chloride 

-  24.1 

50.5 

96.  9  (at  0°C.) 

Methyl  iodide  . 

42.4 

141.9 

46.0 

Methyl  ethyl  ketone 

81.0 

72.1 

103.5 

Methyl  aniline  .  .     . 

193.8 

107.1 

95  5 

Nitrobenzene 

208.3 

123.1 

79.2(atl51.5°C.) 

Octane  

125.8 

114.2 

71.1 

Pentane  ... 

36.3 

72.1 

85.8 

Propyl  alcohol  (w)  . 

97  4 

60.1 

162.6 

Toluene  

110.4 

92.1 

86.8 

o-Toluidene  . 

203.3 

107.1 

95.1 

Water.                                .     . 

100.0 

18.0 

536.6 

o-Xylene  .  .                            .... 

144  0, 

106  1 

82.5 

479 


APPENDIX 

Vapor  Pressure,  Pounds  per  Square  Inch 

10  100 LOOQ IO.OOCL 


K£Y 

Numbers- No  of  carbon 
atoms  per  molecule 
Unpnmed  Hos  -Paraffins 
Primed  Nos«0lefms 
6"  s  Benzene 
44  W  *  Water 


IpOO  IQOOO7  ^^ ^t!OQ 

Pressure,  Millimeters          """ 


Vqpor 


FIG.  1.  Cox  chart  for  extrapolating  vapor-pressure  temperature  curves.  (Walker, 
Lewis,  Me  Adams,  and  GUliland,  "Principles  of  Chemical  Engineering,"  3d  ed.,  McGraw- 
Hill  Book  Company,  Inc.,  New  York,  1937.) 


481 


482  FRACTIONAL  DISTILLATION 

Specific  heat*  Btu/(lb)(DegF)«  Pcu/(Lb)(DegC) 


NO 

LIQUID 

RANGE  DE6.C 

SPECIFIC 
HEAT 

29 

ACETIC  ACID    I00°7o 
ACETONE 
AMMONIA 

-7^1 

37 

AMYL  ALCOHOL 

—  50—  25 

•» 

26 

AMYL  ACETATE 

0—  \00 

M> 

DEGF 
400  — 

30 
23 
27 

ANILINE 
BENZENE 
BENZYL  ALCOHOL 

10-  80 

t< 

^-02 

~ 

J? 

51 

BRINE^STo  Na  C? 

-40-  20 

iol2< 

1 

44 
2 

BUTYL  ALCOHOL 
CARBON'  DISULPHIDE 
CARBON  TETRACHLORIDE 

^LL0°RToERNMZENE 

0-100 

"?§:  II 
8:'°5°o 

50 

E-0.5 

300  -i 

4 
1 

16 

DECANE 
\  D1CHLOROETHANE 
DICHLORO  METHANE 
DIPHENYL 
DIPHENYLMETHANE 
DIPHENYL  OXIDE 

-80-  25 
-30-  60 
-40-  50 
80-  120 
30-  100 
0-200 

60     06A 

0                ?0       9    010 
08      0 

^-0.4 

1 

16 

ft 

DOWTHERM  A 
ETHYL  ACETATE 
ETHYL  ALCOHOL  100°To 

0-200 
-50-  25 
30-  80 

HO   OQ  O'3A 

Z 

200  -E 

46 
50 
25 

ETHYL  ALCOHOL    95  7o 
ETHYL  ALCOHOL     50% 
ETHYL  BENZENE 
ETHYL  BROMIDE 
ETHYL  CHLORIDE 
ETHYL  ETHER 
ETHYL  IODIDE 

20-  80 
20-  80 

o-ioo 

-50-  40 
-100-  25 
0-100 

M      ^Do     O9I 
2^001719      Z]  024 

250  23 
7A026 

27     2^00  31   34 

r-as 

_;.r;— 

39 

ETHYLENE  6LYCOL 

-40-200 

z* 

E 

t-oQ        wfivoD 
37               ^  ?? 

z 

I 

AO  4IQ  45       °  ^l 

3-0.0 

100-i 

•j/         O    /%      AA                  O 

^"S  ~>  T"4                    Q                           *IV 

^^ 

-Z 

46°    047 

z 

i 

NQ 

LIQUID 

WN6EDEGC 

049 

z 

~E 

2A 

rREON"IICCCf3F) 

-20-   70 

—  07 

» 

6 

FREON-I2(CCI2F2) 

-40-    15 

— 

•E 

7^ 

•REON-2l(CHCI2t) 
rREON-22(CHCTF2) 

-20-   70 
-20-   60 

- 

0  -r 

3A 

38 

"REON-II3(CCI2F-CCIF2) 
SLYCEROL 

-20-   70 
-40-   20 

z 

28 

HEPTANE 

0-    60 

—  . 

-~ 

35 
48 

HEXANE 
-iYDROCHLORICACID.30% 

-80-   20 
20-  100 

050                                510 

Z-Q8 

r 

41    ! 

SOAMYL  ALCOHOL 

0-  100 

M. 

„,- 

43 

SOBUTYL  ALCOHOL 

0-  100 

z 

— 

47 

SOPROPYL  ALCOHOL 

-20-    50 

— 

- 

31 

SOPROPYL  ETHER 

-80-   20 

— 

- 

40 

SA 

Y|  ETHYL  ALCOHOL 
METHYL  CHLORIDE 

:8§:  io 

z 

-too  -i 

14 
12 

^T^/Nf^ 

T  ?o8 

Z~a9 

JE 

34   ( 
33 

^ONANE 
XTANE 

*:  8 

z 

z 

3 

5ERCHLORETHYLENE 

•30-  140 

z* 

^r 

ii 

PROPYL  ALCOHOL 
3YRIDINE 

-20-  100 
^50-  25 

52 

z 

,? 

sui:ffii/^oio)aoi  *8% 

-20-  Al 

O                           55O 

E-u> 

23 

TOLUENE 

0-   60 

•• 

WATER 

10-  200 

!•> 

0-  100 

18 

(YLtNt  MElA 

17 

(YLENE  PARA 

0-  100 

FIG.  2.  True  specific  heats  of  liquids.  (Ch&ton,  Cdbwrn,  and  Vernon,  based  mainly 
on  data  from  "International  Critical  Tables,"  McGraw-Hill  Book  Company,  Inc.,  New 
York,  1926-1930.) 


APPENDIX 


483 


0      10     20 

Mol  per  cenf  hydrocarbon 
30      40      50     60      70     * 

0      90 

o.' 

70 

j"4 

—  60 

60  V  o 

o 

\ 

^o 

\ 

S 

—  «to 

50  — 

"^W1^" 

\ 

v 

xK 

—  1ft 

40  —  - 

*  yfc 

V 

\ 

^ 

Ov 

& 

r"\ 

\ 

V 

y 

—  30 

30     ~ 

'*- 

A 

\ 

\  1 

\^*gA 

^ 

\[  * 

>^        ,Q 

—  70 

20     - 

Sjo 

X 

\ 

[Z^2v4~- 

\ 

,  « 

\ 

N-i^Sv-l- 

10 

10  

pS 

VS 

\   XS 

0 

o 

IX    I  ^ 

>^  jrsf 

<^( 

—   fl 

H** 

-*• 

i^a* 

rv       \.  '  »^ 

xS 

_      < 

_       — 

i-i^ik 

I*     &\u                 >VB 

issS 

-JO 

11          "    IV 

1    1 

X<(             >i 

v          '\  ^J 

-eo-' 

•  A               t 

rs  -Atmospheric  bor/ing 

rH  *  At/nosphefc  boiling 
ternperafare,  hydro  cttf 

Ljb^     ^X 

N     N 

S.  i 

L       -,„-    -  ou,   rr 

-9A 

rTV 

"i  <>  i?  'i  *^\ 

CV 

**i!sJSl. 

\       \. 

.. 

-1  h 

'ton  *X 

_j_    ^|r 

-^ 

OU 

0\ 

90  ^^ 

> 

•  X      *» 

-40  — 

l/fo 

Hy^foCorbofi$  /w 

>v^ 

jrecarbons 

2^J 

>  » 

V  w 

~^KffoT$    F 

o 

* 

^<  0 

s 

Kehnes 

v 

0   * 

-60 

AMydes 

r 

Phenols  & 

a 

( 

0  0 

"*"  /v 

~^T  crgsoh 

] 

1 

.    M  (    ,  ,f 

""80 

JO"  2f  '30'    40     50      60     70     80     90    100 
Mo)  per  cert  hydrocarbon 

i'jto.  3.  Composition  of  binary  azeotropes—  hydrocarbons  and  various  solvents. 
[Courtesy  of  Meissner  and  Greenfield,  Industrial  and  Engineering  Chemistry,  40,  438 
(194S).} 


484 


FRACTIONAL  DISTILLATION 


fU 

60 
50 

40 
30 

20 
10 

S 
1 

-20 
-30 
-40 
-50 
-60 
-70 
-80, 

4 

X 

0 

/ 

o  Hydrocarbons  ana 

velds 

/ 

0 

•  Hatogenated  hydrocarbons  and  adds 

/ 

>/ 

\ 

s 

0 

/ 

. 

X 

s^ 

0 

o 

/ 

/ 

0 

\ 

, 

K 

>TJ 

0 

\! 

\ 

s* 

• 

/° 

• 

%vj 

• 

\ 

4' 

'• 

0 

[? 

% 

p 

f 

*• 

>• 

\ 

§ 

>o°y 

( 

N 

> 

( 

1 

0° 

/ 

^ 

^ 

°/ 

> 

t_^ 

^) 

i 

1 

\ 

s 

; 

' 

\ 

/ 

\ 

°°/ 

0 

T  • 

«• 

Atmospheric  boiling 

^ 

\ 

< 

Atmospheric  boiling 
kmperati/re,  hydrocarbonf  *K 

1 

^ 

V 

-'0 

r 

\ 

c 

0 

t 

• 

c 

,,^-K 

/ 

4 

I  c 

/ 

)         10        20        30        40        50  «,     60        70        80        90        101 

Mol  per  cent ' 

*  Abscissa  is  mol  per  cent  halogen  a  ted  hydrocarbons  for  halogenated  hydrocarbons-acids  sys- 
tems and  mol  per  cent  acids  in  hydrocarbons-acids  systems. 

FIG.   4.    Composition   of   binary   azeotropes — hydrocarbons   and   carboxylic   acids 
[Courtesy  of  Mewmer  and  Greenfield,  Ind.  Eng.  Chem.,  40,  438  (1948).] 


APPENDIX 


485 


Table  of  "K*  values 

Alcohols 

ill 

• 

i 

1 

I 

I 

!i 

ii 

1 

Aromattc 

Symbol 

o 

n" 

o 

o 

• 

• 

a 

0 

<K 

4 

Cyclic 
nonaromafic 
Monoolefins 

-2 
-5 

-5 
•8 

-15 

tj 

^ 

« 

v 

« 

+/J 

Diolefins 
Aromatics 

•10 

-10 
-/J 

•20 

-5 

+/5 

^r 

f 

/:/ 

^ 

T  -  No  data             ®  •  Not  plotted 

f 

**ffi-(W)*K 

it 

0 

o 

• 

o2 

, 

/ 

Atmospheric  boiling  points,  *K 

TI  •  Low  boiling  component 
TT  *  High  boiling  component 
Ts  'Solvent 
T2  »  Azeotrope 

, 

^ 

o 

0^ 

•^ 

~ 

4 

=4 

LSs__^ 

nnS 

| 

rfjd 

K 

&«£ 

I* 

o 

u 

R 

^» 

0           0           10          20          30          40          50          60          70          00 
X 

FIG.  5.     Atmospheric  boiling  points  of  binary  azeotropes — hydrocarbons  and  various 
solvents.     [Courtesy  of  Meiaaner  and  Greenfield,  Ind.  Eng.  Chem.,  40,  438  (1948).] 


INDEX 


Absolute  alcohol,  azeotropic  distillation 
of,  313 

two-tower  system  for,  206 
Activity  coefficients,  50,  54 
Adsorption  factor  method,  345 
Allowable  gas  and  liquid  velocities,  439 
Allowable  vapor  velocity,  430 
Amagat's  law,  38 

Analytical  equations,  binary  mixtures, 
174 

finite  reflux  ratio,  176 

minimum  reflux  ratio,  176 

multicomponent  mixtures,  354 

straight  equilibrium  curve,  183 

total  reflux,  174 

use  of,  181 

A.S.T.M.  distillation,  325 
Auxiliaries  for  fractionating  column,  471 
Azeotrope,  21 

effect  of  pressure  on,  204 

example,  205 

Margules  equation  for,  21 

maximum  boiling-point  type,  21 

minimum  boiling-point  type,  20,  95 

pseudo-,  20 

separation  of  binary,  196 
Azeotropic  distillation,*  285,  312 

absolute  alcohol,  313 

agents  for,  287 

diagram  for,  313 

minimum  reflux  ratio,  321 

total  reflux,  321 

Van  Laar  equation  for,  286 
Azeotropic  mixtures,  21,  192 


B 

Batch  distillation,  108,  370 
constant  distillate  composition, 
constant  reflux  ratio,  374 


\ 


4*7 


Batch  distillation,  equations  for    108, 

110 

example,  108 
inverted,  387 
with  liquid  holdup,  380 
without  liquid  holdup,  370 
minimum  vapor  requirements  377 
multicomponent,  383 
split-tower  system,  387 
total  reflux,  376,  384 
Benzene,  dehydration  of,  89 
Binary  mixtures,   analytical  equations 

for,  174 

batch  distillation  of,  370 
rectification  of,  118 
separation  of  azeotropes,  196 
special,  192 

work  of  separating,  162 
Bubble  caps,  design  of,  404 

pressure  drop  through,  396 
Bubble-cap  plates,  diagram  of,  403 
liquid  flow  on,  409 
liquid  gradient  on,  412 
pressure  drop  of,  396 
due  to  liquid  head,  408 
through  risers,  405 
through  slots,  405 
for  vapor  flow,  404 
Bubble-point  curves,  7,  8,  79 


Calculation  of  vapor-liquid  equilibria, 

26 

Cascade  plate,  424 
Chemical   reactions   and   rectification, 

361 

Clark  equation,  60 
Completely  miscible  mixtures,  26 
Complex  mixtures,  laboratory  studies 

of,  332 
rectification  of,  325 


488 


FRACTIONAL  DISTILLATION 


Condensation,  differential  partial,  116 

equilibrium  partial,  116 

fractional,  102 

partial,  115 
Condenser,  441 

leak  in,  192 

partial,  116,  132,  475 

total,  132 

Constant-boiling  mixtures,  20 
Control  of  fractionating  column,  472 
Control  variables,  474 
Cost  of  fractionation,  130 
Critical  region,  8,  79 
Cross  flow  in  column,  448 

D 

Dalton's  law,  28,  46 

Dal  ton's  and  Raoult's  law,  28,  44 

Definitions,  of  fractional  distillation,  1 

of  plate  efficiency,  445 

of  theoretical  plate,  119,  122 
Degrees  of  freedom  in  distillation  sys- 
tem, 215 

for  rectifying  column,  216 
Determination  of  vapor-liquid  equilib- 
ria, 3 
Dew-point    and    bubble-point    curves, 

7,  8,  79 
Diagrams,  Henry's  law,  28 

Raoult's  law,  27 

temperature-composition,  16 

vapor-liquid,  16 
Differential  distillation,  107 
Distillation,  A.S.T.M.,  325 

azeotropic,  285 

batch,  108,  118,  370 

continuous,  118 

degrees  of  freedom  in,  215 

differential,  107,  108 

efficiency  of,  162 

extractive,  285 

multiple,  103 

simple,  107 

steam,  111 

successive,  101 

true-boiling-point  curves,  325 

use  of  instrumentation  in,  362 

vacuum,  113,  362 
Down  pipes,  409 


Duhem  equation,  47 

application  of,  49 

constant  pressure,  48 

constant  temperature,  47 

example  for,  50 

generalization  of,  48 

use  of,  50 
Dumping  of  liquid  in  column,  419 

E 

Efficiency,  of  bubble-cap  plates,  445 
correlation  plot,  457 
effect  on,  of  cross  flow,  448 

of  entrainment,  454 
equation  for,  462 
experimental  data  on,  455 
Murphree,  445 
over-all  column,  445 
of  distillation  column,  162 

methods  of  increasing,  167 
of  packed  towers,  464 
Enthalpy  balances,  120,  139 
below  feed  plate,  145 
general  case,  145 
Enthalpy  calculations,  141,  142 
of  liquid,  139 
vapor,  140 
Entrainment,  424 

Equilibrium,     calculation     of     vapor- 
liquid,  26 

modification  of,  286 
presentation  of  vapor-liquid,  16 
rate  of  approach  to,  476 
vapor-liquid,  3,  18 

Equilibrium  constants,  43,  215,  223 
Equilibrium-curve  design  method,  350 
Escaping  tendency,  26 
Evaporation  rate,  394 
Examples,    adsorption-factor    method, 

345 

analytical  equations,  181,  354 
azeotrope  separation,  205 
azeotropic  distillation,  313 
batch  distillation,  108,  373,  385,  387, 

389 

benzene-toluene-xylene,  219 
binary  batch  distillation,  373 
bubble-cap  plate  design,  433 
chemical  reactions  in  rectification,  363 


INDEX 


489 


Examples,  condenser  leak,  192 
Duhem  equation,  50 
enthalpy-composition   method,    141, 

150,  152 

extractive  distillation,  304 
feed-plate  location,  241,  248 
fugacity  calculations,  44 
gasoline  stabilization,  261 
graphical    correlations    for    column 

design,  347 
immiscible  liquids,  85 
isopropyl  alcohol  stripping,  194 
Margules  equation,  65 
minimum  reflux  ratio,  258 
modified  equilibrium-curve  method, 

350 

net  rate  of  evaporation,  394 
nitric  acid  concentration,  298 
optimum  reflux  ratio,  130 
packed-tower  design,  466 
phenol-water  separation,  198 
plate  efficiency,  463 
Raoult's  and  Dalton's  law,  30 
reduced  relative  volatility  method, 

341 

steam  distillation,  113 
tar-acid  distillation,  236 
Thiele-Geddes  method,  326 
Van  Laar  equation,  65 
Extractive  distillation,  285,  290 
choice  of  agents  for,  287 
diagram  for,  291 
example,  304 
feed-plate  location,  295 
minimum  number  of  plates  for,  293 
minimum  reflux  ratio,  293 
vapor-liquid  equilibrium  for,  286 


6,  f  1, 


148,  241, 


Falling-film  unit,  399 

diagram  of,  400 
Feed  condition,  126 
Feed-plate  location,  126, 

248,  270,  295 
Feed-plate  matching,  265 
Fractional  condensation,  102 
Fractional  distillation,  definition  of.  1 
Fractionating  column,  118 


Fractionating   column,  auxiliaries  for, 
471 

control  of,  472 

design  of,  121,  401 

diagram  of,  404 

performance  of,  445 

transients  in,  476 
Fractionation,  cost  of,  130 

diagram  for,  102 

efficiency  of,  165 

general  methods  of,  101 


Gas-law  deviations,  36,  39 
Gasoline  stabilization,  261 
Graphical  correlation  method,  347 

H 

Heat  of  mixing,  143 

Van  Laar  equation  for,  144 
Heat  economy,  162 

Height  equivalent  to  theoretical  plate 
(H.E.T.P.),  187,  464 

equation  for,  466 

Height  of  transfer  unit  (H.T.U.),  188 
Henry's  law,  27,  93 

diagram  for,  28 
Hydraulic  gradient  on  plate,  412 


Immiscible  liquids,  85 

example  for,  85 

plot  for,  86 

Inhibitors  used  in  distillation,  362 
Instrumentation,  473 

K 

K  values,  41,  42 
Key  components,  217 


Latent  heat  of  vaporization,  142 
Lewis  method,  178 
Lewis  and  Cope  method,  229 
use  of,  231 


490 


FRACTIONAL  DISTILLATION 


Lewis  and  Matheson  method,  219 
Lewis  and  Randall  rule,  37,  39,  56,  83 
Lewis  and  Wilde  method,  329 
Liquid,  enthalpy  of,  140 
Liquid-air  fractionation,  150 
Liquid  head  on  bubble  plate,  408,  418 

gradient,  412 

in  column,  428 
Liquid  phase,  39 
Logarithmic  plotting,  127 

M 

McCabe-Thiele  method,  123 

diagram  for,  124 
Margules  equation,  50,  54,  55 
for  azeotropic  composition,  204 
corrected  for  temperature,  56 
%  evaluation  of,  62 
example  of  use  of,  65 
plot  for,  72 
Mass  transfer,  399 
Maximum  boiling  azeotrope,  21,  95 
Maximum  boiling  mixture,  21 
Mean  free  path,  398 
Methods,  alternate  design,  for  multi- 
component  mixtures,  336 
fractionation,  101 
graphical  correlation,  347 
Lewis  and  Cope,  229 

use  of,  231 

Lewis  and  Matheson,  219 
Lewis  and  Wilde,  329 
McCabe-Thiele,  123 

diagram  for,  124 
Ponchon-Savarit      (see     Ponchon 

Savarit  method) 
reduced  relative  volatility,  341 
Sorel-Lewis,  122 
Sorel's,  118 
Thiele-Geddes,  326 
Minimum  boiling  azeotrope,  20 
Minimum  number  of  plates,  128,  174, 

243,  270,  292,  321,  376,  384 
Minimum  reflux  ratio,  128,  147,  170, 
172,  176,  253,  258,  272,  293,  321, 
379 

equation  for,  129,  253,  258,  294 
Mixing,  heat  of,  143 


Mixtures,  azeotropic,  21,  192 

completely  miscible,  26 

partially  miscible,  14,  21 
Modified  latent  heat  of  vaporization 

method,  158,  276 
Mol  fraction,  definition  of,  16 
Molecular  distillation,  395,  397 

thermal  efficiency  of,  401 
Multicomponent    mixtures,     alternate 
design  methods  for,  336 

batch  distillation,  383 

equation  for  minimum  reflux  ratio, 
249,  255,  258 

equation  for  total  reflux,  243 

feed-plate  location,  241,  248 

rectification  of,  214 
Multicomponent  systems,  72 
Multieffect  systems,  168 
Multilayer  systems,  14 
Multitower  systems,  168 


N 


Nitric  acid,  concentration  of,  296 
Number  of  transfer  units  (N.T.U.),  188 

0 

Open  steam,  134 

Operating  lines,  119,  192,  214 

intersection  of,  125 
Optimum  reflux  ratio,  129,  131 

example,  130 


Packed  towers,  183 

allowable  gas  and  liquid  velocities, 
439 

design  of,  184,  466 

efficiency  of,  464 

example,  466 

flooding  velocity,  440 

pressure  drop  in,  437 
Partial  condensation,  115 

equilibrium  of,  116 

differential  of,  116 
Partial  condenser,  116 
Partial  pressure,  26 


INDEX 


491 


jrtially  miscible  systems,  14,  21,  88, 
197 

rforated  plate,  431 

ase  rule,  16,  81,  216 

enol-water  fractionation,  197 

,te  efficiency,  105,  445 

jorrelations,  457 

equation  for,  462,  463 

a^ample,  463 

ixperimental  data,  455 

tie-design,  example,  433 

f}te  dumping,  419 

j|te  layout,  431 

ite  spacing,  426 

ate  stability,  418 

tnchon-Savarit  method,  146 

feed-plate  location,  148 

minimum  reflux  ratio,  147 

side  streams,  149 

total  reflux,  147 

iynting's  rule,  34,  48 

esentation  of  vapor-liquid  equilib- 
rium data,  16 

essure  drop,  for  bubble-cap  plate,  396 

due  to  liquid  head,  408 

in  packed  towers,  433 

through  risers  and  caps,  405 

through  slots,  405 

tor  vapor  flow,  404 

essure,  partial,  26 

reduced,  43 

sudo-azeotrope,  21 

eudo  mol  fraction,  90 

eudo critical  pressure,  37 

eudocritical  temperature,  37 

?T  relations,  36 


R 

Cult's  law,  26,  30,  46,  64,  93,  220 

•Dalton's  and,  28,  44 

deviation  from,  27 

'diagram  for,  27 

ate  of  evaporation,  394 

relative,  395 

ayleigh's  equation,  108 

^boilers,  471 

ectification,  104 

of  binary  mixture,  118 


Rectification,  with  chemical  reaction, 
361 

of  complex  mixture,  325 

degrees  of  freedom  in,  216 

of  mulibicomponent  mixture,  214 
Reduced  pressure,  43 
Reduced  relative  volatility  method,  341 
Reduced  temperature,  43 
Reflux  ratio,  minimum,  128 

optimum,  129 

total,  129 

Relative  rate  of  evaporation,  395 
Relative  volatility,  30,  395 

modification  of,  203 

use  of,  in  multicomponent  calcula- 
tions, 232 
Retrograde  phenomena,  81 

condensation,  81 


8 


Scatchard  equation,  62 
Simple  distillation,  107 
Solution  deviations,  46,  65 
Sorel-Lewis  method,  122 
Sorel's  method,  118 
Steam  distillation,  111,  135,  396 
Systems,  acetic  acid-water,  109 

acetone-carbon  disulfide,  19 

acetone-chloroform,  19,  66 

ammonia-water,  152 

aniline-water,  202 

benzene-n-propanol,  51 

benzene-toluene,  19 

benzene-water,  89,  202 

butane-hexane,  80 

carbon  dioxide-sulfur  dioxide,  82 

carbon  tetrachloride-carbon  disulfide, 

17 
temperature-composition  diagram, 

16 
yt  x  curve,  18 

ethanol-benzene-water,  314 

ethanol-isopropanol-water,  303 

ether-water,  90 

ethyl  alcohol-water,  46,  52,  163,  165, 
205-207 

gasoline,  261 

isobutahol-water,  19,  202 


492 


FRACTIONAL  DISTILLATION 


Systems,  isobutylene-propane,  44 
isopropyl  alcohol-water,  196 
methanol-water,  130 
methyl  ethyl  ketone-water,  203 
raulticomponent,  72 
multieffect,  168 
multilayer,  14 
multitower,  168 
nitric  acid-water,  297 
oleic-stearic  acids,  397 
phenol-o-cresol,  30 
phenol-water,  94,  198 
split-tower,  387 
toluene-water,  86 
two-tower,  202 


Tar  acid  fractionation,  236 
Temperature-composition  diagram,  16 

for  acetone-carbon  disulfide,  19 

for  acetone-chloroform,  19 

for  benzene-toluene,  19 

for  carbon  tetrachloride-carbon  disul- 
fide, 17 

for  isobutanol-water,  19 
Theoretical  plate,  119, 122 

height  equivalent  to,  187 
Thermal  efficiency,  methods  of  increas- 
ing, 167 

in  molecular  distillation,  401 
Thermodynamic  relations,  32 
Thiele  and  Geddes  method,  336 
Total  reflux,  128,  147,  243,  270,  292, 
321,  376,  384 

analytical  equation  for,  174 
Transfer  unit,  height  of,  188 

number  of,  188 

Transients  in  fractionating  column,  476 
True  boiling-point  distillation,  325 
Two-tower  system,  202 


U 


Unequal  molal  overflow,  138 


Vacuum  distillation,  113, 236, 393, 396 
Van  Laar  equation,  50,  56-59,  63,  98 

for  azeotropic  distillation,  286 

evaluation  of,  62 

example  of  use,  65 

for  extractive  distillation,  286 

for  multicomponent  mixtures,  73 
Van  der  Waals  equation,  57 
Vapor,  distribution  of,  418 

enthalpy  of,  140 

velocity  allowable,  430 
Vapor-liquid  composition  diagram,  16 
Vapor-liquid  equilibria,  3,  18 

calculation  of,  26 

data  on,  22-24 

determination  of,  3 

experimental  determination  of,  3 

presentation  of,  16 
Vapor  phase,  34 
Vapor  pressure,  of  ethanol,  51 

of  water,  51 

Vapor  recompression  system,  170 
Vapor  reuse  system,  170 
Vaporization,  heat  of,  143 
Volatility,  29 

abnormal,  29 

normal,  29 

relative,  30 

W 

Weirs,  409 

Wetted-wall  tower,  116 

Work  for  separation,  of  binary  mix* 

tures,  162 
of  ethanol-water,  163 


y,x  curve,  21 
for  CC14  and  CSa,  18 
ether-water,  90-92 
normal,  20 
propane-isobutylene,  44