Skip to main content

Full text of "Ocean fishery management : discussions and research / Adam A. Sokoloski, editor"

See other formats


(L5  ,     \  nmRs  csz£L(i~ 


NOAA  TR  NMFS  CIRC  371 


A  UNITED  STATES 
DEPARTMENT  OF 
COMMERCE 
PUBLICATION 


NOAA  Technical  Report  NMFS  CIRC-371 


U.S.   DEPARTMENT  OF  COMMERCE 

National  Oceanic  and  Atmospheric  Administration 

National   Marine  Fisheries  Service 


Ocean  Fishery  Management: 
Discussions  and  Research 


ADAM  A.  S0K0L0SKI  (Editor) 


SEATTLE,  WA 
April  1973 


NOAA  TECHNICAL  REPORTS 


National  Marine  Fisheries  Service,  Circulars 


The  major  responsibilities  of  the  National  Marine  Fisheries  Service  (NMFS)  are  to  monitor  and  assess  the 
abundance  and  geographic  distribution  of  fishery  resources,  to  understand  and  predict  fluctuations  in  the  quan- 
tity and  distribution  of  these  resources,  and  to  establish  levels  for  optimum  use  of  the  resources.  NMFS  is  also 
charged  with  the  development  and  implementation  of  policies  for  managing  national  fishing  grounds,  develop- 
ment and  enforcement  of  domestic  fisheries  regulations,  surveillance  of  foreign  fishing  off  United  States  coastal 
waters,  and  the  development  and  enforcement  of  international  fishery  agreements  and  policies.  NMFS  also 
assists  the  fishing  industry  through  marketing  service  and  economic  analysis  programs,  and  mortgage  insurance 
and  vessel  construction  subsidies.     It  collects,  analyses,  and  publishes  statistics  on  various  phases  of  the  industry. 

The  NOAA  Technical  Report  NMFS  CIRC  series  continues  a  series  that  has  been  in  existence  since  1941.  The 
Circulars  are  technical  publications  of  general  interest  intended  to  aid  conservation  and  management.  Publica- 
tions that  review  in  considerable  detail  and  at  a  high  technical  level  certain  broad  areas  of  research  appear  in 
this  series.  Technical  papers  originating  in  economics  studies  and  from  management  investigations  appear  in 
the  Circular  series. 

XOAA  Technical  Reports  NMFS  CIRC  are  available  free  in  limited  numbers  to  governmental  agencies,  both 
Federal  and  State.  They  are  also  available  in  exchange  for  other  scientific  and  technical  publications  in  the  ma- 
rine sciences.  Individual  copies  may  be  obtained  (unless  otherwise  noted)  from  NOAA  Publications  Section,  Rock- 
ville,  Md.  20852.     Recent  Circulars  are: 


315.  Synopsis  of  biological  data  on  the  chum  salmon, 
Oncorhynchus  keta  (Walbaum)  1792.  By  Rich- 
ard G.  Bakkala.  March  1970,  iii  +  89  pp.,  15 
figs.,  51  tables. 

319.  Bureau  of  Commercial  Fisheries  Great  Lakes 
Fishery  Laboratory,  Ann  Arbor,  Michigan.  By 
Bureau  of  Commercial  Fisheries.  March  1970, 
8  pp.,  7  figs. 

330.  EASTROPAC  Atlas:     Vols.  4,  2.     Catalog  No. 

1  49.4:330/ (vol.)  11  vols.  ($4.75  each).  Avail- 
able from  the  Superintendent  of  Documents, 
Washington,  D.C.  20402. 

331.  Guidelines  for  the  processing  of  hot-smoked  chub. 
By  H.  L.  Seagran,  J.  T.  Graikoski,  and  J.  A. 
Emerson.     January  1970,   iv  +  23  pp.,   8   figs., 

2  tables. 

332.  Pacific  hake.  (12  articles  by  20  authors.)  March 
1970,  iii   +  152  pp.,  72  figs.,  47  tables. 

333.  Recommended  practices  for  vessel  sanitation  and 
fish  handling.  By  Edgar  W.  Bowman  and  Alfred 
Larsen.    March  1970,  iv  +  27  pp.,  6  figs. 

335.  Progress  report  of  the  Bureau  of  Commercial 
Fisheries  Center  for  Estuarine  and  Menhaden 
Research,  Pesticide  Field  Station,  Gulf  Breeze, 
Fla.,  fiscal  year  1969.  By  the  Laboratory  staff. 
August  1970,  iii   +   33  pp.,  29  figs.,  12  tables. 

336.  The  northern  fur  seal.  By  Ralph  C.  Baker,  Ford 
Wilke,  and  C.  Howard  Baltzo.  April  1970,  iii  + 
19  pp.,  13  figs. 

337.  Program  of  Division  of  Economic  Research, 
Bureau   of    Commerecial    Fisheries,    fiscal    year 

1969.  By  Division  of  Economic  Research.    April 

1970,  iii  +  29  pp.,  12  figs.,  7  tables. 


338.  Bureau  of  Commercial  Fisheries  Biological  Lab- 
oratory, Auke  Bay,  Alaska.  By  Bureau  of  Com- 
mercial Fisheries.    June  1970,  8  pp.,  6  figs. 

339.  Salmon  research  at  Ice  Harbor  Dam.  By  Wesley 
J.  Ebel.     April  1970,  6  pp.,  4  figs. 

340.  Bureau  of  Commercial  Fisheries  Technological 
Laboratory,  Gloucester,  Massachusetts.  By  Bu- 
reau of  Commercial  Fisheries.  June  1970,  8  pp., 
8  figs. 

341.  Report  of  the  Bureau  of  Commercial  Fisheries 
Biological  Laboratory,  Beaufort,  N.C.,  for  the 
fiscal  year  ending  June  30,  1968.  By  the  Lab- 
oratory staff.  August  1970,  iii  -f  24  pp.,  11  figs., 
16  tables. 

342.  Report  of  the  Bureau  of  Commercial  Fisheries 
Biological  Laboratory,  St.  Petersburg  Beach, 
Florida,  fiscal  year  1969.  By  the  Laboratory  staff. 
August  1970,  iii  +  22  pp.,  20  figs.,  8  tables. 

343.  Report  of  the  Bureau  of  Commercial  Fisheries 
Biological  Laboratory,  Galveston,  Texas,  fiscal 
year  1969.  By  the  Laboratory  staff.  August 
1970,  iii  +  39  pp.,  28  figs.,  9  tables. 

344.  Bureau  of  Commercial  Fisheries  Tropical  Atlan- 
tic Biological  Laboratory  progress  in  research 
1965-69,  Miami,  Florida.  By  Ann  Weeks.  Oc- 
tober 1970,  iv  +  65  pp.,  53  figs. 

346.  Sportsman's  guide  to  handling,  smoking,  and  pre- 
serving Great  Lakes  coho  salmon.  By  Shearon 
Dudley,  J.  T.  Graikoski,  H.  L.  Seagran,  and  Paul 
M.  Earl.     September  1970,  iii  +  28  pp.,  15  figs. 

347.  Synopsis  of  biological  data  on  Pacific  ocean  perch, 
Sebastodes  alutus.  By  Richard  L.  Major  and 
Herbert  H.  Shippen.  December  1970,  iii  +  38 
pp.,  31  figs.,  11  tables. 


Continued  on  inside  back  cover. 


ATMOSAl, 


Iff  NT  of  c 


U.S.  DEPARTMENT  OF  COMMERCE 
Frederick  B.  Dent,  Secretary 

NATIONAL  OCEANIC  AND  ATMOSPHERIC  ADMINISTRATION 
Robert  M.  White,  Administrator 

NATIONAL  MARINE  FISHERIES  SERVICE 
Philip  M.  Roedel,  Director 


NOAA  Technical  Report  NMFS  CIRC-371 

Ocean  Fishery  Management: 
Discussions  and  Research 


ADAM  A.  S0K0L0SKI  (Editor) 


9- 
o 

u 


o 

a 

0) 
Q 


Proceedings  of  a  workshop 
sponsored  by  the  Division 
of  Economic  Research,  National 
Marine  Fisheries  Service, 
November  5-6,  1970 

This  report  contains  a  group 
of  independent  research  studies 
published  on  their  own  merits. 
They  do  not  necessarily  reflect 
the  policies  or  intentions  of 
the  National  Marine  Fisheries 
Service. 


SEATTLE,  WA 
April  1973 


PARTICIPANTS 


Paul    Adam,    Organization    for    Economic    Cooperation 

and  Development 
Harold  B.   Allen,   National   Marine  Fisheries   Service 
Frank  M.  Anderson,  Oregon  State  University 
Frederick  W.  Bell,  National  Marine  Fisheries  Service 
Daniel  W.  Bromley,  University  of  Wisconsin 
Ernest  W.  Carlson,  National  Marine  Fisheries  Service 
Donald  P.  Cleary,  National  Marine  Fisheries  Service 
Francis  T.  Christy,  Jr.,  Resources  for  the  Future 
James  A.  Crutchfield,  University  of  Washington 
John  P.  Doll,  University  of  Missouri 
Richard   F.    Fullenbaum,    National    Marine   Fisheries 

Service 
John  M.  Gates,  University  of  Rhode  Island 
Thomas   Geer,    International    Bank   for   Reconstruction 

and  Development 
William  G.  Gordon,  National  Marine  Fisheries  Service 
Loren  Grant,  Canadian  Department  of  Fisheries  and 

Forestry 
John  E.  Greenfield,  National  Marine  Fisheries  Service 
Murray  L.  Hayes,  National   Marine  Fisheries  Service 
Andreas  A.  Holmsen,  University  of  Rhode  Island 
Paul  Hooker,  University  of  Florida 
A.  M.  HuQ,  University  of  Maine 
Harvey    M.    Hutchings.     National    Marine    Fisheries 

Service 
A.  D.  Insul,  British  White  Fish  Authority 
Joshua  John,  Canadian   Department  of  Fisheries  and 

Forestry 
Milton  G.  Johnson,  National  Oceanic  and  Atmospheric 

Administration 
Edward  Kane,  Boston  College 
E.  A.  Keen,  San  Diego  State  College 
Richard    K.     Kinoshita,    National    Marine    Fisheries 

Service 


Jukka  A.  Kolhonen,  National  Marine  Fisheries  Service 
Richard  J.  Marasco,  University  of  Maryland 
Bruce  Mattox,  University  of  Rhode  Island 
Morton  M.  Miller,  National  Marine  Fisheries  Service 
Paul  Mlotok,  University  of  Rhode  Island 
Darrel  A.  Nash,  National  Marine  Fisheries  Service 
Bruno  G.  Noetzel,  National  Marine  Fisheries  Service 
Virgil  J.  Norton,  University  of  Rhode  Island 
Frederick  L.  Olson,  National  Marine  Fisheries  Service 
Erwin  Penn,  National  Marine  Fisheries  Service 
Giulio  Pontecorvo,  Columbia  University 
Frederick  Prochaska,  University  of  Florida 
R.  Bruce  Rettig,  Oregon  State  University 
Jack  Rich,  Oregon  State  University 
Jack  A.  Richards,  Kansas  State  University 
Jon  C.  Rittgers,  University  of  Rhode  Island 
Richard  Roberts,  Fisheries  Service  of  Canada 
Michael  A.  Robinson,  Food  and  Agriculture  Organiza- 
tion of  the  United  Nations 
Edilberto  L.  Segura,  Columbia  University 
Frederick  J.  Smith,  Oregon  State  University 
Adam  A.  Sokoloski,  National  Marine  Fisheries  Service 
Miller  Spangler,  National  Planning  Association 
Joe  B.  Stevens,  Oregon  State  University 
David  A.  Storey,  University  of  Massachusetts 
John  K.  Sullivan,  National  Marine  Fisheries  Service 
William  M.  Terry,  National  Marine  Fisheries  Service 
B.  G.  Thompson,  National  Marine  Fisheries  Service 
Russell  G.  Thompson,  National  Water  Commission 
Lawrence  W.  Van  Meir,  National  Canners  Association 
John   Vondruska,    National    Marine   Fisheries   Service 
Hoyt  A.  Wheeland,  National  Marine  Fisheries  Service 
Donald  R.  Whitaker,  National  Marine  Fisheries  Service 
Frederick  E.  A.  Wood,  Fisheries  Service  of  Canada 
Robert  Wilson,  Texas  A  &  M  University 


PREFACE 

Until  recent  years  only  biological  or  technical  aspects  of  fisheries  con- 
servation have  advanced  beyond  esoteric  professional  journals  or  smoke- 
filled  back  rooms  to  be  given  serious  consideration  when  formulating  work- 
ing management  programs.  In  recent  years  the  social  sciences,  especially 
economics  with  its  emphasis  on  rational  management,  have  gained  some 
respectability  beyond  mere  conceptual  discussion. 

With  the  mounting  urgency  of  fishery  management  problems  serving 
as  a  catalyst,  the  National  Marine  Fisheries  Service  has  multiplied  its 
research  in  this  area,  aided  in  part  by  the  rapidly  growing  Sea  Grant  pro- 
gram formerly  within  the  National  Science  Foundation  and  now  incorpor- 
ated within  the  National  Oceanic  and  Atmospheric  Administration. 

Within  the  past  two  years  much  progress  has  been  made.  To  aid  in 
assimilating  these  results  and  to  provide  some  sense  of  a  proper  future 
direction  for  both  research  and  the  design  of  management  programs,  the 
National  Marine  Fisheries  Service  convened  a  Workshop  on  November 
5  and  6,  1970.  Invited  were  virtually  all  known  researchers  in  Fishery 
Economics  throughout  the  world,  many  administrators,  and  researchers 
in  related  disciplines. 

What  follows  in  this  circular  are  the  papers  presented  at  this  workshop, 
with  an  introduction  which  makes  a  first  attempt  at  distilling  the  combin- 
ed impact  of  these  papers. 

As  editor  I  wish  to  thank  all  the  authors  for  their  diligent  cooperation. 
The  services  of  the  secretarial  staff  of  the  Division  of  Economic  Research, 
especially  Miss  Carol  Reese,  are  gratefully  acknowledged.  The  generous 
support  of  the  many  institutions  that  absorbed  the  financial  burden  of 
travel  from  distant  geographic  regions  was  necessary  for  the  ultimate 
success  of  this  workshop. 

The  National  Marine  Fisheries  Service  sponsors  the  publication  of 
these  papers,  as  it  sponsored  the  workshop  itself,  to  crystallize  the  issues 
relating  to  fishing  management  and  to  stimulate  further  debate.  As  such 
the  papers  present  the  views  of  the  individual  authors  and  none  of  the 
material  contained  herein  should  be  construed  as  reflecting  official  policy 
statements  of  the  National  Marine  Fisheries  Service. 

Adam  A.  Sokoloski 


in 


CONTENTS 

Page 
INTRODUCTION    1 

The  Status  of  Fisheries  Management  Research:  An  Overview. 

Adam  A.  Sokoloski 1 

ISSUES  IN  FISHERY  MANAGEMENT 7 

Problems  in  Implementing  New  Fishery  Management  Programs. 
Lawrence  W.  Van  Meir 9 

On  the  Utility  of  Bioeconomic  Models  for  Fisheries  Management. 

Giulio  Pontecorvo 12 

Multiple  Objectives  for  Marine  Resource  Management. 

R.  Bruce  Rettig 23 

Economic,  Political,  and  Social  Barriers  to  Efficiency  in 

Selected  Pacific  Coast  Fisheries.  James  A.  Crutchfield 28 

PRODUCTION  FUNCTIONS  AND  BIOECONOMIC  MODELS: 
RESEARCH  IMPLICATIONS   39 

Cross  Section  Production  Functions  for  North  Atlantic  Groundfish  and 
Tropical  Tuna  Seine  Fisheries.  Ernest  W.  Carlson 42 

Optimal  Fishing  Effort  in  the  Peruvian  Anchoveta  Fishery. 

Edilberto  L.  Segura 57 

Natural  Resources  and  External  Economies:  Regulation  of  the 

Pacific  Halibut  Fishery.  Jack  Rich 65 

Production  from  the  Sea.  Frederick  W.  Bell,  Ernest  W.  Carlson, 

Frederick  V.  Waugh 72 

Some  Suggestions  for  the  Development  of  a  Bioeconomic  Theory  of 

the  Fishery.  Russell  G.  Thompson 92 

Practical  Problems  of  Constructing  Bioeconomic  Models  for  Fishery 
Management.  Paul  Adam 96 


(Continued) 


ISSUES  RELATED  TO  FISHERY  MANAGEMENT: 

RESEARCH  RESULTS  : 104 

Management  of  the  Peruvian  Anchoveta  Resource. 

Andreas  A.  Holmsen 106 

A  Stochastic  Investment  Model  for  a  Survival  Conscious  Fishing 

Firm.  Russell  G.  Thompson,  Richard  W.  Callen,  Lawrence  C.  Wolken  ...  112 

Simulation  Experiments  to  Evaluate  Alternative  Hunting  Strategies 
for  a  Deer  Population.  F.  M.  Anderson,  G.  E.  Connolly,  A.  H.  Halter, 
W.  M.  Longhurst 121 

Augmentation  of  Salmon  Stocks  through  Artificial  Propagation: 

Methods  and  Implications.  Joe  B.  Stevens  and  Bruce  W.  Mattox 133 

Limited  Entry:  The  Case  of  the  Japanese  Tuna  Fishery.  E.  A.  Keen 146 

A  Study  of  the  Socioeconomic  Impact  of  Changes  in  the  Harvesting 

Labor  Force  in  the  Maine  Lobster  Industry.  A.  M.  Huq 159 


VI 


INTRODUCTION 

The  Status  of  Fisheries  Management  Research 
An  Overview 

Adam  A.  Sokoloski1 


All  disclaimers  to  the  contrary,  there  is  one 
research  area  near  and  dear  to  the  hearts  of 
virtually  all  economists  conducting  research 
on  marine  resources:  measuring  the  gap  be- 
tween the  "optimum"  management  solution  for 
a  given  fishery  and  current  management  arrange- 
ments. This  is  not  to  say  that  this  gap  has  ever 
been  successfully  measured  for  a  fishery. 

In  recent  years  some  first  approximations 
have  been  made,  however.  These  have  been 
reasonably  consistent  with  a  body  of  economic 
theory  which  has  existed  in  one  form  or  an- 
other for  several  years.  This  theory  is  the 
original  source  of  suggestions  that  the  gap 
existed,  as  casual  observation  of  practice  re- 
vealed   inconsistencies    with    "proper"    theory. 

Initial  empirical  works  unearthed  several 
critical  components  which  are  currently  com- 
plicating the  issue.  These  are  both  empirical 
and  conceptual  in  nature  and  multidisciplin- 
ary  in  scope.  These  may  be  listed  as  follows: 

1.  Existing  yield  functions  need  to  be  ex- 
panded and  alternative  functions  need 
to  be  specified,  both  with  respect  to  such 
factors  as  diminishing  returns  (success 
probabilities  for  effort  on  a  fixed  bio- 
mass)  and  multispecies  interrelation- 
ships. 

2.  The  appropriate  emphasis  for  economics 
and  biology  in  bioeconomic  models. 

3.  The  correct  theoretical  and  empirical 
components  of  effort  series   are   needed 


1  Formerly  of  Division  of  Economic  Research, 
National  Marine  Fisheries  Service;  present  address, 
Environmental  Protection  Agency,  Division  of  Water 
Quality  Standards,  Arlington,  VA  22202. 


to  construct  indices  of  fishing  power  as 
utilized  in  management  programs. 

4.  More  effort  is  needed  in  the  design  of 
"correct"  operational  management  plans: 
the  choice  between  variations  of  licens- 
ing, quota,  auction  and/or  leasing 
schemes. 

5.  A  resolution  of  the  choice  between  long 
run  versus  short  run  "optimal"  solutions. 

6.  An  evaluation  of  the  appropriateness  of 
directly  applying  theoretical  models  to 
fisheries  for  the  purpose  of  deriving  im- 
plied net  gains  from  the  practical  appli- 
cation of  identical  working  models. 

7.  The  role  of  social  transfer  costs  in  the 
evaluation  of  benefits  from  new  manage- 
ment programs. 

8.  The  desirability  of  an  incentive  (pull) 
approach  versus  a  limited  entry  (push) 
form  of  management  program. 

9.  The  place  of  jurisdictional  consideration 
in  program  design  and  operation. 

10.  The  desirability  of  a  multidisciplinary 
objective  simulation  approach  to  the 
measurement  of  management  ramifica- 
tions as  contrasted  to  simultaneous 
equations  with  maximization  and  other 
limiting  assumptions. 

11.  The  role  of  artificial  propagation  in  the 
design  of  total  management  plans. 

12.  The  role  of  competing  uses  for  the  re- 
source base. 

Virtually  all  of  these  items  reflect  the  fact 
that  as  economists  begin  to  penetrate  the  sur- 
face of  the  management  issue  they  gain  a 
greater  appreciation  of  the  vital  role  to  be 
played    by    the    physical    scientist,    usually    a 


biologist  who  has  become  a  population  dy- 
namics expert.  This  involves  more  than  just 
using  the  output  of  the  population  dynamics 
expert;  it  entails  understanding  the  intricacies 
of  this  work  so  that  it  won't  be  misused. 

Here  is  where  the  first  problems  arise.  Once 
familiar  with  population  dynamics  models  the 
economist  falls  prey  to  the  temptation  to  alter 
components  which  may  not  be  ideally  suited  to 
his  needs.  What  results  is  two  versions  of 
population  dynamics  with  one  being  the  result 
of  both  explicit  and  implicit  imperfections  in 
the  other. 

From  this  point  several  ramifications  may 
develop,  depending  on  how  far  each  conceptual 
base  may  have  been  developed  toward  an  actual 
working  management  program.  If  this  has  oc- 
curred original  differences  in  population  dy- 
namics models  will  have  been  magnified.  These 
resultant  differences  generate  a  debate,  and  a 
portion  of  this  debate,  as  currently  stated,  is 
contained  in  the  following  papers.  To  amplify 
let  me  refer  in  greater  detail  to  the  twelve 
points  mentioned  above. 

(1)  The  Need  for  New  Yield  Functions:  The 
biologist's  yield  function  is  the  analogue  of 
the  economist's  production  function.  Produc- 
tion economics  focuses  upon  the  allocation 
of  inputs  to  achieve  production  goals  designat- 
ed as  optimum,  this  proper  allocation  being  the 
most  efficient  (least  cost)  combination  of  these 
inputs.  Partial  derivatives,  giving  the  incre- 
mental contribution  of  each  unit  of  a  particu- 
lar input,  may  be  used  to  construct  efficiency 
indices  roughly  equivalent  to  the  biologist's 
measures  of  the  fishing  power  of  a  vessel. 

These  derivatives  are  obtained  from  general 
form  equations  of  a  linear,  Cobb-Douglas  (con- 
stant elasticity  of  substitution  equal  to  one) 
or  C.E.S.  (any  constant  elasticity  of  substitu- 
tion) type.  Contained  within  these  general  types 
are  certain  assumptions  concerning  constant, 
increasing,  or  decreasing  returns  (output)  from 
increasing  increments  of  a  particular  input. 

Critical  here  is  an  appreciation  of  the  fact 
that  these  are  fundamental  calculations  which 
would  be  carried  out  whether  or  not  any  re- 
lated biological  work  existed.  When  this  work 
does  exist  it  serves  as  a  reference  point  to  the 
economist  as  he  proceeds  systematically 
through  a  series  of  steps  dictated  by  the  classi- 


cal scientific  method  which  has  evolved  for 
his  profession. 

When,  therefore,  an  economist  specifies  a 
function  implying  diminishing  returns  to  ad- 
ditional inputs,  we  have  the  potential  for  debate 
when  the  biologist  has  diminishing  returns 
due  to  population  dynamics  but  constant  re- 
turns from  a  fixed  biomass.  These  two  differ- 
ing approaches  will  lead  to  different  evalua- 
tions of  the  historical  effort  being  exerted  on 
a  fishery,  to  different  estimations  of  the  actual 
yield  curve,  to  different  calculations  of  MSY 
(maximum  sustainable  yield)  and  then  to  dif- 
ferent management  solutions. 

The  issue  becomes  further  complicated  when 
many  species  intermix  and  then  must  be  con- 
sidered simultaneously  when  designing  and 
operating  a  management  program.  The  eco- 
nomic portion  of  this  analysis  is  actually  more 
readily  solved  in  this  case  via  a  standard 
analysis  of  the  joint  product  case,  whereas  the 
biological  literature  still  carries  a  debate  con- 
cerning the  proper  use  of  Beverton-Holt  dy- 
namic pool  models  as  opposed  to  the  Schaefer 
logistic  approach.  This  issue  is  becoming  more 
critical  as  the  trend  in  the  technological  capa- 
bility of  harvesting  units  is  leading  toward 
some  point  in  the  future  where  the  flexibility 
and  maneuverability  of  these  units  will  make 
all  management  considerations  multispecies 
to  correctly  reflect  actual  harvesting  practices. 

(2  &  3)  Economics  and  Biology  in  Measures 
of  Fishing  Power:  For  management  purposes 
what  is  the  appropriate  emphasis  of  economics 
and  biology  in  bioeconomic  models?  One  ex- 
treme suggests  that  it  is  necessary  to  under- 
stand the  complete  microdynamics  of  all  stages 
of  the  food  chain,  an  ecological  approach,  and 
all  forces  that  act  upon  these  stages,  to  proper- 
ly specify  the  results  of  variation  in  fishing 
effort  and,  therefore,  to  suggest  the  optimum 
dimensions  of  that  effort.  This  would  confine 
economists  to  a  role  of  evaluating  the  economic 
costs  and  benefits  of  the  program  suggested 
by  this  detailed  formulation. 

The  opposite  extreme  finds  the  economist 
placing  the  fisherman  in  an  active  role,  where 
he  responds  to  various  market  incentives, 
these  responses  subsequently  becoming  an  in- 
tegral step  in  determining  variations  in  fishing 
effort    and     resultant    success.     Some    would 


suggest  that  prices,  quantities  landed,  and  a 
statistically  acceptable  production  function 
are  all  that  is  needed  to  derive  the  functional 
relationship  necessary  for  management,  as- 
suming that  this  production  function  can  be 
used  to  determine  the  continuing  relationship 
between  effort  and  landings. 

This  last  phrase  is  important,  for  it  may 
well  be  that  a  final  decision  in  the  allocation 
of  research  and  management  resources  will 
depend  upon  the  spin-off,  or  secondary  benefits 
from  certain  research  endeavors  above  and 
beyond  their  direct  contribution  to  manage- 
ment. This  would,  of  course,  be  especially  true 
with  regard  to  the  extreme  of  the  broad-based 
ecological  approach,  with  both  long  run  and 
short  run  considerations,  as  suggested  in  some 
biological  circles.  The  recent  reorganization 
of  certain  agency  functions  within  the  National 
Oceanic  and  Atmospheric  Administration 
may  have  some  bearing  here. 

(4)  Using  Research  Results  to  Design  Oper- 
ational Schemes:  Existing  schemes  which  may 
be  actually  classified  as  direct  measures  to 
limit  entry  have  established  certain  precedents. 
Canadian  programs  in  Atlantic  lobsters  and 
Pacific  salmon  have  emphasized  licenses  for 
principal  capital  inputs,  a  limited  entry  pro- 
gram. The  Inter-American  Tropical  Tuna 
Commission  (IATTC)  has  utilized  quotas  and 
area  restrictions  while  International  Commis- 
sion for  the  Northwest  Atlantic  Fisheries 
(ICNAF)  has  utilized  mesh  size  and  area  re- 
strictions and  recently  seasonal  closures  for 
overfished  species.  The  Union  of  South  Africa 
regulates  via  licenses  issued  through  proces- 
sing plants  which  allocate  these  among  vessels. 

The  debate  concerning  these  existing  types 
and  many  other  hypothesized  forms  may  be 
divided  into  two  subject  areas,  one  concerning 
whether  the  plan  will  actually  lead  to  an  al- 
location of  resources  which  approaches  some 
predetermined  optimum  and  the  other  whether 
the  plan  is  operationally  realistic,  which  may 
invoke  social  and  political  considerations  as 
well  as  those  of  biology,  technology,  and  eco- 
nomics. Mr.  William  Terry,  in  his  opening 
statements  to  the  participants  of  this  work- 
shop, emphasized  the  urgent  need  to  begin 
evaluating  the  proper  mix  of  all  possible  inputs 
into   fisheries   management.    He   asserted   that 


we  must  begin  now  to  define  the  components 
of  the  interface,  looking  beyond  the  immediate 
problems  of  each  discipline.  Consistent  with 
this  he  suggested  the  possible  need  to  develop 
new,  broader  objectives  of  fishing  management. 

Presently  discussed  mechanisms  have  two 
principal  components,  a  way  of  limiting  entry 
into  fisheries  and  a  means  of  allocating  the 
quasi-property  rights  which  result.  Some 
mechanisms  perform  these  functions  simul- 
taneously, such  as  an  auction  system,  whereas 
others,  such  as  a  licensing  system,  require  the 
administrator  to  make  some  judgement  as  to 
the  number  of  licenses  as  well  as  how  they 
shall  be  allocated.  Relating  to  some  historical 
experience  in  the  oyster  fishery,  there  are 
many  unexplored  questions  regarding  the  ap- 
plicability of  leasing  schemes  for  sessile  re- 
sources. 

The  message  here  is  clear.  We  have  devoted 
considerable  effort  to  developing  sophisticated 
conceptual  constructs  for  fisheries  manage- 
ment. Regarding  actual  operational  alterna- 
tives we  have  confined  our  efforts  to  informal 
and  often  exclusively  internal  debates.  Re- 
sponsible researchers  must  soon  assume  the 
task  of  a  thorough  evaluation  of  the  many 
suggested  working  plans.  This  evaluation 
must  be  subsequently  exposed  to  discussion 
via  the  professional  journals  so  that  all  the 
preceding  work  can  truly  be  productive. 

(5)  Long  Run  Versus  Short  Run  Solutions: 
The  fruits  of  economic  modeling  are  proposals 
involving  changes  in  the  quantities  of  labor 
and/or  capital  in  commercial  fishing,  often 
reductions  of  both  in  addition  to  increases 
in  the  capital/labor  ratio.  Capital  inputs  are 
usually  quite  fixed;  indeed,  this  may  be  true 
of  labor  inputs  also.  What  results  is  the  quite 
obvious  conclusion  that  achieving  these  opti- 
mum solutions  will  in  all  cases  involve  extend- 
ed time  periods.  What  then  is  the  preference 
ranking  for  the  many  plans  which  may  have 
to  be  initiated  in  the  interim? 

The  truthful  answer  to  these  questions  is 
that  we  really  don't  know.  We  have  not  fully 
designated  the  compromises  which  would  be 
necessary,  much  less  made  a  careful  evaluation 
of  which  would  be  preferable.  This  suggests 
two  immediate  tasks  to  be  undertaken  by 
the  economist. 


The  first  of  these  relates  to  the  fact  that 
the  responsible  administrator  will  not  wait 
for  the  perfect  solution  to  be  formulated  for 
each  time  horizon.  He  must  formulate  plans 
and  action  programs  on  a  continuing  basis. 
In  this  instance  the  economist  can  indicate 
those  steps  which  can  be  taken  which  will 
lead  toward  the  optimum  solution,  or  at  least 
toward  some  "better"  solution  in  the  tradition 
of  the  theories  of  second  best  as  discussed 
in  the  literature  on  welfare  economics. 

Simultaneously  the  economist  can  perform 
a  second  function,  which  would  be  to  construct 
detailed  interim  plans  and  test  and  evaluate 
these.  These  could  be  constructed  for  alterna- 
tive time  periods  and  based  upon  restrictions 
suggested  by  the  administrator  or  the  other 
disciplines  where  additional  intermediate  term 
planning  and  research  was  also  being  con- 
ducted. 

The  result  would  be  an  array  of  economic 
research  considering  time  horizons  from  the 
present  to  the  long  run  optimum  solution. 
With  such  an  array  it  would  be  easier  to  in- 
corporate the  interdisciplinary  (especially 
social)  aspects  of  the  overall  problem  as  sug- 
gested in  the  other  points  I  am  presenting 
here.  Most  critical  is  the  fact  that  immediate 
steps  are  necessary  if  there  is  to  be  a  fishery 
to  optimize  in  the  long  run. 

(6)  Theoretical  Versus  Working  Models: 
This  point  relates  to  the  previous  issue  and 
also  to  several  of  the  following.  Brieflly  the 
question  here  is  whether  theoretical  models, 
confined  solely  to  a  select  number  of  variables, 
and  seldom  involving  more  than  two  disci- 
plines, can  be  utilized  directly  for  generating 
a  stream  of  benefits  to  be  included  in  the  calcu- 
lation of  a  benefit/cost  ratio  for  a  particular 
management  program.  Some  have  argued  that 
there  is  not  sufficient  realism  in  theoretical 
models  for  these  to  be  applied  directly.  Con- 
versely, it  may  also  be  argued  that  many  of 
the  imperfections  of  this  approach  are  not  so 
much  inherent  in  the  theoretical  models  them- 
selves, but  rather  stem  from  the  use  of  com- 
plementary information  when  performing  B/C 
analysis.  Such  errors  may  be  found  principal- 
ly on  the  cost  side,  where  not  all  indirect  pro- 
gram costs  are  included,  especially  when 
these    costs    may    exceed    the    actual    flow    of 


benefits  in  the  short  run.  What  may  be  the 
most  significant  of  these  cost  components  is 
discussed  next. 

(7)  Social  Transfer  Costs:  Not  long  ago  it 
was  not  possible  to  discuss  limited  entry  ex- 
cept under  the  most  constrained  circumstances. 
Now,  with  the  development  of  more  forceful 
arguments  and  with  a  growing  urgency  in 
certain  fisheries,  limited  entry  plans  are  re- 
ceiving wider  consideration.  As  this  occurs 
the  operational  elements  of  alternative  plans 
are  being  formulated  and  new  questions  are 
resulting.  The  most  prominent  among  these 
relates  to  the  magnitude  of  the  social  transfer 
costs  which  may  result  from  either  the  direct 
or  indirect  reduction  of  the  fishing  labor 
force  in  a  fishery. 

If  displaced  labor  must  be  retrained  and 
relocated,  or  absorbed  on  the  welfare  rolls, 
it  may  be  wise  to  develop  programs  based 
exclusively  on  excluding  excessive  new  entry, 
with  input  balances  to  be  attained  via  attri- 
tion. This  is  tantamount  to  concluding  that 
the  short  run  solution  must  be  contrary  to 
the  suggested  long  run  optimum.  Neverthe- 
less, it  is  the  desired  means  of  achieving  the 
long  run  optimum.  Research  is  now  begin- 
ning on  this  issue  both  within  the  National 
Marine  Fisheries  Service  and  the  Office  of 
Sea  Grant  Programs.  The  results  will  play 
a  critical  part  in  determining  the  character 
of  future  management  plans. 

(8)  Encouraging  Exit  Versus  Limiting  Entry: 
Virtually  all  discussions  of  management  plans 
emphasize  licenses,  or  quotas,  or  some  form 
of  right  which  will  accrue  to  a  reduced  num- 
ber of  harvesting  units.  The  mechanics  of 
reducing  these  units  involve  some  form  of 
exclusion.  Seldom  has  a  plan  been  suggested, 
however,  which  emphasizes  a  program  where- 
by excess  inputs  would  be  attracted  away 
from  the  fishery  by  a  more  rewarding  altern- 
ative. 

To  my  knowledge  such  a  program  has  been 
attempted  once,  a  recent  attempt  to  divert 
excess  capacity  from  the  overfished  haddock 
resource  of  the  Northwest  Atlantic  to  the 
underutilized  pollock  resource.  As  a  limited 
short  term  program  it  met  with  only  limited 
success.    This   is   not   inconsistent   with   other 


similar  non-fishery  programs,  such  as  the  more 
substantial  effort  designed  for  Appalachia.  In 
all  instances,  unanticipated  attractions,  pic- 
turesquely described  as  "psychic  income,"  re- 
sulted in  a  greater  amount  of  labor  immobility 
than  original  calculations  suggested.  Program 
costs  had  to  be  adjusted  accordingly. 

Nevertheless,  if  transfer  costs,  as  discussed 
in  the  previous  point,  can  be  reduced  by  some 
increment  by  an  incentive  program  costing 
less  than  this  increment,  then  the  overall  costs 
of  the  total  management  program  may  be 
reduced  sufficiently  to  result  in  a  favorable 
B/C  ratio.  These  calculations  would  be  over 
and  above  the  more  favorable  political  re- 
sponse to  a  program  which  considered  these 
transfer  costs  as  opposed  to  one  which  did  not. 

The  problem  of  response  to  incentives  may 
be  reduced  in  multiple  species  fisheries  where 
we  wish  to  reduce  pressure  on  one  of  the 
species  and  this  is  technically  possible.  Hard- 
ships resulting  from  restrictions  on  the  king 
crab  resource  were  reduced  by  the  ability  of 
the  harvesting  units  to  adapt  to  alternative 
species.  Indeed,  New  Bedford  scallopers,  13 
vessels  in  all,  journeyed  to  Alaska  when  that 
resource  appeared  (somewhat  falsely)  more 
profitable  than  their  traditional  fishery.  If 
they  could  have  been  induced  to  leave  earlier 
then  perhaps  the  degree  of  depletion  in  the 
Atlantic  could  have  been  reduced. 

(9)  Jurisdictional  Issues:  Fisheries  research- 
ers interested  in  formulating  management 
plans  usually  focus  on  specific  fisheries  in 
their  entirety.  This  is  appropriate  for  every 
"discipline"  except  one,  the  area  of  legal- 
political  considerations.  Fish  do  not  respect 
jurisdictional  boundaries  and  this  has  long 
been  a  critical  operational  issue  in  fisheries 
management. 

Resolution  of  these  jurisdictional  issues  will 
involve  more  lead  time  than  biological  and 
economic  questions.  Developing  interstate  co- 
operative mechanisms  and  widely  accepted 
international  arrangements  which  will  be  polit- 
ically acceptable  while  incorporating  biologic- 
al, economic,  and  social  factors  will  be  a 
herculean  task,  witness  the  slow  progress  of 
developing  a  national  quota  system  in  ICNAF 
and  the  200-mile  dispute  with  countries 
bounding  the  yellowfin  tuna  fishing  areas. 


The  individual  disciplines  can  contribute  to 
solving  this  problem  by  orienting  their  work 
so  that"  the  trade-offs  between  alternative  juris- 
dictional arrangements  can  be  readily  assessed 
in  each  disciplinary  dimension.  As  the  U.S. 
develops  new  coastal  zone  and  contiguous 
zone  legislation  and  as  all  nations  prepare 
for  another  Law-of-the-Sea  conference  it  be- 
comes increasingly  necessary  that  these  trade- 
offs be  specified  in  the  near  future. 

(10)  The  Potential  of  Simulation  Models: 
Much  of  the  population  dynamics  research 
done  to  date  has  involved  single  or  multiple 
equation  regression  techniques  of  constrained 
maximization.  Within  the  capabilities  of  these 
techniques  one  (biology)  or  at  most  only  two 
(biology-economics)  disciplines  would  be  con- 
sidered, and  even  then  only  a  limited  number 
of  variables  in  each.  Many  of  the  twelve  points 
discussed  here  are  not  included  within  these 
analyses.  At  best  they  are  appended  on  an 
ad  hoc  basis. 

To  formalize  this  ad  hoc  process  one  would 
set  out  specifically  to  systematize  these  multiple 
considerations  via  a  simulation  model,  where 
each  consideration  would  appear  sequentially 
leading  to  outputs  which  would  represent 
many  combinations  of  these  interactions.  With- 
in this  framework  each  specialist  would  not 
be  trying  to  extend  his  own  area  to  include 
other  disciplines  in  the  process  of  specifying 
optimum  solutions.  Rather  he  would  merely 
be  characterizing  his  own  special  consider- 
ations, which  might  be  one  of  several  sub- 
routines in  the  entire  simulation  program. 
The  proper  manner  in  which  these  inputs 
would  be  combined  would  be  a  joint  responsi- 
bility of  all  researchers  providing  the  principal 
inputs. 

To  be  feasible,  each  separate  input  area 
must  have  reached  a  sufficient  stage  of  sophisti- 
cation and  accuracy  to  be  of  use  in  a  simula- 
tion model.  I  believe  this  judgment  can  now 
be  made.  This  suggests  that  the  work  that 
has  been  initiated  at  the  University  of  Wash- 
ington and  Massachusetts  Institute  of  Tech- 
nology should  be  expanded  to  encompass  all 
major  fisheries.  Work  on  other  water  resource, 
game  resource  management  problems  provide 
an  additional  base  of  expertise  to  facilitate 
development  of  these  models. 


Initially  these  models  would  include:  (1)  an 
assessment  of  the  resource  base,  (2)  a  popu- 
lation dynamics  model,  (3)  cost  and  earning 
functions,  (4)  demand  functions,  including 
provision  for  foreign  trade  flows,  (5)  exit-entry 
functions  based  on  profitability,  (6)  character- 
ization of  existing  and  alternative  regulatory 
constraints,  and  (7)  a  depiction  of  the  social 
response  function,  with  some  reference  to 
transfer  costs.  These  models  would  originally 
be  constructed  for  each  of  the  principal  fish- 
eries of  Alaska,  the  Pacific  Northwest,  the 
tuna  fisheries,  the  shellfish  and  menhaden 
fisheries  of  the  Gulf  and  the  Middle  Atlantic 
and  lobster  and  groundfish  in  the  North 
Atlantic.  Ultimately  multispecies  regional 
models  would  be  developed,  leading  to  a 
national  model  which  would  characterize  the 
entire  U.S.  fishing  industry. 

Initial  failures  in  the  construction  of  these 
models  will  suggest  immediate  research  needs. 
The  output  of  each  model  will  indicate  the 
sensitivity  of  each  component  of  the  model 
for  each  fishery. 

(11)  Artificial  Propagation  and  Fishery 
Management:  With  few  exceptions,  when  we 
identify  a  fishery  which  has  excess  capital 
and/or  labor  in  relation  to  the  sustainable 
resource  base  we  recommend  reduction  in 
these  inputs.  A  Canadian  fishermen's  group 
has  eloquently  phrased  another  course,  that 
is,  expand  the  resource  base.  This  would 
especially  be  recommended  if  the  incremental 
returns  from  dollar  expenditures  on  expansion 
exceeded  the  incremental  benefits  from  dollars 
spent  withdrawing  inputs. 

At  this  time  such  a  possibility  could  only 
be  anticipated  for  Pacific  salmon.  Several 
factors  could  enhance  these  trade-offs,  among 
these  being  the  possibility  that  demand  rising 
faster  than  costs  would  bring  the  cost  of 
hatchery  production  into  a  more  favorable 
light  and  a  full  realization  of  the  political 
resistance  to  withdrawing  excess  inputs.  With 
the  further  development  of  hatchery  tech- 
nology other  fisheries,  perhaps  shellfish,  may 
be  supplemented  by  artificial  propagation  and 
rearing.  As  this  occurs  it  will  be  necessary 
to  include  the  dimensions  of  this  alternative 
as  an  new  subroutine  in  the  simulation  models 
discussed  in  point  10  above. 


(12)  Competing  Uses:  A  new  dimension,  an 
additional  complication,  has  entered  upon  the 
scene  of  fishery  management,  suggesting  new 
priorities  here  as  it  has  elsewhere.  It  comes 
under  the  banner  of  ecology,  an  old  word 
with  new  urgency.  With  the  scarcity  of  natural 
resources  increasing  relative  to  multiple  de- 
mands, and  with  the  new  insistence  upon 
quality  in  addition  to  (or  rather  than)  merely 
quantity,  the  management  of  coastal  resources 
has  suddenly  taken  on  a  new  dimension.  Man- 
agement of  commercial  fisheries  will  be  obliged 
to  reflect  this  trend. 

Coastal  fisheries  must  now  be  managed  as 
part  of  the  total  coastal  resource.  No  sug- 
gestion has  yet  been  made  as  to  how  this 
will  be  done.  Suffice  to  say  that  such  critical 
issues  as  fishery  tolerances  to  certain  water 
quality  levels  and  the  interrelationship  be- 
tween sports  and  commercial  fisheries  will 
be  critical  issues.  I  will  forego  an  attempt  to 
treat  this  issue  in  a  few  brief  paragraphs 
here,  acknowledging  the  likelihood  that  the 
next  fisheries  workshop  will  certainly  treat 
this  area  as  one  of  its  principal  topics. 


With  this  general  background  on  the  princi- 
pal issues  in  fisheries  management  we  can 
now  look  briefly  at  the  workshop  contributions 
to  summarize  their  contents.  With  these  papers 
serving  as  the  stimulus  the  discussions  at 
the  workshop  inevitably  revolved  around  two 
related  issues:  (1)  the  necessity  for  develop- 
ing short  term  models  due  to  the  extreme 
urgency  of  resource  management  problems 
in  many  fisheries  and  (2)  the  need  to  assume 
the  full  responsibility  for  measuring  all  social 
costs  associated  with  alternative  resource  use 
plans  and  to  suggest  ways  by  which  these 
social  costs  can  be  minimized. 

At  the  conclusion  of  this  workshop  one 
was  definitely  left  with  the  impression  that 
if  significant  steps  cannot  be  made  in  both 
of  these  areas  in  the  near  future  (2-4  years) 
then  serious  questions  will  have  to  be  raised 
about  the  utility  of  the  bioeconomic, 
socio-political  research  and  planning  which 
we  are  conducting.  In  this  light  much  of  the 
work  reported  at  the  workshop  provides  some 
encouragement  that  progress  will  be  made 
on  these  issues. 


ISSUES  IN  FISHERY  MANAGEMENT 


The  opening  paper  by  Van  Meir  appropri- 
ately cites  the  Burkenroad  observation  that 
fisheries  should  be  managed  for  people  not 
fish,  a  trite,  but  occasionally  overlooked  ad- 
monition. He  emphasizes  that  the  critical 
element  now  is  that  time  is  running  out  in 
many  fisheries.  The  solution  is  to  replace  com- 
mon rights  with  private  rights,  these  rights 
to  be  consistent  and  in  balance  with  allow- 
able yield.  The  program  should  not  only 
permit,  but  also  promote  economic  efficiency 
both  in  the  short  run  and  in  the  long  run. 

To  begin  limited  entry  programs  we  must 
emphasize  three  areas:  (1)  a  resolution  of 
jurisdictional  conflict,  (2)  an  educational 
program  which  will  communicate  the  poten- 
tial benefits  and  dispel  the  idea  that  the 
scheme  is  to  be  a  government  monopoly  and, 
(3)  trial  programs  which  will  demonstrate 
how  limited  entry  operates  in  practice. 

Van  Meir  suggests  that  in  practice  we  must 
be  willing  to  accept  a  second-best  solution, 
i.e.,  agree  with  biologists  on  harvesting  maxi- 
mum sustainable  yield  (MSY)  and  proceed  to 
specifying  the  most  efficient  way  of  doing  this. 
We  must  develop  a  system  which  will  insure 
that  fishing  rights  will  be  allocated  to  the 
most  efficient  producer  at  any  point  in  time. 

Van  Meir  concludes  by  suggesting  a  system 
for  doing  this.  It  is  here  that  he  introduces 
the  first  real  element  of  controversy.  He  sug- 
gests a  licensing  mechanism.  Licenses  would 
be  allotted  so  as  to  include  all  grandfather 
rights.  They  would  be  reduced  by  attrition 
with  the  total  number  changing  as  technology 
changes.  Monopoly  powers  would  be  restricted 
and  rents  would  be  redistributed  via  license 
fees  or  taxes. 

Undoubtedly  this  is  a  reasonable  step  toward 
a  politically  palatable  solution.  Others  would 
argue  that  there  are  other  schemes  that 
would  be  more  appropriate  for  certain  fish- 
eries. They  would  argue  that  this  proposal 
contains  the  same  faults  as  U.S.  agricultural 
programs  of  the  past  decade,  where  a  central 


authority  is  granted  the  right  to  determine 
the  number  of  licenses.  To  do  so  it  must  use 
existing  measures  of  technological  capacity 
and  technological  change,  when  both  of  these 
may  change  substantially  under  the  exogenous 
influence  of  a  newly  introduced  licensing 
scheme.  Some  alternative  suggestions  would 
allow  both  the  rate  of  technological  change 
and  the  size  and  number  of  property  rights 
to  be  determined  within  the  market  mechan- 
ism. The  paper  presented  by  Holmsen  refers 
briefly  to  one  alternative.  Also  the  paper  by 
Carlson  could  serve  as  a  basis  for  preliminary 
calculations  of  the  appropriate  number  of 
licenses  in  the  tuna  and  groundfish  fisheries. 

Pontecorvo  introduces  several  broad  con- 
ceptual issues,  among  these  being  the  need 
for  short  run  models  which  can  be  utilized 
directly  in  resource  management.  If  the  short 
run  is  critical  we  should  examine  those  models 
which  appear  to  be  more  satisfactory  for  the 
short  run. 

Pontecorvo  focuses  upon  the  difficulties  of 
choosing  biological  models  and  combining 
these  with  economic  models  for  both  short 
run  and  long  run  analysis  —  to  determine 
optimum  solutions.  He  cites  violent  fluctua- 
tions in  the  Pacific  red  salmon  resource  as 
a  characteristic  which  militates  against  the 
use  of  even  short  run  models,  and  also  where 
the  costs  of  improving  the  information  flow 
may  exceed  the  benefits.  Further  complica- 
tions arise  due  to  instability  on  the  economic 
side  (demand  and  the  general  state  of  the 
economy)  and  changing  political  and  social 
considerations.  One  suggestion  here  is  that  a 
program  geared  to  catch  some  average  level, 
less  than  the  allowable  yield  during  the 
highest  year,  may  be  the  desirable  economic 
solution  —  one  case  where  we  would  suggest 
taking  less  than  MSY. 

Pontecorvo's  position  on  social  and  political 
issues  is  that  these  are  fully  accounted  for 
(albeit  incorrectly)  in  the  economists'  assump- 
tions of  full  employment  and  factor  mobility. 


The  economists'  assumption  of  human  rational- 
ity forces  the  social-political  ordering  into 
the  same  ordering  as  economics.  The  more 
reality  deviates  from  this  ordering  the  more 
the  economic  conclusions  must  be  altered  by 
subsequent  ad  hoc  social  and  political  con- 
siderations. 

This  can  be  extended  to  multiple  use  issues 
as  well.  Often  we  treat  the  fishery  as  if  it  were 
the  only  user  of  the  resource.  Future  regu- 
latory organizations  will  have  to  incorporate 
such  considerations  directly  and  this  will  affect 
the  design  of  these  organizations. 

Rettig  adds  to  the  mounting  chorus  warning 
of  the  social  implications  of  certain  fisheries 
management  plans.  He  suggests  that  these 
may  lead  us  to  actually  restructure  the  ob- 
jectives of  these  plans.  Absence  of  these  con- 
siderations may  be  one  reason  for  our  failure 
to  initiate  revised  management  programs. 
Other  reasons  for  failure  may  be  the  present 
existence  of  a  severe  divergence  between  the 
objectives  of  administrators  and  researchers, 
incompletely  specified  models,  or  the  mere 
absence  of  sufficient  educational  programs. 

Regarding  the  incorrectly  specified  models 
Rettig  makes  the  intriguing  observation  that 
market  imperfections  on  the  buyers'  side  could 
alter  the  optimum  solution.  Ignoring  this 
fact  would  actually  result  in  a  further  mis- 
allocation  of  resources.  He  suggests  a  further 
evaluation  of  inter-market  linkages  before 
making  irreversible  management  steps. 

Additional  issues  which  must  be  faced  are 
the  multispecies  management  problems  and 
the  absence  of  a  reasonable  discount  rate  in 
the  sustainable  yield  curve.  This  relates  to 
some  degree  to  his  final  conclusion  that  we 
must  include  so  many  diverse  factors  that  in 
the  end  our  "theory"  may  be  useless.  Never- 
theless, like  many  others  as  well  as  partici- 
pants at  this  workshop,  he  can  see  no  other 
alternative  but  to  follow  this  course  unless 
we  intend  to  ignore  realism  and  the  needs 
of  fishery  administrators. 

In  the  last  of  four  general  papers  on  the 
issues  in  fishery  management,  Crutchfield  re- 
views the  inputs  to  fishery  modeling  work 
now  developing  for  four  Pacific  Northwest 
fisheries:  anchovy,  salmon,  king  crab,  and 
halibut. 


These  models  have  three  basic  components: 
economics,  biology,  and  law.  In  the  economics 
portion  the  cost  and  earnings  and  profit  and 
loss  statements  for  representative  vessels  are 
developed,  related  to  certain  catch  rates,  tech- 
nological factors  and  market  conditions  (pro- 
duct price,  interest  rate,  alternative  employ- 
ment). By  this  manner  the  complete  operation 
of  vessels  in  the  selected  fisheries  can  be  speci- 
fied and  from  this  it  is  possible  to  construct 
an  exit-entry  function  which  would  relate  to 
changes  in  these  economic  variables,  indepen- 
dently, or  as  affected  by  biological  and/or 
legal  variables. 

The  biological  elements  of  this  model  include 
gross  stock  parameters  and  a  yield-effort 
function  which  generates  catch  rates,  these 
serving  as  direct  input  into  both  the  economic 
model  and  into  population  dynamics  compo- 
nents of  the  biological  model.  In  the  case  of 
the  salmon  fishery  separate,  though  similar 
models,  are  developed  for  five  different  stocks 
at  ten  locations,  a  50-cell  matrix.  Any  per- 
tinent species  interactions  are  also  included. 

The  legal  portion  of  this  model  specifies 
the  existing  regulatory  structure  which  may 
determine  the  components  of  both  the  biological 
and  economic  models,  determining  what  is  fished 
for,  when,  how,  and  to  what  extent.  As  in 
other  portions  of  the  model,  alternative  legal 
structures  will  be  posited  to  allow  for  alterna- 
tive patterns  of  resource  utilization. 

The  ultimate  purpose  of  this  model  is  to 
take  a  complete  interdisciplinary  approach 
to  fisheries  management.  Alternative  manage- 
ment programs  will  be  specified.  Among  these 
an  optimum  plan  will  be  identified,  with  the 
sequence  of  steps  which  would  most  effective- 
ly lead  toward  this  plan.  In  its  most  extensive 
form  this  model  will  consider  multiple  species 
management  cases  such  as  anchovy-mackerel- 
tuna  in  California  and  salmon-tuna-crab- 
halibut  of  the  Pacific  northwest.  As  empha- 
sized by  Crutchfield,  in  its  present  form  the 
model  emphasizes  the  multidisciplinary  nature 
of  the  management  problem  and  will  readily 
incorporate  many  of  the  suggestions  made 
at  this  workshop. 

A.  A.  S. 


Problems  in  Implementing  New  Fishery  Management  Programs 


Lawrence  W.  Van  Meir1 
ABSTRACT 


Even  though  an  "optimum"  management  program,  in  an  economic  sense,  may 
never  be  achievable  in  the  management  of  commercial  fisheries,  changes  can  be 
initiated  which  will  allow  individual  governments  to  realize  economic  gains  over 
the  status  quo  in  harvesting  common  property  fishery  resources.  These  changes 
primarily  involve  jurisdictional  issues;  country  quotas  for  international  fisheries; 
accord  between  the  Federal  government  and  the  states;  and  a  within-industry 
system  for  allocating  fishing  rights.  A  system  of  vessel  licensing  is  described  with 
reference  to  the  ultimate  use  of  licenses  on  units  of  fishing  effort. 


The  management  of  fisheries  is  intended 
for  the  benefit  of  man,  not  fish,  therefore, 
effect  of  management  upon  fish  stocks  cannot 
be  regarded  as  beneficial  per  se. 

Martin  D.  Burke?iroad 

These  words  by  Burkenroad  were  published 
almost  20  years  ago.  This  statement  is  a  par- 
ticularly cogent  phrasing  of  the  crux  of  the 
question  of  fishery  management  for  it  raises 
both  the  question  of  what  benefits  will  be 
sought  in  managing  fisheries  and  the  question 
of  to  which  men  will  these  benefits  accrue. 
These  are  the  two  64  dollar  questions  in  the 
area  of  fishery  management  policy. 

In  spite  of  Burkenroad's  admonition  that 
the  conservation  of  fish  stocks  per  se  cannot 
be  regarded  as  beneficial,  and  articles  and 
studies  on  the  economic  aspects  of  fishery 
management  that  have  appeared  in  the  last 
decade,  most  fishery  management  programs 
remain  oriented  to  the  conservation  of  fish 
stocks  with  no  consideration  of  the  economic 
results  that  may  be  obtained.  We  still  resort 
to  practices  that  either  encourage  the  ineffi- 
cient use  of  vessels,  gear,  and  labor,  or  that 
limit  and  impede  the  efficient  use  of  these 
economic  inputs  as  a  means  of  conserving 
fish  stocks.  This  does  not  mean  that  we  do 
not    advocate    conservation,    but    rather    that 


Staff  Economist,  National  Canners  Association. 


we  state  more  completely  our  objectives  for 
the  conservation  program. 

Time  is  running  out  on  us.  With  technologi- 
cal development  yielding  a  3.5  to  4.0%  annual 
increase  in  the  productivity  of  labor  in  the 
economy,  the  fishing  industry  will  find  itself 
in  an  ever  increasing  economic  squeeze  if 
positive  steps  cannot  be  taken  to  include  eco- 
nomic objectives  in  fishery  management.  We 
may  continue  to  conserve  fish  stocks  but  it 
will  not  be  for  the  benefit  of  U.S  fishermen  or 
fishing  communities. 

The  entire  problem  of  fishery  management 
of  course  stems  from  the  common  property 
status  of  fishery  resources.  In  the  past,  when 
scientific  evidence  indicated  that  a  particular 
fish  stock  was  being  overfished,  or  in  danger 
of  being  overfished,  the  solution  was  to  place 
a  quota  on  the  fishery  and/or  add  regulations 
that  either  impaired  the  efficiency  of  fishing 
gear,  or  in  some  cases  required  the  use  of 
inefficient  gear  and  fishing  methods.  The  con- 
sequences of  such  programs  have  been  com- 
pletely discussed  in  other  articles  and  are 
not  the  purview  of  this  paper.  Instead  I  want 
to  concentrate  on  the  question  of  what  must 
be  accomplished  to  change  the  situation,  and 
how  it  is  to  be  done. 

Obviously,  common  property  status  must 
be  replaced  by  explicit  fishing  rights.  More- 
over, in  the  process  of  conserving  fish  stocks 
we  must  do  so  by  bringing  these  specific  fishing 
rights  in  balance  with  the  allowable  yield  of 
the  resource  in  a  manner  that  not  only  permits 
efficiency  but  also  actually  promotes  economic 


efficiency  and  technological  development  in 
both  the  short  and  the  long  run.  In  short, 
some  system  must  be  developed  to  limit  the 
amount  of  labor  and  capital  employed  in  har- 
vesting the  allowable  catch  and  at  the  same 
time  assure  that  the  labor  and  capital  is  used 
in  an  economically  efficient  manner. 

Various  economic  advantages  can  result 
from  limited  entry.  Catch  per  vessel  and  fisher- 
man employed  will  increase,  thus  increasing 
wages  and  return  on  investment.  The  overall 
value  of  fish  landed  may  be  increased  in  some 
cases  if  the  fishery  management  program  re- 
sults in  a  better  marketing  pattern.  Labor 
and  capital  employed  in  processing  and  dis- 
tribution can  be  brought  into  better  balance 
with  the  volume  of  fish  processed,  thus  realiz- 
ing economic  gains  in  these  sectors. 

A  number  of  complex  problems  must  be 
overcome  in  order  to  realize  the  fruits  of 
limited  entry.  In  the  first  place,  the  question 
of  fishery  jurisdiction  must  be  solved.  In  the 
case  of  international  fisheries,  the  solution  to 
jurisdiction  will  necessitate  some  system  of 
national  quotas.  Once  national  quotas  have 
been  agreed  to  and  established,  then  each 
individual  nation  can  institute  its  own  pro- 
gram for  harvesting  its  quota.  Jurisdictional 
problems  also  exist  between  States  and  be- 
tween the  Federal  government  and  the  States. 
Many  fish  stocks  are  fished  by  fishermen  from 
more  than  one  State.  In  the  case  of  pelagic 
fish,  the  fish  may  migrate  through  the  waters 
of  several  States  or  between  international 
waters  and  waters  under  the  jurisdiction  of 
several  States.  Moreover,  a  specific  fishery 
may  involve  waters  under  the  jurisdiction  of 
several  States  and  the  Federal  government. 
Consequently,  no  one  jurisdiction  or  authority 
by  itself  can  come  to  grip  with  the  problem. 
Certain  enabling  legislation  will  be  needed  at 
both  the  Federal  and  State  levels  of  government. 

An  important  prerequisite  to  solving  the 
jurisdictional  and  legal  questions  will  be  a 
thorough  understanding  of  the  concept  of 
limited  entry,  and  the  need  for  limited  entry, 
on  the  part  of  the  fishermen,  government  of- 
ficials, and  congressional  representatives. 
Many  individuals  in  commercial  fishing  today 
are  convinced  of  the  necessity  for  a  limitation 
on  the  entry  of  labor  and/or  capital  in  those 
fisheries    that    are    fully    exploited.    However, 


these  individuals  are  still  in  a  minority.  Some- 
how, the  problems  -we  are  facing  in  many  of 
our  fisheries,  and  the  effectiveness  of  limited 
entry  in  dealing  with  these  problems,  must 
be  brought  to  the  attention  of  the  rest  of  the 
commercial  fishing  industry  in  a  meaningful 
way. 

One  reason  why  many  commercial  fisher- 
men are  wary  of  limited  entry  proposals  may 
be  because  they  have  not  been  presented  a 
specific  proposal  to  study  and,  hence,  are 
understandably  cautious  about  embracing  a 
new  concept  without  having  some  idea  of 
how  they  might  fare  under  the  new  regime. 
Thus,  specific  proposals  likely  will  have  to  be 
worked  out  and  presented  to  industry  as  a 
step  in  overcoming  their  resistance  to  the  idea. 

A  second  reason  why  some  people  are  sus- 
picious of  the  concept  of  limited  entry  is  be- 
cause they  have  formed  the  opinion  that 
limited  entry  is  a  scheme  to  put  the  govern- 
ment in  monopolistic  control  of  fisheries  to 
enable  them  to  extract  either  the  monopoly 
profit  or  economic  rent  from  the  fishery.  Econo- 
mists may  have  contributed  to  this  image  in 
their  writings  on  objectives  of  fishery  manage- 
ment. 

From  the  viewpoint  of  the  economist,  pro- 
gress toward  a  more  rational  fishery  manage- 
ment program  will  be  a  process  of  accepting 
second  best  solutions.  One  of  the  first  instances 
in  which  we  need  to  be  willing  to  accept  a 
second  best  solution,  at  least  initially,  is  on 
the  objective  of  a  fishery  management  pro- 
gram. If  we  accept  the  historical  precedence 
of  "maximum  sustainable  yield"  (MSY)  and 
seek  agreement  with  the  biologist  on  the  im- 
portance of  harvesting  the  MSY  as  efficiently 
as  possible,  it  should  be  possible  for  the  econo- 
mist and  the  biologist  to  approach  industry 
with  a  common  argument.  The  improvement 
in  returns  to  capital  and  labor  in  moving 
from  present  management  methods  to  a  method 
which  achieves  a  reasonable  degree  of  effi- 
ciency in  harvesting  the  MSY,  will  represent 
the  major  share  of  total  improvement  in  re- 
turns that  might  result  from  any  other  man- 
agement objective.  As  a  starting  point  I  would 
suggest  that  the  emphasis  be  placed  on  im- 
proving the  returns  to  labor  and  capital  in 
the  fishery  management  program  while  de- 
leting the  argument  for  either  seeking  to  maxi- 


10 


mize  net  economic  return  or  economic  rent. 

If  we  are  going  to  accept  MSY  as  the  basis 
for  managing  a  fishery,  what  economic  ob- 
jectives should  we  try  to  build  into  a  new 
fishery  management  program?  As  I  mention- 
ed earlier,  the  first  objective  would  be  to 
seek  the  optimum  amount  of  effort  to  harvest 
the  MSY,  or  at  least  a  reduction  in  effort, 
as  a  means  of  improving  the  catch  and  re- 
turn per  unit  of  effort  in  the  fishery.  In  ad- 
dition to  this,  we  should  also  seek  to  build 
into  the  management  program  some  means  of 
insuring  continued  efficiency  through  time. 
This  means  that  over  time  the  management 
program  must  allocate  the  fishing  right  to 
those  economic  resources  that  are  most  ef- 
ficient in  fishing.  Thirdly,  the  program  must 
stimulate  the  development  and  adoption  of 
technological  advancements  in  fishing.  How 
can  these  objectives  be  attained  in  fishery 
management? 

The  system  I  foresee  consists  of  a  com- 
mercial fishing  license  issued  either  by  the 
Federal  government  or  by  joint  agreement  of 
the  Federal  government  and  the  individual 
States  concerned.  Each  license  issued  would 
represent  a  specific  amount  of  fishing  effort. 
Initially,  the  number  of  licenses  and  fishing 
effort  would  have  to  accommodate  all  vessels 
and  crews  that  have  historically  been  employ- 
ed in  the  fishery.  However,  as  vessels  were 
retired  from  use,  licenses  would  be  cancelled 
until  normal  attrition  reduced  the  amount  of 
effort  to  the  desired  level.  When  the  number 
of  licenses  have  been  reduced  to  the  optimum 
number,  a  market  for  the  licenses  would  be 
allowed  to  develop.  Licenses  could  be  sold  or 
leased.  Thus,  the  more  efficient  manager  of 
a  fishing  enterprise  would  be  given  the  op- 
portunity to  lease  or  buy  fishing  rights  from 
the  less  efficient  operator.  The  total  number 
of  licenses  could  be  adjusted  over  time  as 
productivity  of  fish  stocks  and  technology 
merited.  In  this  manner,  the  management 
program  would  work  toward  allocating  the 
limited  fishing  rights  to  those  fishing  firms 
that  were  most  efficient.  Moreover,  it  would 
now   be   advantageous   to   the   fishing  firm   to 


seek  means  of  improving  its  efficiency  and  to 
adopt  new  technology  that  improved  efficiency. 
Limitations  could  be  placed  on  the  number  of 
licenses  that  any  one  company  could  own  or 
control  as  a  means  of  preventing  someone 
from  developing  a  monopoly  over  the  fishery. 

A  licensing  scheme  of  this  nature  would 
generate  a  certain  amount  of  economic  rent 
in  the  fishery.  This  economic  rent  could  either 
be  taxed  away  in  the  form  of  the  license  fee 
or  could  be  allowed  to  accrue  to  the  resources 
employed  in  the  fishery.  If  the  rent  is  taxed 
away,  it  should  be  used  for  administration  of 
the  management  program  and  for  research 
on  and  development  of  the  resource  itself.  If 
the  rent  is  allowed,  either  in  total  or  in  part, 
to  accrue  to  the  resources  employed  in  the 
fishery,  it  would  in  turn  be  redistributed  in 
the  economy  through  taxes,  and  would  ulti- 
mately be  built  into  the  cost  of  production 
through  the  cost  of  fishing  licenses.  In  either 
case,  this  should  not  be  a  serious  deterrent 
to  initiating  a  licensing  system. 

One  of  the  most  difficult  problems  to  cope 
with  in  the  licensing  system  will  be  the  adop- 
tion of  the  effort  base  for  the  licenses.  Licenses 
could  be  based  on  vessels,  tonnage  of  fish,  or 
an  index  of  fishing  effort.  Vessels  would  be 
the  least  adequate  base  for  issuing  fishing 
licenses  because  of  the  variation  in  fishing 
ability  from  vessel  to  vessel.  It  would  certainly 
seem  technically  possible  to  develop  an  index 
of  fishing  effort  and  to  define  licenses  in  units 
of  fishing  effort.  Ideally  the  units  of  fishing 
effort  would  be  the  link  between  the  economic 
aspects  of  harvesting  and  the  biological  model 
used  in  assessing  stock  and  yield  character- 
istics. 

The  steps  toward  a  more  rational  system  of 
managing  fisheries  will  no  doubt  be  a  system 
of  compromises.  Perhaps  the  way  to  facilitate 
these  compromises  with  the  least  loss  of  time 
and  effort  will  be  to  include  the  commercial 
fishing  industry  in  the  actual  development 
of  the  specifics  of  new  management  programs. 
The  situation  is  sufficiently  urgent  that  this 
should  be  given  the  highest  priority  in  our 
"new  look"  toward  the  oceans. 


11 


On  the  Utility  of  Bioeconomic  Models  for  Fisheries  Management ' 


GlULIO  PONTECORVO2 


ABSTRACT 

Short  run  and  long  run  biological  and  economic  models  are  inevitably  bound  to- 
gether in  any  comprehensive  plan  to  manage  commercial  fisheries.  While  these 
disciplines  can  be  treated  rigorously,  political  and  social  considerations  can  be 
considered  only  generally  and  therefore  on  an  ad  hoc  basis.  Within  this  framework 
long  run  models  are  useful  primarily  for  goal  setting.  More  work  must  be  done  in 
developing  short  run  models  which  will  measure  the  immediate  biological  and  economic 
impacts  of  alternative  management  steps  in  addition  to  immediate  political  and  social 
ramifications.  Emphasis  would  then  be  placed  upon  the  economic  sources  of  short  run 
instability,  with  an  initial  economic  rationalization  of  the  fishery  providing  the  funds 
for  subsequent  management  and  biological  forecasting  which  will  concentrate  on  ex- 
tending management  from  a  rationalized  fishery  at  a  given  harvesting  level  to  rational- 
ized fishing  at  some  optimum  level. 


BIOLOGICAL  MODELS 
The  Yield  from  Ocean  Resources 

A  19th  century  view  was  that  the  high  seas 
fisheries  were  inexhaustible.  We  who  are  pre- 
occupied by  our  failure  to  control  our  own 
numbers  and  the  possibility  of  worldwide  eco- 
logical disaster  can  at  best  regard  such  notions 
as  quaint.  Nevertheless,  certain  implications 
of  the  19th  century  view  of  the  oceans,  with 
its  infinitely  elastic  aggregate  supply  curve 
for  fish,  are  worth  considering  here  in  our 
attempt  to  understand  the  biologists'  concept 
of  the  maximum  sustainable  physical  yield.3 

The  need  for  biological  regulation  of  fish- 
eries arose  because  the  economic  interests 
involved  in  exploiting  these  resources  became 
aware,  through  rapidly  declining  yield-effort 
relationships,  that  the  supply  of  any  particular 
species  was  limited.4  Recognition  of  the  exis- 


1  I  wish  to  express  my  thanks  for  helpful  sugges- 
tions to  Dr.  Brian  Rothschild  and  Dr.  EdilbertoSegura. 

2  Professor  of  Economics,  Graduate  School  of  Busi- 
ness, Columbia  University. 

3  The  elasticity  of  the  supply  function  may  be  with 
respect  to  the  inexhaustibility  of  one  stock  of  fish  or 
as  indicated  below  it  may  arise  from  a  process  of  sub- 
stitution of  one  stock  for  another. 

4  The  concept  of  limit  is  highly  ambiguous.  It  may 
be  thought  of  as  the   maximum   sustainable  yield;   the 


tence  of  these  limits  led  the  biologist  to  in- 
vestigation of  the  characteristics  of  particular 
populations  in  order  to  find,  if  possible,  a 
level  of  exploitation  that  would  be  safe.  A  safe 
level  is  defined  as  that  maximum  rate  of  ex- 
ploitation that  would  preserve  the  stock,  and 
yet  allow  the  catch  to  continue  at  the  maxi- 
mum rate  through  time.5  The  imposition  on 
the  stock  of  the  appropriate  (to  achieve  the 
maximum  sustainable  yield)  level  of  fishing 
mortality  involves  the  movement  from  one 
long  run  equilibrium  condition  of  the  fish 
population  to  another,  with  both  equilibriums 
considered  stable.6 

The  development  of  the  argument  thus  far 


level  of  scarcity  of  fish  beyond  which  it  would  not 
pay  to  fish,  i.e.,  an  economic  limit,  the  level  that  maxi- 
mizes the  net  economic  yield  from  the  resource,  or  in 
complete  destruction  of  the  stock. 

5  For  the  development  of  an  alternative  view  of 
the  relationship  between  the  maximum  sustainable 
yield  and  the  net  economic  yield,  see  (Schaefer,  1970a). 
Schaefer  also  develops  the  necessary  conditions  for 
adequate  biological  management  (p.  9ff.) 

H  The  first  equilibrium  is  the  natural  or  unexploited 
state  of  the  stock.  The  second  is  the  condition  of  the 
stock  being  exploited  at  the  maximum  sustainable  yield. 
There  is  a  question  about  the  relative  stability  of 
the  two  equilibriums  both  per  se  and  also  because 
of  the  effect  of  fishing  effort  on  ecological  conditions. 
More  simply,  populations  that  are  heavily  exploited 
by  fishermen  may  show  greater  fluctuations  in  stock 
size. 


12 


has  already  led  to  a  conceptual  difficulty.7 
There  is  some  evidence  that  in  certain  of  the 
populations  which  historically  were  overfished 
the  costs  to  society  of  rehabilitation  of  the 
populations  exceeded  the  benefits  from  the 
subsequent  higher  yields  which  resulted  from 
the  imposition  of  the  biological  control  pro- 
gram required  to  obtain  the  maximum  sus- 
tainable yield. 

The  debate  between  economists  and  biolo- 
gists over  the  "success"  or  "failure"  of  the 
International  North  Pacific  Halibut  Commis- 
sion as  an  instrument  in  fisheries  manage- 
ment is  an  illustration  of  this  type  of  difficulty.8 

Aggregate  Yield  and  Biological 
Models  of  Particular  Species 

A  more  subtle  set  of  difficulties  is  involved 
in  the  interrelationships  between  an  aggregate 
yield  function  and  the  partial  yield  functions 
derived  by  the  biologists  for  particular  popu- 
lations. In  the  recent  development  of  the  eco- 
nomic   literature   on    fisheries   the   economists 


7  More  precisely,  if  biological  overfishing  has  occurred, 
and  if  a  population  is  pushed  well  beyond  the  second 
equilibrium  point,  does  it  enhance  the  material  well- 
being  of  the  society  to  spend  time  and  money  (labor 
and  capital)  to  restore  it  to  the  second  equilibrium? 

8  It  is  also  appropriate  to  point  out  that  the  Halibut 
Commission  has  been  an  important  training  ground 
for  biologists  interested  in  population  dynamics  and 
fisheries  management  problems,  a  benefit  not  included 
in  the  economic  calculus.  And  also  it  is  important 
to  note  that  there  should  be  a  clear  distinction  between 
population  dynamics  as  part  of  an  academic  discipline 
and  the  administrative  process  of  a  social  institution 
such  as  a  commission.  The  two  activities  have  differ- 
ent goals,  but  in  the  field  the  practitioners  interact  so 
closely  it  is  difficult  for  outsiders  to  observe  the  dis- 
tinction. For  a  biologist's  view  of  the  Halibut  Com- 
mission's work,  see  Schaefer  (1970a,  p.  14).  Another 
illustration  may  be  in  the  sea  lamprey  control  pro- 
gram. 

Since  the  economist,  like  St.  George,  traditionally 
defends  the  general  welfare  —  the  maximization  of 
Gross  National  Product  —  of  the  society  against  the 
onslaught  of  particular  interests  it  seems  appropriate 
to  me  to  make  this  argument  an  economic  one.  What 
is  omitted,  however,  from  the  economic  analysis  is 
adequate  consideration  of  the  place  (role)  of  any  par- 
ticular species  in  the  ecological  structure.  This  omis- 
sion, together  with  the  usual  inadequacy  of  the  defini- 
tion of  the  appropriate  rate  of  social  discount,  is 
probably  sufficiently  important  to  make  one  want  to 
proceed  with  care  with  a  decision  to  fish  out  a  resource. 
It  does  not  follow,  however,  that  such  a  resource 
should  be  "saved"  by  a  costly  biological  rehabilitation 
program. 


have  looked  to  the  biologists'  yield  function 
for  a  particular  species  as  the  source  of  the 
production  function  upon  which  subsequent 
economic  analysis  of  the  fishery  can  rest. 
This  development  is  logical  in  that  the  most 
money  for  biological  research  has  been  spent 
(for  the  most  part)  on  those  populations  that 
are  economically  interesting.  More  directly 
put,  the  market  demand  for  fish  has  been  an 
important  determinant  of  the  direction  of 
application  of  biological  work. 

In  the  historical  development  of  ocean  fish- 
eries the  interaction  between  market  forces 
and  biological  limits  on  the  supply  represented 
by  specific  fish  populations  has  been  a  typical 
case  of  exploitation  at  an  extensive  margin.  In 
the  long  run,  for  certain  fisheries  (given  a  posi- 
tive income  elasticity  of  demand)  operating  on  a 
particular  stock  of  fish,  there  has  been  a  ten- 
dency for  the  fishery  to  extend  itself  both  geo- 
graphically and  temporally.  If,  after  this  ex- 
tension has  taken  place,  we  assume  that  through 
the  imposition  of  biological  regulation  the 
supply  function  for  the  stock  in  question  be- 
comes infinitely  inelastic,  economic  adjustments 
will  take  place  and  the  fleet  will  tend  to  move 
on  to  another  similar  stock  or  perhaps  to  a  com- 
pletely differentiated  stock. 

Thus  we  have  the  development  of  fisheries 
management  essentially  on  an  ad  hoc  basis 
as  a  response,  often  belated,  to  the  expansion 
of  fishing  effort  against  a  finite  supply  of  fish. 
The  continuous  expansion  of  fisheries  at  the 
margin  (taken  collectively)  has  resulted  in  an 
aggregate  supply  curve  which  has  been  elastic. 
World  production  of  protein  from  the  oceans 
has  risen  and  is  expected  to  continue  to  rise. 
The  ultimate  limit  will  be  determined  by  a 
trade  off  between  the  capacity  of  the  basic 
chemical  biological  processes  of  the  oceans 
to  produce  protein  and  the  cost  of  collecting 
it.9  At  the  same  time  the  expansion  in  world 
fisheries  has  tended  to  conceal  the  condition 
of  the    specific    stocks    already    exploited    and 


9  The  inputs,  the  labor  and  capital  utilized  to  bring 
about  the  increase  in  aggregate  output  of  fish  are 
not  on  the  average  highly  specialized.  Both  are  able 
to  shift  from  one  -fishery  to  another.  Vessel  construc- 
tion and  reconditioning  is  a  relatively  easy  process. 
Labor  immobility  is  a  larger  problem  but  it  is  less 
acute  in  high  seas  than  in  inshore  fisheries. 


L3 


complicate  the  problem  of  management  (Stew- 
art and  Pontecorvo,  1970). 

For  a  first  approximation  we  may  visualize 
the  fisheries  management  apparatus  as  re- 
quiring several  aggregate  yield  functions  as 
operational  concepts.  The  plurality  is  neces- 
sary because  our  concepts  of  aggregate  yield 
are  ambiguous:  it  may  be  a  regional  concept, 
a  concept  associated  with  a  particular  level 
of  the  food  chain,  a  concept  that  involves  a 
set  of  stocks  that  are  either  economically  or 
ecologically  consistent  in  some  way,  etc.  The 
ambiguities  involved  derive  from  the  inade- 
quacy of  biological  knowledge  of  aggregate 
yield,  what  the  economists  might  call  macro 
biological  ocean  processes,  and  also  from  the 
open  access  common  property  status  of  the 
stocks.  This  latter  situation  permits  each 
nation  state  to  define  its  output  goals  in  terms 
of  its  own  tastes  (exploitation  of  alternative 
species)  and  then  to  proceed  to  bargain  for 
its  share  in  purely  nationalistic  terms. 

In  the  absence  of  adequate  goals  at  the 
level  of  aggregate  yield,  fishery  administrators 
are  left  dealing  with  partial  equilibrium 
systems,  i.e.,  yield  functions  for  particular 
species,  a  circumstance  which  makes  them 
particularly  vulnerable  to  pressure  from  eco- 
nomic interests,  fluctuations  in  the  stocks  and 
the  interaction  between  the  two. 

Population  Dynamics 

The  population  models  developed  by  the 
biologists  are  basically  consistent  with  economic 
models.  Difficulties  in  the  process  of  data  col- 
lection, statistical  problems  in  fitting  functions, 
and  the  development  of  accurate  forecasts  are 
all  familiar  ground.  The  question  of  the  ade- 
quacy and  the  cost  of  basic  data  does  require 
further  comment.  The  observation  of  wild 
populations  is  a  time-consuming  and  costly 
process.  The  fishermen  are  close  observers  of 
the  behavior  of  these  populations  and  the  com- 
mercial catch  is  therefore  an  important  data 
source.  Several  types  of  bias  may  be  involved 
in  using  catch  data,  the  most  obvious  being 
that  the  data  are  restricted  largely  to  what  the 
fishermen  want  to  catch  when  they  want  to 
catch  it. 

Perhaps  more  important,  however,  are 
certain  problems  inherent  in  the  structure  of  the 


biological  models.  Biologists  distinguish  a 
number  of  types  of  biological  models,  among 
which  are  the  logistic  and  the  dynamic  pool- 
type. 

The  logistic  model  results  in  a  parabolic 
yield  curve  with  a  well-defined  maximum, 
and  this  has  been  utilized  by  many  economists. 
The  maximum  point  on  the  yield  curve  repre- 
sents the  maximum  sustainable  yield.10  At 
this  point  the  stock  is  roughly  half  as  abundant 
as  in  its  initial  state  or  maximum  size.  Two 
assumptions  of  interest  to  economists  lie  be- 
hind this  model,  the  first  "that  the  rate  of 
increase  in  the  stock  responds  immediately 
to  changes  in  population  density;  second, 
that  the  rate  of  natural  increase  at  a  given 
weight  of  stock  is  independent  of  its  age  (or 
size)  composition." 

Naturally  the  adequacy  of  the  assumptions 
and  the  intrusion  of  exogenous  forces  affect 
the  adequacy  of  the  model.  But  the  matter 
of  greatest  concern  to  the  economist  lies  in 
the  first  assumption.  If  there  is  not,  as  the 
biological  evaluation  of  this  type  of  model 
suggests  there  is  not,  an  instantaneous  ad- 
justment between  changes  in  population 
density  and  population  rate  of  growth  the 
economist  for  one  becomes  immediately  inter- 
ested in  the  time  dimension  of  the  adjustment 
mechanism  and  the  lag  function  that  may  be 
utilized  to  describe  it.11  Unless  we  limit  our- 
selves to  consideration  of  long  run  equilibrium 
solutions  the  integration  of  the  biological 
yield  function  into  an  economic  system  will 
require  the  specification  of  the  lag  function. 
For  many  purposes,  particularly  exposition, 
it  may  be  adequate  to  define  the  biological 
system  in  terms  of  equilibrium  points.  How- 
ever, if,  as  appears  to  be  the  case,  the  time 
lags  are  significant,  i.e.,  if  they  are  of  such 
duration  as  to  influence  economic  variables 
(price,    fishing    effort,    entry    and    exit),    then 


10  Holt,  (1962  p.  141-142),  has  suggested  that  this 
particular  function  may  be  flat  topped  which  "simply 
means  that  the  biological  facts  are  not  very  relevant 
to  determining  where  fishing  becomes  stabilized  over 
quite  a  range  of  variations  in  the  situation."  See 
alsoGulland  (1968). 

11  Prof.  G.  Paujik  has  pointed  out  to  me  that  the 
time  lags  involved  are  a  function  of  the  species  to 
which  the  model  is  applied.  In  general  tropical  species 
fit  the  first  assumption  fairly  well  but  those  in  temperate 
zones  show  much  slower  time  rates  of  adjustment. 


11 


equilibrium  models  of  this  type  lose  a  great 
deal  of  their  utility  as  a  basis  for  regulation. 
The  short  run  characteristics  of  the  economic 
adjustment  process  will  be  discussed  later  in 
this  paper  but  at  this  point  it  should  be  clear 
that  the  short  run  economic  and  biological 
adjustment  processes  are  inexorably  bound 
up  with  each  other. 

A  second  type  of  model,  distinguished  with- 
in the  biological  literature,  the  dynamic  pool, 
presents  an  even  thornier  set  of  difficulties 
for  fisheries  management.12  There  are  two  prob- 
lems inherent  in  dynamic  pool  models.  The  first 
is  that  the  maximum  sustainable  physical  yield 
is  defined  as  a  limit  that  can  be  reached  only 
by  the  expenditure  of  infinite  fishing  effort 
(infinite  cost).  The  second  difficulty  is  more 
analogous  to  those  found  in  trying  to  maximize 
the  net  economic  yield  from  the  resource.  Vari- 
ous degrees  of  overfishing  and  underfishing  are 
quantities  of  output  which  are  deviations  from 
the  eumetric  yield  curve  or  curve  of  best  yield, 
i.e.,  catching  the  appropriate  size  of  fish  for 
that  level  of  fishing  effort.  Deviations  from  the' 
eumetric  curve  are  controlled  by  making 
changes  in  the  selectivity  of  the  gear  utilized. 
The  necessary  conditions  for  making  these 
gear  adjustments  is  a  knowledge  of  the  con- 
dition of  the  stock  and  a  reasonable  degree  of 
flexibility  in  the  regulatory  process.  The  ab- 
sence of  an  operationally  definable  maximum 
sustainable  yield  plus  the  necessity  of  adjust- 
ing regulatory  technique  is  a  requirement  on 
management  that  is  similar  to  the  adjustments 
in  output  level  and  inputs  that  would  be  re- 
quired by  changes  in  price  and  cost  under 
economic  regulations. 


Forecasting  with  Biological  Models 

In  the  commercial  fisheries  the  forecasting 
problem  is  a  mix  of  the  complexity  of  the  life 
cycle  of  the  individual  species  and  the  avail- 
ability of  resources  to  carry  on  the  necessary 
biological    research    programs.    For    the    bulk 


of  the  populations  fished  the  effort  devoted 
to  biological  research  is  simply  insufficient 
to  provide  sophisticated  forecasts.  And  while 
at  first  glance  the  cure  for  this  particular  in- 
adequacy would  appear  to  be  simple  it  is  not. 
In  general  it  seems  unlikely  that  sufficient 
funds  for  meaningful  broad-based  biological 
programs  can  be  obtained  except  from  the 
income  generated  by  the  fisheries  themselves.13 
If  this  hypothesis  is  approximately  valid  then 
it  suggests  that  economic  rationalization 
(realization  of  the  potential  net  yield)  is  a 
necessary  condition  for  achieving  the  level  of 
funding  of  biological  research  sufficient  to 
allow  the  development  of  dependable  forecasts. 

The  type  of  forecast  made  depends  upon 
the  behavior  characteristics  of  the  species. 
The  Bristol  Bay  red  salmon  fishery  has  been 
studied  intensively  by  three  agencies;  the 
National  Marine  Fisheries  Service,  the  Alas- 
kan Department  of  Fish  and  Game,  and  the 
Fisheries  Research  Institute  of  the  University 
of  Washington. 

Table  1  presents  a  summary  of  salmon  runs 
and  a  rough  measure  of  the  accuracy  of  the 
forecasts  for  the  recent  decade.  Despite  the 
investment  in  research  and  the  heavy  payoff 
for  accurate  forecasts  in  the  Bristol  Bay 
fishery  it  is  clear  that  the  existing  forecasts 
are  not  completely  satisfactory.14  Furthermore, 
even  if  forecasting  in  this  fishery  was  100% 
accurate  the  instability  on  the  supply  side 
would  (does)  cause  severe  economic  problems. 

Few  other  species  present  the  forecasting 
problems  of  the  red  salmon.15  In  the  simpler 
cases  it  is  possible  to  estimate  the  stocks, 
and  then  assign  under  biological  regulations 
catch  limits  in  some  form.  In  subsequent 
time  periods  the  limits  may  be  adjusted  to 
allow   for  errors   in   estimation   of  stock   size. 


12  From  the  viewpoint  of  the  analysis  of  the  biological 
condition  of  the  stock  it  appears  to  have  certain  ad- 
vantages over  the  logistic  model. 


13  This  follows  from  the  common  property  status  of 
the  resource  which  means  that  the  rate  of  return  to 
the  firm  or  the  nation  on  investment,  in  research  is 
zero  in  long  run  equilibrium. 

14  Crutchfield  and  Pontecorvo,  (1969),  especially 
Chapter  7,  develop  the  rationale  for  the  high  payoff 
for  accurate  forecasts  in  Bristol  Bay. 

15  Schaefer  (1970a),  discusses  the  approach  to  the 
maximum  sustainable  yield  that  may  be  utilized  with 
species  such  as  the  Peruvian  anchoveta,  halibut,  etc. 
Even  within  these,  more  stable  populations  there  is 
room  for  substantial  disagreement  about  the  appropri- 
ate level  of  yield.  For  further  examples  see  (Schaefer, 
1967;  1970b  and  Segura,  1972). 


15 


Certain  deductions  can  be  made  about  the  bio- 
logical and  economic  implications  of  this  lagged 
response  approach  to  maximum  sustainable 
yield.  A  priori,  it  appears  that  its  economic 
viability  is  a  function  of  the  stability  of  the 
population.  And  since  we  have  already  sug- 
gested that  those  populations  which  are  fished 
heavily,  i.e.,  those  exploited  close  to  or  even 
beyond  the  maximum  sustainable  yield,  tend 
to  show  greater  fluctuations,  it  follows  that 
accurate  forecasts  of  short  run  supply  changes 
will  require  a  continuous  extensive  biological 
research  program.16 

Political  Science  and  Sociology 

The  discussion  thus  far  has  been  aimed 
at  understanding  the  nature  and  limitations 
of  the  maximum  physical  yield  as  a  biological 
construct  and  as  a  tool  in  fisheries  manage- 
ment. The  technical  side  of  the  problem  is 
however  just  one  part  of  it.  As  one  distin- 
guished fisheries  administrator  put  it: 

I  wish  to  inquire  whether  social  and 
political  problems  are  included  within  the 
scope  of  fisheries  economics.  If  so,  we  may 
be  able  to  arrive  at  a  fairly  broad  and  com- 
prehensive view  on  matters  of  fishery  regu- 
lation. If  not,  then  I  think  they  must  be  treated 
as  separate  aspects  (McHugh,  1962). 

Neither  the  economist  nor  the  biologist 
will,  based  on  what  can  be  learned  from  the 
individual  disciplines,  accept  responsibility 
for  the  social  and  political  problems  associ- 
ated with  the  fisheries.  They  have  both  been 
guilty  of  implying  that  social  and  political 
objectives  will  best  be  met  by  choosing  the 
alternative  they  espouse.  However,  since  social 
and  political  objectives  themselves  are  apt 
to  be  as  disparate  as  are  the  biological  and 
economic,  the  debaters  have  grasped  at  only 
those  aspects  of  social  and  political  policy 
that  have  best  fit  their  needs  at  the  moment. 


Table    1.    —    Run    of    sockeye    salmon    to   Bristol    Bay, 
1960-1970.* 


16  One  alternative  would  be  to  limit  fishing  effort 
to  that  sufficient  to  harvest  only  the  lower  portion  of 
the  range  of  variation  in  the  stock.  This  would  give 
a  small  output  at  low  cost  with  little  or  no  require- 
ment for  investment  in  biology.  Since,  however,  this 
would  be  a  disequilibrium  situation  with  long  run 
excess  profits  in  all  probability  it  could  not  be  sus- 
tained in  the  face  of  the  economic  pressures  to  expand. 


Millions  o 

ffish 

1960 

36.3 

1961 

18.0 

Range                          7.7-53.1 

1962 

10.4 

Median                             17.5 

1963 

6.8 

Mean                                20.8 

1964 

10.7 

Coefficient 

of  variation                       74% 

1965 

53.1 

1966 

17.5 

Approximate  forecasting 
errors: 

1967 

10.3 

1960-1970+40% 

1968 

7.7 

1969 

18.5 

Anticipated  accuracy  of 
forecasts  in  near  future: 

1970 

39.6 

±  20%  in  4  years  out  of  5 
±50%  in  5  th  year 


I  am  deeply  indebted  to  Dr.  Donald  E.  Rogers  of  the  Fisheries 
Research  Institute  of  the  University  of  Washington  for  assem- 
bling the  complex  data  on  the  runs  and  forecasts  of  Bristol 
Bay  and  Western  Alaska  Sockeye  (from  which  Table  1  is  ex- 
cerpted). As  noted  in  the  text,  the  purpose  in  presenting  these 
figures  is  to  emphasize  the  year-to-year  variations  in  supply. 


They  have  not  dealt  in  a  rigorous  analytical 
way  with  these  problems. 

In  this  it  seems  fair  to  say  that  the  biolo- 
gists have  been  the  political  realists  while 
the  economists,  to  the  extent  that  they  have 
dealt  with  the  question  of  labor  mobility, 
income  distribution,  and  the  impact  of  barriers 
to  entry  on  the  scale  of  enterprise,  have  been 
closer  to  social  realities. 

Economists  have  insisted  correctly,  in  my 
judgment,  in  their  discussion  of  the  problems 
of  fisheries,  that  the  general  economic  wel- 
fare of  the  state  and  the  individuals  in  it  are 
best  served  by  maximizing  the  net  economic 
yield  from  the  resource.  Their  occasional 
willingness  to  temporize  their  position  arises 
for  the  following  reason.  For  species  of  fish 
with  a  high  unit  value  such  as  lobster,  red 
salmon,  etc.,  the  discrepancy  between  the 
maximum  sustainable  physical  yield  and  the 
net  economic  yield  is  not  apt  to  be  very 
large,   the   former   being   a   second   best   solu- 


10 


tion  that  does  not  deviate  significantly  from 
the  economic  optimum.17 

Of  greater  significance  is  the  political  appeal 
of  the  idea  of  achieving  the  maximum  sus- 
tainable physical  yield.  The  economist  cate- 
gorically rejects  the  idea  that  it  is  "good"  to 
maximize  the  output  of  any  commodity  just 
because  it  is  physically  possible.  To  the  poli- 
tician negotiating  fisheries  agreements  both 
nationally  and  internationally  it  is  important 
to  be  able  to  state  that  the  agreement  makes 
possible  the  utilization  of  all  the  fish  avail- 
able for  all  time,  none  will  be  "wasted."  The 
simplistic  political  argument  runs  as  follows: 
the  production  of  food,  particularly  protein 
food,  is  good.  The  maximum  sustainable 
physical  yield  is  the  most  food  that  can  be 
obtained.  Any  other  definition  of  optimum 
output  such  as  the  net  economic  yield  would 
either  represent  less  food  (a  waste)  or  if  it 
was  greater  than  the  maximum  physical  yield 
it  would  be  a  threat  to  the  stock. 

The  danger  in  this  political  exposition  of 
the  problem  is,  of  course,  that  it  conceals  the 
underlying  complexities  of  the  biological  pro- 
cess as  well  as  the  interaction  between  those 
processes  and  economic  variables.18 

THE  NET  ECONOMIC  YIELD  AND  LONG 
RUN  PARTIAL  EQUILIBRIUM  MODELS 

Economic  and  Biological  Models 

As  a  first  approximation  we  may  assert 
that  the  utility  of  economic  models  in  fisheries 
management  is  symmetrical  with  the  biolog- 
ical counterparts  upon  which  they  rest.19  The 

17  There  is  for  the  noneconomist  a  possible  con- 
fusion here.  Both  the  output  that  will  maximize  the 
net  economic  yield  from  the  resource  and  the  output 
that  will  maximize  the  physical  yield  in  the  long  run 
are  points  derived  from  the  same  biological  yield 
function.  However,  the  maximization  of  the  net  eco- 
nomic yield  requires,  given  the  common  property  status 
of  fish  stocks,  an  economic  control  mechanism  as  well. 

18  -\ye  have  ignored  any  discussion  of  what  has 
been  referred  to  as  social  problems.  A  good  illustration 
of  the  interaction  of  all  the  forces  can  be  found  in  the 
Norwegian  coastal  fisheries.  For  reasons  that  are 
political,  social,  and  national,  the  Norwegian  govern- 
ment has  seen  fit  to  subsidize  coastal  fisheries.  These 
subsidies  have  been  indirect:  education  for  dependents, 
health  care,  transportation,  etc.,  and  direct:  price 
supports  for  raw  fish,  vessel  construction  subsidies, 
etc.  A  key  objective  of  this  policy  is  to  maintain  the 
population   living   along  the  coast  of  western   Norway, 


argument  developed  about  the  inadequacy 
of  partial  models  and  the  difficulty  of  utilizing 
long  run  equilibrium  systems  for  manage- 
ment decisions  are  applicable  to  both  biology 
and  economics. 

By  ignoring  underlying  definitional  prob- 
lems as  well  as  those  which  result  in  short 
run  fluctuations  in  output,  the  biological  con- 
cept of  the  maximum  physical  yield  can 
present  a  facade  of  stability.  No  such  facade 
exists  with  the  net  economic  yield.  The  ap- 
propriate level  of  output  is  defined  by  the 
interrelationship  between  market  price  and 
costs.  Since  the  price  of  most  fish  products 
is  determined  in  markets  that  are  describable 
as  workably  competitive,  it  is  clear  that  when 
the  physical  yield  function  is  transformed 
into  a  revenue  function  the  appropriate  level 
of  output  will  shift  in  response  to  price 
changes.20 

The  voluminous  literature  on  the  economics 
of  uncertainty  is  suggestive  of  the  magnitude 
of  the  administrative  and  political  problems 
it  creates  for  the  regulatory  mechanism,  and 
also  the  lengths  administrators  will  go  to 
minimize  it.  But  how  much  uncertainty  actu- 
ally would  be  created  by  the  imposition  of 
regulatory  practices  aimed  at  maximizing 
the  net  economic  yield?  The  major  sources 
of  disequilibrium  in  most  fisheries  appear  to 
be  attributable  to  two  forces,  short  run  vari- 
ations in  the  supply  of  fish  and  shifts  in  the 
demand  function.  In  this  context  it  is  impor- 
tant to  distinguish  between  stability  in  the 
demand  function  and  stability  of  price.  Within 
the  economic  models  price  changes  are  caused 
by  short  run  variations  in  supply  which  cause 


the  traditional  farmer  fisherman.  Other  considerations 
are  military,  economic  (balance  of  payment),  and 
political  in  that  it  is  important  to  participate  in  the 
exploitation  of  stocks  as  a  claim  against  any  future 
regulation  that  might  involve  national  quotas,  etc. 

19  There  are  of  course  important  differences.  For  an 
exposition  of  certain  properties  of  variable  propor- 
tions diminishing  returns  unique  to  the  fisheries,  see 
F.  W.  Bell,  and  E.W.Carlson  (1970). 

20  See  Crutchfield  and  Pontecorvo,  (1969,  pp.  28- 
88).  The  usual  assumption  is  that  cost  functions  are 
linear  and  stable.  This  follows  from  the  size  of  labor 
markets,  the  general  availability  of  the  type  of  capital 
instruments  required  in  most  fisheries,  and  especially 
from  the  small  scale  of  most  fisheries  encompassed  by 
the    partial    equilibrium    systems    analyzed. 


17 


a  change  in  price  and  are  therefore  the  pri- 
mary source  of  uncertainty.  This  effect  will 
be  dampened  if  the  demand  function  is  highly 
elastic,  a  condition  that  seems  applicable  to 
many  fisheries. 

In  the  long  run,  however,  the  situation  is 
different.  The  dynamics  of  short  run  supply 
changes  continue  through  time  but  in  the 
long  run  the  demand  function  tends  to  be 
responsive  to  income  changes  and  therefore 
shifts  to  the  right,  adding  to  the  degree  of 
uncertainty. 

Given  the  high  level  of  uncertainty  inherent  in 
fisheries,  the  workability  of  fisheries  as  in- 
dustries is  highly  dependent  on  their  economic 
efficiency,  i.e..  their  ability  to  operate  profit- 
ably and  not  dissipate  the  rent  from  the  re- 
source among  redundant  inputs. 


Sources  of  Short  Run  Economic  Instability 

If,  for  the  moment,  however,  we  limit  our 
argument  to  the  short  run  and  we  hypothe- 
size that  the  source  of  instability  is  on  the 
supply  side  then  it  is  correct  to  say  that 
regulations  aimed  at  either  maximizing  phys- 
ical yield  or  net  revenue  are  not  significantly 
different  from  each  other  in  terms  of  the  level 
of  uncertainty  involved.  An  analogy  may  be 
useful  in  putting  this  problem  in  better  per- 
spective. 

Academic  economists  are  virtually  unani- 
mous in  the  opinion  that  some  degree  of  flexi- 
bility in  exchange  rates  is  desirable.  Central 
bankers  and  to  a  lesser  extent  businessmen 
are  generally  opposed  to  the  creation  of  the 
uncertainty  that  flexible  rates  would  bring. 
At  the  heart  of  this  debate  are  two  different 
views  of  the  stability  of  the  underlying  system. 
If  equilibrium  not  disequilibrium  is  the  norm 
then  problems  involving  the  short  run  adjust- 
ment process  are  minimal.  But  if  short  run 
variations  are  inherent  and  important  in  'the 
system  then  the  regulatory  process  must  be 
flexible  to  be  consistent  with  the  dynamics  of 
the  short  run. 

The  short  run  economic  adjustment  process 
in  fisheries,  unlike  certain  industries,  is  par- 
ticularly responsive  to  changes  in  market  con- 
ditions. Common  property  and  easy  entry,  a 
relatively  low  level  of  specialization  of  inputs, 


and  the  possibility  of  shifts  of  economic  units 
between  fisheries,  have  combined  to  create 
this  condition. 

Let  us  define  the  short  run  to  be  a  period 
sufficiently  brief  to  exclude  new  entry.  By 
new  entry  we  mean  that  fishing  effort  would 
be  carried  on  by  units  of  capital  and  labor 
with  no  previous  experience  in  the  fishery 
in  question.  Fisheries  (except  in  initial  growth 
stages  or  periods  of  significant  technological 
change)  tend  to  be  characterized  by  excess 
capacity.  In  these  circumstances  short  run 
shifts  in  the  price/cost  ratio  brought  about 
by  changes  in  either  supply  or  demand  can 
generate  wide  swings  in  fishing  effort.  This 
effort  may  come  from  greater  productivity 
(longer  hours,  better  organization,  harder 
work),  by  the  activation  of  units  previously 
participating  in  this  fishery  but  currently  "on 
the  beach,"  or  by  the  response  of  economic 
units  that  work  in  several  fisheries  on  a  part- 
time  basis  to  the  enhanced  profit  position  in 
this  one.21  If  the  excess  profits  continue  in 
the  face  of  the  increase  in  fishing  effort,  new 
entry  will  take  place  fairly  rapidly. 

The  labor  and  capital  utilized  in  many 
fisheries  may  be  characterized  as  having  three 
elements,  one  is  a  core  of  labor  and  capital 
that  is  primarily  identified  with  the  particu- 
lar fishery  in  question  and  this  core  may  ex- 
pand and  contract  its  efforts  in  response  to 
market  conditions  but  it  lacks  the  mobility 
required  to  shift  rapidly  to  alternative  fisheries. 

The  second  element  is  a  stock  of  standby 
capacity,  either  currently  employed  elsewhere 
or  unemployed,  that  can  and  does  respond  to 
changes  in  profit  prospects.  These  two  com- 
ponents, possibly  each  individually  if  restric- 
tions on  productivity  are  considered,  represent 
more  capacity  than  is  needed  to  harvest  any 
average  level  of  catch  and  perhaps  even  more 
capacity  than  is  required  to  land  the  upper 
limits  of  the  frequency  distribution  of  the 
abundance  of  the  stock. 


21  These  responses  are  not  symmetrical,  i.e.,  fishing 
effort  increases  more  rapidly  in  response  to  profit  op- 
portunities than  does  exit  to  a  reduction  in  earnings. 
Inertia,  the  possibilities  of  windfall  gains  and  the 
inadequacy  of  the  forecasts  of  short  run  supply  all 
contribute  to  the  asymmetry.  It  is  particularly  im- 
portant to  note  that  we  have  not  mentioned  changes 
in  technology.  In  many  fisheries  the  relationship  be- 
tween technological  change  and  fishing  effort  is  cir- 
cumscribed by  the  regulatory  process. 


IK 


The  third  element  is  the  continuous  threat 
of  entry.  If  the  excess  profits  observed  in  one 
time  period  continue,  or  if  they  are  expected 
to  continue,  entry  will  take  place.  Expecta- 
tions and  the  competitive  illusion  play  a  role 
in  all  industries  but  the  great  flexibility  in 
the  capital  instruments  employed  in  fisheries 
tend  to  make  the  interaction  between  market 
conditions,  expectations,  and  capacity  par- 
ticularly close  (Pontecorvo  and  Vartdal,  1967). 

Supply  fluctuations,  excess  capacity,  the 
rapidity  of  responsiveness  to  changes  in  the 
market,  and  the  influence  of  expectations  all 
contribute  to  short  run  instability  in  fisheries. 
Control  of  capacity,  improvements  in  fore- 
casting supply  in  order  to  reduce  uncertainty, 
plus  recognition  that  capacity  sufficient  to 
capture  some  average  level  of  catch  less  than 
the  maximum  sustainable  yield  may  be  ap- 
propriate, are  all  elements  in  a  management 
program  geared  to  meeting  the  conditions  im- 
posed by  the  short  run  dynamics  of  fisheries. 

Long  Run  Equilibrium 

If  short  run  economic  objectives  are  defin- 
able in  the  terms  indicated  above  it  is  ap- 
propriate to  inquire  next  about  the  long  run 
equilibrium  conditions.  Economic  analysis  of 
fisheries  has  accepted  as  given  the  biological 
yield  function  for  the  species  in  question,  as 
well  as  the  usual  assumptions  of  static  equi- 
librium analysis  of  full  employment  and  factor 
mobility.  In  these  circumstances  the  condition 
of  Pareto  optimality  is  roughly  fulfilled  if  the 
policy  recommendations  (essentially  creation 
of  a  set  of  regulations  aimed  at  maximizing 
the  rent  of  the  resource  and  in  all  probability 
requiring  barriers  to  entry)  required  to  ration- 
alize the  fishery  are  met.  Within  the  frame- 
work of  economic  analysis  (maximization  of 
Gross  National  Product)  this  is  a  necessary 
and  sufficient  condition  for  making  the  maxi- 
mization of  the  net  economic  yield  the  ap- 
propriate goal  of  fisheries  management.  Any 
alternative  is  less  satisfactory  in  that  it  will 
result  in  a  lower  level  of  material  well-being 
(GNP).22 

22  A  crucial  assumption  is  that  the  opportunity 
cost  for  labor  is  positive.  Most  of  the  attacks  on  the 
concept  of  economic  regulation  of  fisheries  assert  the 
contrary.  Perhaps  the  needed  empirical  investigation 
of  this  point  could  start  with  a  classification  such  as 
suggested  in  Approaches  to  Fisheries  Management. 


Attacks  on  this  goal  have  come  from  two 
sources,  biologists  and  fisheries  administra- 
tors, and  also  from  within  the  economics  pro- 
fession. The  position  of  the  former  group  rests 
in  large  part,  in  my  opinion,  on  a  funda- 
mental misconception  concerning  the  mean- 
ing of  economic  optimization.  Economics  is 
not  sufficient  to  explain  (or  optimize),  par- 
ticularly in  the  short  run,  the  entire  set  of 
variables  involved  in  a  fishery.  The  economist 
accounts  for  the  social  problems  by  his  as- 
sumption of  full  employment  and  factor  mo- 
bility. He  does  not  normally  account  for  po- 
litical factors  except  indirectly  in  his  under- 
lying assumption  of  human  rationality  which 
tends  to  force  the  political  preferences  into 
the  same  ordering  as  the  economic. 

A  bioeconomic  position  dominates  thinking 
about  fisheries  management  simply  because 
there  is  no  body  of  social  or  political  theory 
sufficiently  powerful  (relative  to  welfare  maxi- 
mization in  economics  or  population  dy- 
namics in  biology)  to  force  a  modification  of 
either  the  biological  or  economic  position. 
In  these  circumstances,  which  appear  unlike- 
ly to  change  in  the  foreseeable  future,  political 
and  social  considerations  can  only  be  con- 
sidered on  an  ad  hoc  basis.  More  specifically, 
it  is  normally  true  that  the  biological  optimum 
and  economic  optimum  are  consistent  with 
each  other  in  that  both  will  protect  the  stock. 
The  economic  goal  is  more  general  and  there- 
fore preferable  in  that  in  addition  to  protect- 
ing the  stock  it  also  provides  the  maximum 
economic  benefits  to  society.  Any  deviation 
from  the  economic  maximum  involves  there- 
fore a  cost,  a  cost  measurable  in  terms  of 
output  foregone.  Nothing  in  this  argument 
suggests  that  ad  hoc  reasons  are  not  sufficient 
grounds  (given  the  weakness  in  the  two  under- 
lying assumptions  in  certain  circumstances) 
to  make  an  alternative  objective  (political, 
military,  social,  etc.)  either  the  primary  or  a 
subsidiary  goal  of  fishery  management. 

In  this  circumstance  the  economist's  primary 
concern  would  be  to  calculate  the  cost  of  the 
alternative.  The  latter  calculation  presupposes 
that  the  alternative  can  be  specified,  a  con- 
dition that  is  seldom  met.  What  tends  to 
emerge  as  the  management  goals  in  fisheries 
under  the  long  run  equilibrium  condition  that 
dominates   today's   thinking  is   an   unspecified 


19 


mix  of  all  factors  including  purely  adminis- 
trative considerations. 

The  attack  within  the  profession  has  raised 
an  appropriate  question  about  the  implica- 
tions of  limits  on  entry  for  Pareto  optimiza- 
tion. Another  position  also  has  been  advanced 
based  on  an  assumption  questioned  through- 
out this  paper  that  the  maximum  physical 
yield  is  so  fundamentally  sound  in  the  oper- 
ational sense  that  its  utility  as  a  tool  out- 
weighs its  defects.23 

In  the  competitive  model  there  is  no  limit 
on  entry  beyond  that  provided  by  what  Knight 
has  called  the  "social  function  of  ownership." 
In  the  fishery  with  the  resource  being  common 
property  the  objective  of  the  economist  in 
calling  for  a  limit  on  entry  is  to  provide  the 
ownership  function  while  retaining  the  Pareto 
optimum  conditions  inherent  in  the  competi- 
tive model. 

Two  questions  may  be  raised  about  this 
goal  and  the  procedures  necessary  to  achieve 
it.  Will  the  barriers  to  entry  result  in  a  situ- 
ation that  goes  beyond  competition,  i.e.,  does 
the  creation  of  property  rights  just  restore 
the  conditions  that  would  be  found  with  pri- 
vate property  operating  under  competition 
or  does  it  also  imply  the  creation  of  monopoly 
power  so  that  buying  and  selling  is  no  longer 
on  a  competitive  basis?  The  second  question 
is  an  integral  part  of  the  first.  Does  the  estab- 
lishment of  barriers  to  entry  and  the  subse- 
quent economic  regulation  of  the  fishery  in 
the  public  interest  require  the  creation  of  a 
regulatory  mechanism  so  costly  and  complex 
as  to  be  self  defeating?24 

Economic     theory     does     not     provide     an 


23  See  Wantrup  (1970,  p.  18):  "While  maximum  sus- 
tainable yield  constitutes  a  relevant,  operational,  and 
noncontroversial  objective  of  conservation  policy,  this 
is  quite  different  for  the  objective  of  'maximum  net 
economic  yield'  •  even  if  its  realization  through 
limitation  of  entry  could  be  agreed  upon  by  the  fishing 
industry."  Also  (Wantrup,  1962,  p.  292):  "My  approach, 
therefore,  would  be  set  more  modest  regulation  goals 
which  would  concern  themselves  more  with  the  re- 
source base  than  with  rent.  We  are  dealing  then 
with  matters  we  can  measure.  If  we  try  to  maximize 
rent  as  a  policy  goal,  then  we  get  into  an  area  where 
I  for  one  would  put  out  a  'caveat'  sign." 

24  Virtually  the  entire  literature  on  the  economics 
of  fisheries  has  commented  on  these  two  questions 
albeit  in  a  not  very  satisfactory  manner.  For  the 
details  on  control  plans  to  limit  entry  see  Sinclair 
(1962)  and  Royce  et  al.  (1963). 


answer  to  these  questions.  It  is  possible,  how- 
ever, based  on  experience  with  the  social  con- 
trol of  industry,  to  advance  certain  tentative 
hypotheses.  No  control  mechanism  based  solely 
on  biological  considerations  is  workable  in 
the  long  run  in  the  face  of  economic  and 
other  pressures.  Therefore,  the  cost  of  the 
control  mechanism  is  ultimately  a  joint  bio- 
economic  cost.  Even  in  these  circumstances 
it  may  be  that  the  costs  of  control  are  dis- 
proportionately large  relative  to  the  value  of 
the  resource  left  unprotected.25 

A  second  hypothesis  is  that  if  the  cost  of 
bioeconomic  control  is  not  excessive  then  the 
capacity  to  regulate  the  fishery  in  the  public 
interest,  i.e.,  to  preserve  the  stock  and  prevent 
the  emergence  of  significant  monopoly  power, 
is  well  within  the  power  of  regulatory  pro- 
cesses and  tax  arrangements  that  have  proved 
themselves   workable   in   other   circumstances. 

APPROACHES  TO  FISHERIES 
MANAGEMENT 

A  major  constraint  in  the  development  of 
a  consistent  monetary  policy  is  that  the  mone- 
tary system  itself  is  continuously  evolving. 
The  analogy  seems  applicable  to  regulatory 
problems  involving  ocean  resources.  The  pat- 
tern of  resource  exploitation  in  the  oceans 
and  the  law  of  the  sea  are  changing  rapidly. 
In  addition,  military  uses  of  the  oceans,  while 
not  a  new  phenomenon,  are  being  transformed 
and  at  the  same  time  the  very  existence  of 
the  ocean  in  the  way  we  have  known  it  is 
threatened  by  the  effects  of  the  population  ex- 
plosion and  the  rising  level  of  real  income. 
Furthermore,  our  capacity  to  deal  effectively 
with  ocean  living  resource  problems  is  limited 
by  the  inadequacy  of  our  scientific  knowledge 
of  life  processes  in  the  ocean,  the  generally 
weak  economic  condition  of  fisheries,  and  the 
nationalistic  interests  involved. 

It  is  beyond  the  scope  of  this  paper  to  go 
into  these  structural  questions.  It  is  clear, 
however,  that  in  the  future  the  organizations 
by  which  fisheries  are  to  be  regulated  must 
be  prepared  to  negotiate  the  basic  issues  of 
control  of  the  environment  and  the  priorities 


25  It  might  be  desirable  to  protect  the  resource  for 
other  reasons,  i.e.,  because  it  was  unique,  etc. 


liO 


appropriate  to  multiple  use  situations  in  the 
oceans  with  external  forces.  In  these  activities 
the  strength  of  their  bargaining  position  will 
depend  heavily  on  their  having  rationalized 
both  the'  economic  and  biological  sides  of  the 
fisheries.  Regardless  of  what  potential  eco- 
nomic yields  may  be  or  what  social  pressure 
for  employment  is  present,  a  realized  net 
economic  yield  of  zero  from  fishing  does  not 
provide  an  adequate  base  for  defending  ocean 
space  for  commerical  fisheries  in  competition 
with  the  oil  industry,  recreational  use,  power 
generation,  etc. 

In  recent  years  progress  has  been  made 
by  governments,  fisheries  commissions,  and 
academic  researchers  in  the  analysis  of  fish- 
eries problems.  What  are  the  elements  of  this 
analysis  that  may  be  utilized  to  help  reorient 
our  approach  to  fisheries  management? 

The  long  run  partial  equilibrium  systems 
constructed  thus  far  make  a  major  contribu- 
tion by  an  exposition  of  the  problems  in  static 
terms.  It  is  clear,  however,  that  they  are  in- 
adequate for  resource  management.  In  most 
circumstances  they  do  provide  limits  within 
which  the  regulatory  process  may  operate. 
In  the  analysis  of  particular  species  the  dis- 
tinction between  the  net  economic  yield  and 
the  maximum  sustainable  physical  yield  is 
subject  to  empirical  verification  depending  on 
the  unit  value  of  the  species,  but  in  any  case 
it  is  a  second  order  question.  In  any  set  of 
priorities  established  for  fisheries  manage- 
ment the  first  is  to  move  toward  meeting  the 
criteria  of  economic  efficiency,  probably  by 
establishing  limits  on  entry.  Once  the  fishery 
is  rationalized  then  the  solution  to  the  prob- 
lem of  the  appropriate  level  of  output  should 
be  greatly  simplified.  This  follows  from  the 
nature  of  the  adjustments  that  must  be  made 
in  the  process  of  economic  rationalization. 
The  administrators  will  be  forced  to  consider 
simultaneously  the  appropriate  amount  of 
fishing  effort  (amount  of  inputs)  relative  to 
the  forecast  of  the  frequency  distribution  of 
supply  and  the  impact  of  productivity  changes 
on  fishing  effort.  Once  the  fishery  is  defined 
in  this  way,  the  economic  implications  and 
advantages  of  various  levels  of  output  will 
be  more  apparent  to  all  and  self  interest, 
which  today  drives  producers  toward  over- 
fishing the  resource,  will   move  them  toward 


limiting  the  catch  to  maximize  the  net  yield 
from  the  resource.26 

Given  recognition  of  the  long  run  bio- 
economic  limits,  the  adequacy  of  the  regula- 
tory mechanism  may  be  evaluated  in  terms 
of  how  well  it  handles  the  short  run  maxim- 
ization problem  and  its  success  in  restructur- 
ing the  fishery  from  a  disequilibrium  position 
(excess  capacity)  to  one  of  equilibrium.  This 
latter  will  require  evaluation  of  the  possibil- 
ities inherent  in  aggregative  yield  functions 
and  clarification  of  the  goal  of  a  workably 
competitive  structure  for  the  fisheries.  The 
rents  captured  in  the  rationalization  process 
are  available  to  finance  this  transition.  In 
these  Utopian  circumstances  the  essence  of 
the  internal  regulatory  mechanism  will  be 
found  in  the  interaction  between  changes  in 
biological  supply,  prices,  and  technology. 

LITERATURE  CITED 

BELL,  F.W.,  and  E.  W.  CARLSON,  1970.  The  Pro- 
ductivity of  the  Sea  and  Malthusian  Scarcity.  Work- 
ing Paper  Number  48,  National  Marine  Fisheries 
Service.  Draft  Manuscript. 

CRUTCHFIELD,  JAMES  A.,  and  GIULIO  PONTE- 
CORVO,  1969.  The  Pacific  Salmon  Fishery:  A  Study 
of  Irrational  Conservation.  Published  for  Resources 
for  the  Future,  Inc.,  The  Johns  Hopkins  Press,  Balti- 
more. 

GULLAND,  J.  A.,  1968.  Population  Dynamics  of  the 
Peruvian  Anchoveta.  FAO  Fisheries  Technical  Paper 
Number  72.  Rome. 

HOLT,  S.  J.,  1962.  Comments  made  in  discussion 
of  Dickie,  L.  M.  Effects  of  Fishery  Regulations  on 
the  Catch  of  Fish.  In:  Economic  Effects  of  Fishery 
Regulation,  R.  Hamlich,  ed.  FAO  Fishery  Report 
Number  5.  Rome.  pp.  141-142. 

McHUGH,  J.  L.,  1962.  Comments  made  in  discussion 
of    Dickie,    L.    M.    Effects    of    Fishery    Regulations    on 


26  This  statement  is  more  than  a  pious  hope  but 
less  than  a  certainty.  Its  validity  depends  in  part  on 
the  nature  of  the  frequency  distribution  of  catch  among 
the  participants  in  the  fishery,  i.e.,  if  the  fishery  were 
the  property  of  a  monopolist  he  would  operate  at  the 
level  of  the  net  economic  yield.  Only  under  certain 
assumptions  will  this  be  true  of  the  behavior  of  a  set 
of  competitors.  Their  recognition  of  the  desirability 
of  maximizing  aggregate  net  revenue  will  come  as  a 
process  of  education  as  studies  of  the  characteristics 
of  the  fishery  reveal  the  advantages  inherent  in  various 
alternatives. 


21 


the  Catch  of  Fish.  In:  Economic  Effects  of  Fishery 
Regulation,  R.  Hamlich,  ed.  FAO  Fishery  Report 
Number  5.  Rome.  p.  147. 


SEGURA,  EDILBERTO  E.,  1972.  Optimal  Fishing- 
Effort  in  the  Peruvian  Anchoveta  Fishery.  This  publi- 
cation. 


POXTECORVO.  GIULIO,  and  K.  VARTDAL,  JR., 
1967.  Optimizing  Resource  Use:  The  Norwegian 
Winter  Herring  Fishery.  StatsjeSkonomisk  Tidsskrift 
Number  2. 


SINCLAIR,  SOL,  1962.  License  Limitation  —  British 
Columbia.  In:  Economic  Effects  of  Fishery  Regula- 
tion, E.  Hamlich,  ed.  FAO  Fishery  Report  Number  5. 
Rome.  pp.  306-328. 


ROYCE.  W.,  et  al.,  1963.  Salmon  Gear  Limitation  in 
Northern  Washington  Waters.  University  of  Washing- 
ton Publication  in  Fisheries.  New  Series.  11(1).  Seattle. 

SCHAEFER,  M.  B..  1967.  Dynamics  of  the  Fishery  for 
the  Anchoveta  off  Peru.  Boletin,  Instituto  del  Mar 
del  Peru,  Callao. 

SCHAEFER,  M.  B.,  1970a.  Investigation,  Conserva- 
tion and  Management  of  the  Fisheries  of  the  High 
Seas.  Paper  presented  at  the  Preparatory  Conference 
on  Ecology  and  Science  Policy,  April  20-26.  The 
Center  for  the  Study  of  Democratic  Institutions,  Santa 
Barbara,  California. 

SCHAEFER,  M.  B.,  1970b.  Men,  Birds  and  Anchovies 
in  the  Peru  Current  —  Dynamic  Considerations.  Trans- 
actions of  the  American  Fisheries  Society.  99(3).  p.  461. 


STEWART,  C.  S.,  and  GIULIO  PONTECORVO,  1970. 
Problems  of  Resource  Exploitation:  The  Oil  and  Fish- 
ing Industries.  Chapter  I  in  Ocean  Enterprises,  The 
Center  for  the  Study  of  Democratic  Institutions,  Santa 
Barbara,  California. 

WANTRUP,  S.  V.  CIRIACY,  1962.  Comments  made 
in  discussion  of  Pontecorvo,  Giulio,  Regulation  of 
the  North  American  Lobster  Fishery.  In:  Economic 
Effects  of  Fishery  Regulation,  R.  Hamlich,  ed.  FAO 
Fishery  Report  Number  5.  Rome.  p.  292. 

WANTRUP,  S.  V.  CIRIACY,  1970.  The  Economics 
of  Environmental  Policy.  Paper  presented  at  the 
Preparatory  Conference  on  Ecology  and  Science 
Policy,  April  20-26.  The  Center  for  the  Study  of  Demo- 
cratic Institutions,  Santa  Barbara,  California. 


22 


Multiple  Objectives  for  Marine  Resource  Management1 


R.  Bruce  Rettig2 


ABSTRACT 

Management  decisions  suggested  by  recent  bioeconomic  models  have  been  largely 
disregarded  by  fishery  managers.  This  negligible  impact  may  be  due  to  error  on  the 
part  of  management,  an  incomplete  grasp  of  the  role  of  noneconomic  objectives,  and/or 
the  possibility  that  more  sophisticated  economic  models  might  yield  markedly  different 
results.  More  sophisticated  models  are  suggested  which  consider  the  problem  of  second 
best,  risk  and  uncertainty,  transaction  and  adjustment  costs,  and  income  redistribution. 
Creation  of  analytical  systems  amenable  to  treatment  of  noneconomic  variables  along 
with  economic  variables  is  suggested. 


During  the  past  two  decades,  a  growing  body 
of  economists  has  been  articulating  a  rationale 
for  management  of  ocean  fisheries  which  is 
based  upon  the  principle  of  maximum  sustain- 
able net  economic  yield.  The  usual  paradigm 
emphasizes  the  lack  of  clearly  denned  property 
rights  and  arrives  at  a  conclusion  of  a  need  for 
limited  entry,  most  commonly  suggested  through 
a  system  of  licensing  and/or  taxes.  While  the 
better  analyses  have  often  hedged  their  con- 
clusions with  a  set  of  qualifications,  even  these 
balanced  policy  programs  are  rejected  by 
authorities  actually  responsible  for  fishery 
management. 

That  articulate  arguments  from  a  respected 
cross-section  of  the  economics  profession  con- 
tinue to  carry  only  minor  weight  with  their 
intended  audience  is  quite  disconcerting.  This 
paper  consists  of  an  examination  of  two  possible 
reasons  for  the  treatment  of  the  bioeconomic 
models  to  date.  The  first  possibility  is  a  diver- 
gence between  what  public  authorities  consider 
appropriate  objectives  to  pursue  and  the  assump- 
tions of  goals  implicit  or  explicit  in  the  bio- 
economic models.  The  second  possibility  is  that 
bioeconomic  models  are  incompletely  specified 
and  that  more  complete  models  would  be  better 
received.  A  third  possibility  that  will  not  be 
handled  in  this  paper  is  that  existing  analysis 
is    correct    and    that    all    that    remains    is    to 


1  Oregon  Agriculture  Experiment  Station,  Technical 
Paper  2996.  Research  for  this  paper  was  supported  by 
N.S.F.  Institutional  Sea  Grant.  GH-45. 

2  Department  of  Agricultural  Economics,  Oregon 
State  University. 


educate  the  resource  managers  on  the  merits  of 
implementing  the  correct  suggestions  already 
available. 


THE  ELUCIDATION  AND 
LEGITIMIZATION  OF  SOCIAL  GOALS 

The  characteristic  of  the  fishery  which  lies 
at  the  root  of  the  problem  is  the  lack  of  clearly 
defined  property  rights  over  the  fishing  ground. 
The  severe  depletion  of  such  fisheries  as  the 
Pacific  halibut  fishery  and  the  sardine  fishery 
off  the  California  coast  stand  as  stark  testimony 
to  the  value  of  property  rights.  That  the  problems 
associated  with  fisheries  can  be  easily  related  to 
the  absence  of  property  rights  is  seen  by  con- 
sidering the  central  economic  functions  of 
property  as  set  forth  recently  by  Bjork  (1969, 
p.  65): 

First,  it  provides  incentives  for  the  creation  and 
improvement  of  assets.  Second,  it  provides  incen- 
tives for  efficient  control  of  existing  assets.  Third, 
it  rations  the  use  of  scarce  assets  to  ensure  that 
they  will  be  used  for  those  purposes  which  society 
values  most  highly. 

Bjork  argues  that  property  rights  exist  largely 
because  stable  market-oriented  societies  value 
the  performance  of  these  functions  so  highly. 

Investment  incentives,  efficient  allocation  of 
fishing  effort  and  living  fish,  and  distribution  to 
him  who  values  the  resource  most  highly  are 
indeed  the  central  objectives  behind  the  bio- 
economic models  of  current  interest.  These  are 
not  apparently  the  sole  objectives  of  the  societies 


23 


whose  mandates  the  resource  managers  must 
have  in  order  to  perform  their  duties  viably.  It 
is  useful  to  lump  these  other  objectives  into 
two  categories  —  equity  in  the  distribution  of 
income,  and  noneconomic  objectives.  Equity  of 
income  distribution  is  relevant  both  in  con- 
straining changes  from  the  status  quo  and  in 
defining  acceptable  distributions  of  income.  This 
latter  can  be  illustrated  by  the  popularity 
among  some  authors  of  giving  some  supra- 
national agency  control  over  all  marine  resources 
and  by  current  claims  of  U.S.  nationals  to  the 
natural  right  of  the  citizens  of  a  country  over 
fish  which  swim  over  "their"  continental  shelf. 
Noneconomic  objectives  are  important  in  the 
arguments  of  some  concerning  the  inherent  evil 
of  blocking  entry  to  a  fishing  ground.  Of  course, 
several  objectives  can  be  classified  more  than 
one  way.  The  repercussions  of  sudden,  unexpected 
unemployment  of  fishermen  can  be  called  adjust- 
ment costs  or  could  be  appropriately  titled 
sociological  phenomena. 

In  any  case,  it  is  not  appropriate  for  econo- 
mists to  identify  conclusions  of  their  positive 
models  with  normative  policy  proposals.  Society 
is  not  composed  of  economic  men  (Boulding, 
1969).  Rather  the  economist  must  first  try  to 
maximize  economic  gain  subject  to  noneconomic 
constraints.  However,  when  noneconomic  ob- 
jectives are  not  postulated  at  unique  target 
levels,  the  tradeoffs  between  economic  and  non- 
economic  objectives  must  be  considered.  I  will 
return  to  this  practical  problem  after  consider- 
ing some  problems  of  analyzing  economic 
objectives. 


SECOND-BEST  FISHERIES,  OR 
WHEN  IS  AN  OPTIMUM  NOT  OPTIMAL? 

The  application  of  the  theory  of  common 
property  resources  to  ocean  fisheries  leads 
inexorably  to  two  conclusions.  First,  it  is 
possible  to  observe  an  allocation  of  human  and 
capital  resources  in  fishing  with  social  marginal 
products  which  are  negative.  Second,  the  optimal 
allocation  of  resources  is  one  in  which  social 
marginal  revenue  product  equals  social  oppor- 
tunity cost  of  factors  used  to  catch  fish.  This 
is  to  say  that  less  fishing  effort  should  be 
employed  than  would  be  required  to  harvest 
the    maximum    sustainable    yield    unless    the 


alternative   use   of  the   marginal   resources   is 
valueless. 

Crutchfield  and  Pontecorvo  (1969,  p.  35)  have 
pointed  out  that  the  need  for  intervention  in 
fishery  management  hinges  upon  the  assump- 
tion of  competitive  behavior  by  the  downstream 
purchasers  of  the  resource. 


A  monopsonist  would  impose  a  rational  solution 
on  the  fishery,  i.e.,  he  would  capture  the  rent  by 
offering  sellers  a  price  that  would  permit  only  the 
most  efficient  exploitation  of  the  resource  to  take 
place,  and  the  malallocation  of  resources,  which 
results  from  the  combination  of  free  entry  and 
common  property,  would  be  avoided.  If,  in  turn,  the 
product  market  in  which  he  sells  is  highly  competi- 
tive, monopsony  could  provide  a  near-optimal  level 
of  output  and  real  costs. 


While  Crutchfield  and  Pontecorvo  go  on  to 
point  out  that  the  industry  which  purchases 
fish  from  fishermen  is  a  competitive  oligopsony 
and  does  not  lead  to  a  socially  optimal  solution, 
their  argument  still  holds  qualitatively.  When 
an  -oligopsonistic  industry  faces  a  group  of 
competitive  sellers  the  price  paid  for  the  output 
of  the  competitive  sellers  is  less  than  the  social 
valuation  for  an  incremental  unit,  i.e.,  the  social 
marginal  revenue  product  exceeds  the  market 
price.  This  divergence  alters  the  optimal  inter- 
vention in  the  market  structure  of  fish  buyers 
and  alters  the  optimal  fishery  management 
scheme,  assuming  control  over  the  two  cannot 
be  coordinated. 

If  the  market  between  fishermen  and  fish 
processors  is  not  perfectly  competitive,  the 
correction  of  fisheries  resource  allocation  ignor- 
ing this  fact  could  actually  misallocate  resources. 
Equating  the  social  marginal  product  of  fishing 
effort  multiplied  by  observed  market  price  to 
social  opportunity  cost  of  factors  used  to  catch 
fish  would  secure  a  level  of  fishing  effort  less 
than  the  one  where  the  true  value  of  social 
marginal  product  equals  social  opportunity  cost 
of  factors.  If  the  demand  for  fish  is  elastic 
throughout  the  relevant  range,  the  correct 
solution  would  still  occur  prior  to  attainment 
of  maximum  sustainable  physical  yield.  The 
simple  maximum  sustainable  physical  yield 
criterion  is  still  in  error  and  the  error  is  still  in 
the  same  direction,  but  the  error  is  smaller  than 
the  simpler  analysis.  Thus  the  efficiency  loss 
from  using  the  physical  rule  is  smaller  than 
previous  analysis  has  suggested. 


24 


On  the  other  hand,  if  a  social  suboptimum 
does  not  exist  in  an  ocean  fishery,  the  dis- 
solution of  oligopsonistic  structure  in  the  fish 
processing  industry  would  further  misallocate 
resources.  Assume  that  effort  had  entered  until 
the  effort  level  which  would  secure  maximum 
physical  sustainable  yield  had  been  exceeded. 
Breaking  up  the  monopsony  and  allowing  the 
higher  social  marginal  valuation  to  be  revealed 
would  further  overcapitalize  the  fishery  and  lead 
to  a  lower  physical  sustainable  yield. 

While  the  argument  has  not  proven  that 
suboptimization  in  either  fisheries  or  fish 
processing  industries  will  lead  to  a  misallocation 
of  resources,  the  possibility  of  such  an  event 
may  lead  to  a  desire  to  gather  more  information 
about  upstream-downstream  linkages  in  these 
industries  before  taking  large  irreversible  policy 
actions.  It  also  may  comfort  those  who  wish  for 
maximum  physical  sustainable  yield  in  the 
fisheries  and  those  who  are  hesitant  about 
breaking  up  what  appears  to  be  an  oligopsonistic 
fish  distribution  chain  in  the  near  future. 

An  extension  of  the  preceding  analysis  may 
lead  one  to  observe  that  imperfectly  competitive 
factor  markets  from  one  side  allows  the  theoreti- 
cal possibility  that  the  unorganized  fishermen 
may  be  able  to  organize  and  bargain  collectively 
for  higher  prices  without  third-party  effects. 
Nevertheless,  the  tenor  of  this  piece  has  sug- 
gested that  third-party  effects  may  possibly  be 
involved  and  that  the  public  interest  may  imply 
that  this  bargaining  should  be  observed  by 
representatives  of  the  third  parties,  such  as 
the  Government. 

Directly  parallel  to  the  problem  of  second 
best  in  the  economy  is  an  ecological  second  best. 
If  two  species  are  competitors  in  the  ocean,  an 
increase  in  the  sustainable  yield  of  one  may 
reduce  the  sustainable  yield  of  the  other.  Like- 
wise, increasing  the  sustainable  yield  of  a 
species  may  increase  the  sustainable  yield  of 
its  predators  and/or  decrease  the  sustainable 
yield  of  its  prey. 

RISK,  UNCERTAINTY,  AND 
INTERTEMPORAL  CHOICE 

Anthony  Scott  (1962)  has  pointed  out  and 
Plourde  (1970)  has  recently  given  a  concise 
proof  that  looking  for  solution  values  on  steady- 
state    curves    (which    these    sustainable    yield 


curves  are)  is  akin  to  ignoring  the  existence  of 
a  positive  rate  of  time  discount.  This  becomes 
obvious  when  one  assumes  an  infinite  rate  of 
time  discount  and  immediately  finds  the  in- 
stantaneous yield  curve  to  be  the  only  relevant 
one  for  consideration. 

When  one  backs  away  from  steady  state 
solutions,  tries  to  pose  the  relevant  horizon 
curve,  and  tries  to  determine  the  role  of  time 
discount,  one  realizes  that  he  is  postulating 
expected  values  of  a  probability  distribution  of 
possible  yield  curves  with  only  a  vague  aware- 
ness of  the  yields  in  short  and  long  run  which 
will  occur.  It  may  be  useful  to  separate  problems 
of  uncertainty  into  two  categories.  On  the  one 
hand,  demand  for  particular  fish  species  is 
uncertain.  It  is  one  thing  to  extrapolate  desires 
for  particular  fish  species  into  the  near  future. 
It  is  quite  something  else  to  fail  to  realize  that 
current  demand  is  dependent  upon  current 
techniques  of  processing  and  marketing  fish. 
The  rapid  rise  of  consumption  of  frozen  fish 
steaks  and  fillets  in  recent  years  is  only  sug- 
gestive of  changes  which  we  can  expect  in 
the  future. 

Major  research  programs,  such  as  those 
supported  by  the  Sea  Grant  college  system,  are 
currently  attempting  to  reduce  uncertainty 
about  fish  supply.  A  number  of  important  areas 
need  to  be  resolved.  To  manage  the  supply  of 
anchovy,  one  needs  to  know  the  biological 
production  function  of  anchovy.  As  already 
suggested,  ecological  parameters  are  needed  to 
manage  both  independent  fisheries  and  biologi- 
cally interdependent  species.  Thus  we  need  to 
understand  the  nature  of  supply  of  all  possible 
species  which  might  occupy  an  ecological  niche. 
It  is  interesting  to  examine  the  controversy 
over  total  yield  of  food  from  the  sea.  As  Chapman 
has  pointed  out  frequently  in  recent  years, 
the  wide  divergence  in  estimates  really  depends 
upon  the  trophic  level  assumed.  Consequently, 
not  only  is  there  uncertainty  about  the  supply 
of  any  particular  species  of  fish,  but  there  is 
uncertainty  about  the  relevant  definition  of 
supply  offish. 

THE  COST  OF  MOVING  TO  A  PARTIAL 
EQUILIBRIUM  POINT 

Equilibrium  points  in  static  analysis  are 
illuminating  for   recommendations   concerning 


25 


direction  of  change.  There  are  several  reasons 
for  realizing,  however,  that  one  may  not  choose 
to  move  to  the  point  of  maximum  rent.  This 
reservation  is  strongest  in  short-run  analysis, 
but  several  parameters  in  a  bioeconomic  model 
can  be  expected  to  shift  in  the  long  run. 

Before  proposing  that  a  fishery  should  be 
managed  at  the  point  of  maximum  net  economic 
yield,  one  must  first  show  that  the  present  value 
of  the  fishery  at  maximum  sustainable  net 
economic  yield  is  greater  than  the  value  of  the 
status  quo  by  more  than  the  transactions  costs 
of  moving  to  the  new  point.  This  was  brought 
out  dramatically  by  Wantrup  (FAO  1962)  and 
also  in  a  comment  by  Crutchfield  to  the  effect 
that  not  reducing  the  existing  level  of  fishing 
effort  could  conceivably  be  a  country's  cheapest 
unemployment  or  welfare  policy  in  cases  where 
the  excess  number  of  fishermen  truly  had  no 
viable  alternative  to  fishing.3 

However,  even  if  the  maximum  sustainable 
net  economic  yield  point  is  greater  in  value  than 
status  quo  by  all  relevant  costs  of  change,  this 
does  not  preclude  the  possibility  that  some 
other  point,  intermediate  to  those  two,  might 
be  more  desirable.  In  summary,  the  proof  of 
superiority  of  a  theoretical  optimum  over  a 
status  quo  position  leads  to  an  argument  for 
direction  of  change  in  effort,  but  does  not  show 
the  magnitude  of  change  until  the  costs  of  such 
a  change  are  themselves  considered. 


PROBLEMS  POSED  BY 
REDISTRIBUTION  OF  INCOME 

Such  a  recommendation  as  moving  to  the 
point  of  maximum  net  economic  yield  is  roughly 
akin  to  the  statement  that  a  readjustment  is 
recommended  whenever  the  dollar  value  to 
potential  gainers  is  greater  than  the  dollar 
value  to  potential  losers.  This  is  the  famous 
Kaldor-Hicks  criterion  for  an  improvement'  in 
social  welfare.  The  criticisms  of  this  criterion 
are  now  well-known  (Rothenberg,  1961)  but  of 
them  all,  the  most  commonly  cited  is  the 
inability  to  judge  among  different  income  dis- 
tributions. To  say  that  one  state  is  better  than 


3  The  comment  was  made  in  August  1969,  during 
discussions  after  a  panel  presentation  given  at  Oregon 
State  University. 


another  state,  when  even  one  individual  is  worse 
off  in  the  former  state,  is  to  make  those  inter- 
personal utility  comparisons  which  the  eco- 
nomics profession  has  largely  disavowed. 

To  confess  that  economists  have  no  straight- 
forward technique  for  judging  among  alternative 
income  distributions  does  not  alter  the  fact  that 
judgments  can  and  will  be  made.  It  does  mean 
that  economists  should  try  to  describe  the 
effects  of  alternative  management  decisions 
upon  the  distribution  of  income.  In  addition  to 
this,  economists  can  realize  that  the  general 
interpretation  of  equity  seems  to  frequently 
preclude  drastic  changes  in  the  distribution  of 
income.  Management  schemes  in  bioeconomic 
models  should  include  systems  which  compen- 
sate losers  whenever  possible.  It  may  be  wise  to 
attach  grandfather  clauses,  unemployment  relief 
funds,  and  the  like  to  licensing  or  other  limited 
entry  proposals.  The  costs  of  preserving  stability 
and  the  existing  distribution  of  income  should 
not  be  overlooked,  creating  a  parallel  to  the 
awesome  headaches  of  our  contemporary  farm 
program. 


AN  OPERATIONAL  PROPOSAL  — 
THE  USE  OF  TARGET  VARIABLES 

Academics  who  would  wish  to  have  a  voice 
in  public  policy  cannot  devote  themselves  to 
being  solely  naysayers.  Decisions  must  and  will 
be  made.  While  there  are  many  reservations 
which  must  be  made  about  bioeconomic  models, 
there  is  also  something  intrinsically  appealing 
about  them.  How  can  one  use  the  information 
concerning  net  economic  yield  and  still  consider 
other  objectives? 

One  possible  technique  is  to  simply  array  a 
group  of  options  for  the  authority  who  has 
received  society's  mandate.  However,  it  is  likely 
that  the  fishery  management  body  itself  will  be 
somewhat  removed  from  direct  interaction  with 
the  society.  Hence,  they  will  probably  need  to 
infer  relative  values  from  another  source.  Our 
experience  with  many  operating  agencies  has 
tended  to  show  that  "the  wheel  that  squeaks 
gets  the  grease."  Thus,  the  use  of  operating 
agency  discretion  may  not  reflect  the  values  of 
the  underlying  society. 

A  second  technique  is  that  proposed  in  most 
bioeconomic    models.    Namely,    net    economic 


26 


yield  would  be  maximized  subject  to  certain 
constraints  on  acceptable  rules  for  redistributing 
income  and  constraints  with  respect  to  minimum 
levels  of  the  assorted  noneconomic  objectives. 
There  is  no  denying  that  this  system  solves  the 
problem  of  weighting  the  various  objectives 
by  simply  avoiding  the  problem.  While  weights 
are  not  explicitly  chosen,  they  are  implicit  in 
the  levels  of  the  noneconomic  variables  selected. 
Consequently,  some  technique  will  have  to  be 
devised  for  continuous  reconsideration  of  non- 
economic  objectives  with  periodic  adjustments 
in  the  level  of  the  constraints  being  made  by 
an  authority  responsible  to  society. 

The  technique  of  constrained  maximization 
will  operate  best  where  clear  threshold  levels 
of  other  variables  can  be  designated.  In  cases  of 
international  fisheries  management,  it  will 
operate  best  when  the  parties  to  the  international 
agreement  can  agree  on  the  objectives  other 
than  net  economic  yield  and  when  relative 
weights  of  more  than  one  species  can  be  specified 
where  more  than  one  commercially  important 
species  is  affected  by  the  management  decision. 
When  this  is  not  true,  a  third  technique  of 
explicitly  agreeing  on  relative  weights  of  several 
objectives  and  maximizing  the  weighted  function 
may  be  superior. 

It  may  well  be  that  developing  a  general 
theory  of  fishery  management  is  to  develop  an 
empty  theory.  Special  consideration  will  be 
needed  for  different  species  of  fish  and  different 
groups  of  nations.  Nonetheless,  it  is  apparent 


that  management  of  ocean  fisheries  is  desired. 
It  is  also  apparent  that  biological  criteria  are 
not  sufficient  to  manage  a  resource  in  a  world 
in  which  there  are  more  goals  than  merely 
consuming  one  particular  species  of  fish.  It  is 
thus  incumbent  upon  us  to  try  to  specify  public 
policy  actions  which  public  authorities  can 
undertake  to  achieve  the  best  possible  mix  of  a 
large  assortment  of  goals. 


LITERATURE  CITED 

BJORK,  GORDON  C,  1969.  Private  Enterprise  and 
Public  Interest,  Prentice-Hall,  Inc.,  p.  65. 

BOULDING,  KENNETH,  1969.  Economics  as  a  Moral 
Science,  American  Economic  Review,  59(  1) :  1-12. 

CRUTCHFIELD,  JAMES,  and  GIULIO  PONTECORVO, 
1969.  The  Pacific  Salmon  Fisheries,  The  Johns  Hopkins 
Press,  p.  35. 

FAO,  1962.  Economic  Effects  of  Fishery  Regulation. 
FAO  Fisheries  Report  No.  5.  Rome. 

PLOURDE,  C.  G.,  1970.  A  Simple  Model  of  Replenishable 
Natural  Resource  Exploitation,  American  Economic 
Review,  60(3):  518-522. 

ROTHENBERG,  JEROME,  1961.  The  Measurement  of 
Social  Welfare,  Prentice-Hall,  Inc.,  pp.  80-103. 

SCOTT,  ANTHONY,  1962.  The  Economics  of  Regulating 
Fisheries,  FAO  Fisheries  Report  No.  5,  Economic- 
Effects  of  Fishery  Regulation,  p.  32. 


21 


Economic,  Political,  and  Social  Barriers  to 
Efficiency  in  Selected  Pacific  Coast  Fisheries 


James  A.  Crutchfield1 
ABSTRACT 

Multidisciplinary  models  are  being  developed  for  the  salmon,  halibut,  king  crab  and 
anchovy  fisheries  as  an  aid  in  fisheries  management.  These  models  will  provide  estimates 
of  economic  rent  in  these  fisheries,  with  an  evaluation  of  alternative  management 
structures  available  to  capture  these  net  benefits.  The  character  of  the  models  for  each 
of  these  differing  fisheries  is  described,  including  reference  to  the  nature  of  the  products, 
markets,  processors,  harvesters,  regulators,  stocks,  and  locations  sectors  of  these  fisheries. 
Introductory  observations  are  made  on  the  future  role  of  multifishery  modeling  studies. 


INTRODUCTION 

In  June  1970  the  University  of  Washington 
and  the  University  of  Rhode  Island  were  funded 
by  the  National  Marine  Fisheries  Service  to  take 
a  first  step  in  identification  and  quantification 
of  the  economic  costs  of  institutional  barriers 
to  the  efficient  use  of  commercially  fished  marine 
stocks.  Anyone  familiar  with  the  American  flag 
fisheries  will  recognize  that  the  time  and  financial 
limitations  of  these  one  or  two  year  studies 
preclude  any  definitive  findings  applicable  on  a 
broad  scale.  Nevertheless,  first  steps  must  some- 
how be  taken,  and  the  two  university  teams, 
together  with  their  NMFS  counterparts,  share 
the  view  that  a  convincing  demonstration  of 
substantial  economic  gains  from  the  elimination 
of  obvious  sources  of  inefficiency  is  one  of  the 
most  important  of  these  steps.  Hopefully,  it  will 
represent  one  phase  of  a  broad-based  attack  on 
the  problems  of  modernizing  the  American 
fisheries  and  rationalizing  the  objectives  and 
techniques  of  management. 

This  paper  presents  a  summary  progress 
report  of  the  Pacific  Coast  studies.  The  project 
has  two  objectives.  In  the  short  run  it  is  intended 
to  provide  reasonable  estimates  of  potential  net 
economic  rent  in  representative  Pacific  Coast 
fisheries,  and  to  explore  the  feasibility  of  alter- 
native management  regimes  to  realize  at  least 
a  portion  of  these  net  benefits.  The  importance 


University  of  Washington. 


of  this  objective  is  underscored  by  increasing 
pressure  for  tangible  evidence  that  the  overall 
activities  of  the  National  Marine  Fisheries 
Service  can  be  translated  into  economic  benefits: 
an  outcome  that  is  anything  but  likely  under 
present  institutional  arrangements  in  the  fish- 
eries. In  the  face  of  increasingly  insistent  de- 
mands on  the  inshore  waters  of  the  United 
States,  and  the  likelihood  of  severe  budget 
stringency  for  an  indefinite  period,  a  convincing 
demonstration  of  the  net  benefits  that  can  be 
generated  by  the  elimination  of  unnecessary 
barriers  to  efficient  harvesting  of  marine  stocks 
may  well  determine  the  future  existence  of  a 
strong  federal  fisheries  function. 

The  longer  term  objective  of  the  study  is  to 
develop  primary  data  and  modeling  capacity  to 
test  fully  alternative  management  and  develop- 
ment regimes.  Previous  studies  of  individual 
segments  of  the  American  fisheries  (Crutchfield 
and  Zellner,  1962;  Crutchfield  and  Pontecorvo, 
1969;  and  Bell  and  Carlson,  1970)  have  been 
concerned  primarily  with  maximum  potential 
net  economic  rent  in  long  run  terms,  with 
varying  assumptions  as  to  acceptance  or  modifi- 
cation of  existing  legal  and  other  constraints. 
It  is  clear,  however,  that  a  full  reevaluation  of 
fishery  management  objectives  requires  a  much 
broader  frame  of  reference  and  a  larger  kit  of 
tools.  Since  it  is  politically  unlikely  that  all 
barriers  to  efficiency  will  be  removed  simul- 
taneously, it  would  be  most  useful  to  develop  a 
modeling  technique  that  would  permit  us  to 
look  at  a  wide  variety  of  measures  or  combina- 
tions of  measures,  at  relatively  low  cost  but  with 


2H 


real  numbers  to  provide  real  estimates  of 
economic  and  biological  effects. 

There  is  equally  urgent  need  for  a  quantified 
model  that  can  be  manipulated  in  terms  of 
multiple  objectives:  economic  efficiency,  income 
distribution,  structural  unemployment,  and  per- 
haps others.  The  modeling  technique  lends  itself 
well  to  assessment  of  a  range  of  management 
measures  that  might  be  undertaken  to  achieve 
multiple  objectives,  or  to  maximize  certain  ele- 
ments subject  to  constrained  values  for  others. 

There  are  both  biological  and  economic  reasons 
for  development  of  a  more  sophisticated  model 
than  the  long  term  equilibrium  constructs  used 
in  earlier  work.  Short  term  adjustments  of  both 
fish  stocks  and  fishermen  to  altered  parameters 
must  be  scrutinized  much  more  carefully.  Simi- 
larly, the  usual  analysis  of  yield  functions,  and 
of  bioeconomic  models  based  on  them,  is  cast  in 
terms  of  a  single  fishery,  while  most  American 
fishing  gear  either  exploits  more  than  one  species 
or  is  capable  of  doing  so.  Even  before  the  eco- 
nomic numbers  to  be  used  in  a  more  complex 
process  model  of  this  sort  can  be  developed,  it 
is  possible  to  derive  a  great  deal  of  knowledge 
of  immediate  benefit  in  assessing  alternative 
management  regimes  by  framing  appropriate 
functional  relations  in  model  form  and  testing 
their  sensitivity  to  various  assumptions  as  to 
quantitative  values. 

In  short,  it  would  be  highly  desirable  to 
develop  a  set  of  models  specific  to  individual 
fisheries  but  geared  to  a  central  common  frame- 
work that  would  permit  comparison  among 
fisheries.  Obviously,  this  will  not  be  done  in  a 
day  or  a  year ;  but  if  a  good  start  can  be  made  in 
isolating  the  functions  that  must  be  quantified 
and  delineating  data  requirements,  the  ultimate 
payoff  in  terms  of  flexibility  and  low  operating 
cost  will  make  possible  a  dynamic  concept  of 
fisheries  management  that  can  really  utilize 
increases  in  scientific  knowledge,  improved 
technology,  and  more  flexible  administrative 
arrangements. 

THE  FISHERIES 


The  Pacific  halibut  operation  is  a  mature  fishery, 
relatively  simple  in  economic  structure,  and 
employing  only  a  single  type  of  gear.  It  has 
been  under  a  carefully  conceived  regulatory 
program  for  a  sufficiently  long  period  to  gen- 
erate excellent  data  on  both  biological  and 
economic  variables. 

The  Pacific  salmon  fishery  stands  at  almost 
the  opposite  extreme.  It  is  complex  in  every 
sense  —  biological  and  economic  —  that  can 
be  imagined.  It  is  subject  to  inherent  data 
limitations  since  it  is  based  on  populations  that 
are  in  constant  short  run  disequilibrium,  and 
it  is  now  regulated  on  such  an  irrational  basis 
that  great  improvement  is  possible  with  rela- 
tively simple  alterations  in  management  tech- 
niques. 

The  California  anchovy  fishery,  barely  ex- 
ploited at  the  present  time,  represents  one  of 
the  largest  single  latent  resources  available  to 
American  flag  fishermen.  On  the  assumption 
that  present  legal  limitations  on  commercial 
exploitation  are  removed,  the  potential  physical 
yield  from  the  fishery  is  almost  as  great  as  the 
total  United  States  landed  catch.  The  possibility 
of  creating  a  new  and  highly  attractive  industry 
under  controlled  entry  conditions  is  intriguing, 
to  say  the  least.  Data  on  the  California  anchovy 
are  still  rather  limited,  but  the  basic  stock 
information  is  being  developed  rapidly,  and 
both  the  biological  and  economic  analysis  can 
borrow  extensively  from  the  broad  experience 
of  the  Peruvian  anchoveta  fishery. 

The  king  crab  fishery  of  the  North  Pacific 
presents  a  classic  example  of  the  speed  with 
which  modern  technology,  under  conditions  of 
open  entry,  can  lead  to  overinvestment,  over- 
fishing, and  potential  economic  disaster.  In 
addition,  the  hastily  conceived  regulations  now 
in  effect  present  some  of  the  worst  examples  of 
efficiency-reducing  techniques,  coupled  with 
obvious  efforts  to  redistribute  income  from  one 
set  of  fishermen  to  another.  Data  are  woefully 
inadequate  in  this  fishery,  but  its  economic 
value  and  potential  make  it  an  excellent 
case  study. 


The  four  fisheries  chosen  for  analysis  were 
selected  for  characteristics  which  make  them 
broadly  representative  of  the  kinds  of  problems 
to  be  faced  in  future  fishery  management  pro- 
grams geared  more  closely  to  economic  objectives. 


THE  MODELING  FRAMEWORK 

The  simulation  approach  which  serves  as  the 
basis  for  the  longer  run  aspects  of  this  project 
is  hinged  on  a  general  model  which  is  adaptable 


2!) 


to  each  of  the  specific  fisheries  to  be  studied. 
It  involves  a  large  computer  program  which 
provides  a  framework  for  studying  short  term 
and  long  term  effects  of  alternative  regulatory 
policies  on  economic  and  biological  performance 
of  various  sectors  of  a  fishery.  The  basic  program 
is  written  in  a  version  of  FORTRAN  IV.  A 
schematic  of  the  basic  model  is  shown  in  Figure 
1.  The  sectors  simulated  by  the  program  are: 
products,  markets,  processors,  harvesters,  regu- 
lators, stocks,  and  locations.  General  operation 
of  the  model  is  as  follows.  The  stock  sector 
"grows"  the  resources  and  determines  the 
amount  of  each  stock  which  is  available  for 
harvest  in  each  location.  Harvesters  operate  in 
locations  of  their  choosing,  catch  a  portion  of 


the  available  stock,  and  sell  it  to  processors. 
Processors  convert  their  purchases  into  finished 
goods  and  offer  them  for  sale  in  the  markets 
which  are  available.  Demand  (and  the  marketing 
activity  of  the  processors)  determine  sales  by 
each  processor  of  each  product  in  each  market. 
The  regulators  are  free  to  impose  restrictions 
of  various  types  on  the  activities  of  both 
processors  and  harvesters.  In  operation  the 
program  compiles  statistics  on  the  operation  of 
the  system  and  prints  out  monthly  and  annual 
summaries  of  these  statistics.  The  detail  of  the 
printout  is  optional. 

Each  harvesting  group  operates  as  a  semi- 
independent  unit,  constrained  only  by  links  to 
a  location  and  one  set  of  processors.  At  the  start 


Product 
Sector 


Consumer 
Market  Sector 


Processor 
Sector 


Product 
1 


Harvester 
Sector 


Location 
Sector 


Stock 
Sector 


Regulator 
Sector 


Harvester 

1 


Stock 
1 


Regulator 
1 


Market 


Harvester 
2 


Harvester 


Stock 
2 


Regulator 
2 


Regulator 


Figure  1.  —  Structure  of  fishery  simulator. 

30 


of  each  month  a  group  of  fishing  units  moves 
from  its  initial  location  to  the  harvesting  location, 
operates  there  for  a  specified  number  of  days, 
and  sells  its  catch  to  the  processor  which  will 
produce  the  maximum  profit  to  the  group.  If 
the  processor  cannot  absorb  the  group's  total 
supply,  the  group  sells  what  it  can  and  moves 
to  the  next  most  profitable  processor  and  so  on. 
A  group  may  only  sell  to  the  set  of  processors 
linked  to  the  harvester  that  owns  the  group. 

Any  catch  unsold  at  the  end  of  the  month  is 
recorded  and  discarded  since  it  has  no  economic 
value.  The  harvesting  time  for  each  group  is 
limited  not  only  by  harvester  and  regulator 
decisions,  but  by  the  group's  harvesting  capacity. 
Primary  operating  costs  for  the  harvester  built 
into  the  model  are  distance  costs,  time  costs, 
harvest-proportional  costs,  and  license  fees. 

A  processor  is  a  managerial  entity  that 
operates  in  one  physical  location,  buying  stocks 
from  harvesters,  transforming  them  into  finished 
products,  and  selling  them  in  markets.  As  the 
program  is  now  set  up,  each  processor's  share 
of  the  market  can  be  made  to  depend  on  his 
previous  market  share  and  on  marketing  ex- 
penditures and  product  price  relative  to  those 
of  other  processors  linked  to  the  same  market. 
The  cost  structure  of  the  processors  is  in  standard 
accounting  terms. 

A  regulator  is  an  agency  that  imposes  restric- 
tions on  the  activities  of  harvesters  or  processors 
in  any  of  the  variety  of  ways  now  employed  or 
discussed  in  the  literature.  These  include: 
(1)  license  fees;  (2)  size  limits;  (3)  gear  efficiency 
limits;  (4)  effort  limits;  (5)  operating  limits  for 
processors;  (6)  seasonal  closures;  (7)  monthly 
quotas;  (8)  annual  quotas. 

The  model  can  be  run  in  either  of  two  basic 
operational  modes:  as  a  conventional  computer 
simulation  model,  with  built-in  decision-making 
algorithms  specifying  the  behavior  of  processors, 
harvesters,  and  regulators;  or  with  human  inter- 
vention at  intervals  to  allow  for  intuitive  and 
heuristic  decisionmaking. 

A  stock  is  any  type  of  renewable  marine 
resource.  It  is  treated  in  the  model  as  linked  to 
a  given  location,  and  the  quantities  available 
to  harvesters  at  any  given  time  are  computed 
continuously. 

It  should  be  stressed  that  each  entity  is  a 
subroutine,  and  can  be  designed  to  any  degree  of 
complexity  warranted  by   the   purpose  of  the 


routine  and  the  adequacy  of  the  data  base. 
Similarly,  the  degree  of  detail  for  a  readout  on 
monthly  or  annual  bases  can  be  predetermined. 
The  model  can  be  programmed  not  only  to 
maximize  specific  objective  functions,  but  can 
accommodate  dynamic  feedback  factors  in  assess- 
ing different  kinds  of  management  alternatives. 
It  can  also  handle  a  wide  range  of  spatial 
distributions  of  stocks  and  harvesters  without 
difficulty. 

THE  ANCHOVY  FISHERY 

Figure  2  shows,  in  schematic  form,  a  pre- 
liminary version  of  the  model  of  the  California 
anchovy  fishery.  This  model  reflects  the  activities 
of  the  types  of  vessels  presently  exploiting  the 
fishery,  and  therefore  attempts  to  deal  with  the 
complications  imposed  by  their  harvesting  of 
bluefin  tuna  and  mackerel  as  well  as  anchovy. 
It  is  also  complicated  by  the  interaction  between 
the  markets  for  sport  fishing  bait  and  for  meal 
and  oil,  both  of  which  now  absorb  considerable 
quantities  of  anchovy.  Sufficient  data  are  avail- 
able to  permit  some  preliminary  conclusions  as 
to  the  economic  return  from  this  limited  fishery, 
which  is  now  prosecuted  at  a  level  so  low  that 
the  more  fundamental  problems  in  the  stock 
sector  are  not  really  involved.  These  preliminary 
findings  suggest,  as  one  might  suspect,  that  the 
return  to  vessels  fishing  for  anchovy  on  a  full- 
time  basis  during  a  nine  months  open  season 
would  be  substantially  more  attractive  than  the 
returns  from  mixed  operations. 

Accordingly,  the  model  which  will  be  used  to 
test  regulatory  alternatives  will  probably  be 
based  on  the  assumption  that  a  specialized 
fleet  of  vessels  optimized  for  the  anchovy  fishery 
will  develop  once  catch  quotas  are  established 
at  levels  sufficiently  high  to  induce  long  term 
investment  in  fishing  and  processing  equipment 
by  major  firms  in  the  meal  industry.  Preliminary 
work  in  a  dissertation  by  Dr.  Dennis  Paulaha 
(1970)  provides  excellent  data  on  the  type  of 
vessel  and  gear  best  adapted  to  the  fishery. 
Work  on  the  complex  stock  model  is  reasonably 
well  advanced,  and  it  is  expected  that  a  fairly 
sophisticated  and  realistic  approximation  to  the 
behavior  of  the  anchovy  stocks  under  various 
rates  of  exploitation  can  be  developed. 

The  economics  of  the  anchovy  operation  are 
relatively  simple  to  simulate,  since  total  produc- 


31 


Meal 


Canned 
JM 


Live 
Bait 


Monterey 
Reducers 


Los  Angeles 
Canners  5. 
Reducers 


Calif. 
Fish  & 
Game 


Monterey 
Fleet 


Wetfish 
Fleet 


Northern 
Anchovy 


Jack 

Mackerel 


Calif. 

current 

System 


Bait 
Processors 


*-->--> - 


Special 

Anchovy 

Fleet 


Rait 
Fleet 


Figure    2.    -  -    Graphic    representation    of   the    logical    re- 
lationships  between   sectors   of  the  Simplified  Northern 
Anchovy  Fishery  System.  (Area  quotas  of  total  system 
are    considered    to    be    levied    as    production    quotas    in 
the  simplified  system.) 


tion  from  the  fishery,  even  at  a  catch  level  of 
one  million  tons,  would  still  produce  only  a 
small  fraction  of  total  American  fish  meal 
consumption.  Market  price  can  thus  be  taken 
as  given  to  the  California  meal  producer,  and 
the  estimated  net  economic  rent  available  under 
various  assumptions  as  to  management  regime 
can  then  be  calculated  on  the  basis  of  alternative 
forecasts  of  the  time-path  offish  meal  prices. 

THE  PACIFIC  SALMON  FISHERY 

In  Appendix  I  the  general  format  of  a  pre- 
liminary program  for  modeling  the  Pacific 
salmon  fishery  is  presented.  The  objectives  in 
modeling  this  extraordinarily  complex  opera- 
tion are  partly  methodological  and  partly  aimed 
at  answering  specific  management  problems  of 
real    significance.    The    complications    are    ap- 


parent. Five  separate  species  of  salmon  are 
involved,  and  since  each  river  usually  contains 
more  than  one  species  (and  separate  races  of 
the  same  species),  the  number  of  "management 
units"  which  should,  in  theory,  receive  separate 
treatment  in  modeling  the  stock  sector  is 
probably  from  eight  to  ten  thousand] 

"The  salmon  fishery"  is  actually  a  large  num- 
ber of  geographically  separate  operations,  linked 
in  varying  degrees  by  the  mobility  of  the  gear 
involved.  Several  types  of  gear  are  used,  and  the 
relative  importance  of  each  type  varies  from 
area  to  area.  Finally,  salmon  deteriorate  very 
rapidly  unless  processed  soon  after  being  cap- 
tured, which  creates  a  large  number  of  primary 
markets  in  which  processors  generate  several 
different  end  products  from  each  of  the  types 
of  salmon  purchased. 

Even    with    the   prodigious   capacity   of  the 


32 


modern  computer,  overall  modeling  of  a  fishery 
this  complex  is  obviously  severely  limited  by 
available  data,  and  the  marginal  cost  of  generat- 
ing the  necessary  data  is  very  high.  In  one 
sense,  then,  the  broader  modeling  exercise  is 
intended  to  provide  some  guidelines  to  the 
limitations  on  the  technique  in  dealing  with 
highly  complex  fisheries. 

On  the  other  hand,  the  program  is  flexible 
enough  to  permit  specific  consideration  of  im- 
portant policy  questions  in  separable  segments 
of  the  salmon  fishery.  For  example,  the  Columbia 
River,  Puget  Sound,  and  British  Columbia 
fisheries  are  plagued  by  serious  problems  arising 
from  the  spectacular  growth  of  the  ocean  troll 
fishery.  Since  the  trollers  take  large  numbers 
of  immature  fish,  they  do  a  considerable,  though 
unknown,  amount  of  damage  in  returning  under- 
sized fish  to  the  water.  The  troll  fishery  is 
inherently  highly  inefficient  from  both  biological 
and  economic  points  of  view.  A  substantial 
part  of  its  catch  is  made  up  of  two  and  three 
year  old  chinooks  and  two  year  old  cohoes 
which  would  almost  certainly  gain  substantially 
more  in  body  weight  than  would  be  lost  to 
natural  mortality  if  allowed  to  mature  another 
year.  It  is  possible,  with  a  restricted  model,  to 
test  the  biological  and  economic  impact  of  the 
elimination  or  limitation  of  the  troll  fishery  in 
specified  areas  under  varying  assumptions  as 
to  the  resulting  net  increment  in  weight  and  the 
distribution  of  the  troll  catch  among  other 
types  of  gear.  This  analysis  is,  incidentally, 
crucial  to  another  public  policy  issue  of  major 
proportions  —  the  allocation  of  chinook  and 
coho  salmon  among  commercial  and  recreational 
users. 

Earlier  work  by  a  University  of  Washington 
team  on  the  Puget  Sound  salmon  fishery  (Royce, 
et  al.,  1963)  indicates  that  modeling  permits 
surprisingly  accurate  prediction  of  the  net 
economic  benefits  and  catch  distribution  effects 
by  area  of  different  techniques  for  reduction  of 
gear  and  expansion  of  intraseasonal  fishing 
time.  The  earlier  study  was,  for  strategic  rea- 
sons, constrained  by  the  assumption  that  any 
reduction  of  gear  must  be  proportional  for  each 
type  of  gear.  It  is  obviously  desirable  to  develop 
the  capability  to  test  quickly  and  inexpensively 
the  effects  of  altering  gear  mix  by  area  and  by 
time  period.  Since  any  gear  reduction  program 
in  the  salmon  fishery  will   inevitably  involve 


intensely  partisan  political  negotiations,  a 
display  of  the  impact  of  a  wide  range  of  alterna- 
tives is  essential  if  any  progress  is  to  be  made. 
Finally,  the  model  can  be  used  to  predict  the 
impact  of  recent  court  decisions  requiring  that 
Indian  fishermen  must  be  granted  a  "prior 
claim"  on  any  total  catch  permitted  under 
regulation. 

THE  PACIFIC  HALIBUT  FISHERY 

This  fishery  presents  a  far  simpler  set  of 
modeling  problems.  The  stocks  have  been  under 
intensive  study  for  more  than  40  years,  and  a 
wealth  of  reliable  statistical  information  is 
available  on  both  the  stock  and  harvesting 
sectors.  In  addition,  the  use  of  standard  gear 
(whatever  its  economic  validity)  makes  analysis 
much  simpler,  as  does  the  widespread  use  of 
a  standard  accounting  system  for  halibut  vessels 
devised  by  the  Fishing  Vessel  Owners  Associa- 
tion. The  fact  that  halibut  is  marketed  almost 
entirely  in  fresh  and  frozen  form  further 
simplifies  the  analysis.  The  principal  gain  in 
the  modeling  exercise  for  this  fishery  will  be 
the  ability  to  incorporate  badly  needed  studies 
of  the  effects  of  introducing  different  types  of 
gear,  potentially  much  more  efficient,  if  and 
when  limitation  of  the  number  of  operating 
units  becomes  possible.  It  will  also  be  possible 
to  introduce  into  the  analysis  the  effect  of 
shifting  many  of  the  halibut  vessels  from  their 
present  multipurpose  form  into  larger,  specializ- 
ed units  —  a  process  which  would  almost 
certainly  follow  any  effective  gear  limitation 
program. 

THE  KING  CRAB  FISHERY 

It  is  doubtful  that  the  king  crab  fishery  will 
be  amenable  to  very  effective  empirical  work 
in  the  near  future.  Not  only  are  data  extremely 
limited,  but  the  fishery  is  based  on  a  relatively 
long-lived,  slow-growing  animal,  and  it  is 
currently  in  a  state  of  disequilibrium.  Con- 
sequently, the  fragmentary  statistical  informa- 
tion on  catch,  effort,  and  economic  returns 
from  the  fishery  cannot  be  considered  representa- 
tive of  long  term  equilibrium  values.  Neverthe- 
less, the  situation  in  the  king  crab  fishery  with 
respect  to  stock  depletion  is  so  serious,  and  the 


33 


regulatory  methods  already  adopted  so  question- 
able, that  some  analysis  of  this  fishery,  even 
with  very  limited  data,  is  clearly  necessary  if 
we  are  to  avoid  serious  and  perhaps  irreparable 
mistakes. 

CONCLUSIONS 

If  the  approach  embodied  in  this  study  proves 
to  be  as  useful  as  expected,  it  is  considered 
possible  that  the  techniques  could  be  extended  to 
provide  a  broader  approach  to  multifishery 
cases.  The  seasonal  nature  of  the  availability  of 
fish  and  of  weather  conditions  on  the  Pacific 
Coast  suggests  that  an  optimal  harvesting 
technique  for  virtually  all  species  (with  the 
possible  exception  of  halibut  and  some  other 
bottomfish)  will  involve  multipurpose  gear 
exploiting  multiple  species  in  different  geo- 
graphic locations.  For  example,  salmon,  crab, 
and  albacore  fishing  by  combination  units  may 
be  significantly  more  attractive  economically 
(always  assuming  some  control  over  entry) 
than  the  present  hodgepodge  of  vessels  involved 
in  each.  We  do  have  combination  vessels  at 
present,  of  course,  but  they  are  not  designed  to 
any  set  of  specifications  that  present  data  would 
make  available  if  an  integrated  view  of  the 
fisheries  available  to  each  type  of  gear  were 
taken  as  the  frame  of  reference. 

The  discussion  above  suggests  the  nature  of 
the  outputs  to  be  expected  from  these  models 
in  the  short  run.  We  are  still  limited  to  syn- 
thetic numbers  in  many  of  the  sectors  at  present, 
but  these  are  being  systematically  whittled 
down.  It  cannot  be  stressed  too  strongly  that 
the  whittling  down  process  can  be  done  far 
more  economically  and  effectively  once  the 
sensitivity  of  the  desired  outputs  to  the  various 
parameters  involved  has  been  established  by 
the  model.  Moreover,  some  of  the  fisheries  and 
some  elements  of  the  model  have  now  reached 
a  point  where  reasonably  hard  data  are  avail- 
able which  can  be  manipulated  to  provide  at 
least  rank-ordering  of  a  number  of  management 


options.  While  the  overall  program  is  clearly 
geared  to  longer  range  objectives,  short  run 
outputs  of  real  usefulness  in  management 
planning  can  be  expected,  and  will  increase  in 
number  and  predictive  value  as  the  work 
progresses. 

Members  of  this  workshop  and  other  interested 
scientists  and  economists  are  urged  to  com- 
municate to  us  the  nature  of  their  interest  in 
the  problems  addressed  by  the  University  of 
Washington  team.  In  addition,  it  might  be 
mutually  advantageous  if  visits  to  the  University 
of  Washington  could  be  arranged  to  permit 
actual  operating  experience  with  these  models. 


LITERATURE  CITED 

BELL,  FREDERICK  W.,  1970.  Estimation  of  the 
Economic  Benefits  to  Fishermen,  Vessels  and  Society 
from  Limited  Entry  to  the  Inshore  U.S.  Northern 
Lobster  Fishery,  Marine  Technology  1970  Preprints  — 
Vol.  1.  Marine  Technology  Society,  6th  Annual  Con- 
ference and  Exposition  June  29-July  1,  1970,  Wash- 
ington, D.C. 

CRUTCHFIELD,  JAMES  and  GIULIO  PONTECORVO, 
1969.  The  Pacific  Salmon  Fisheries.  The  Johns  Hopkins 
Press,  Baltimore,  Maryland. 

CRUTCHFIELD,  J.A.  and  A.  ZELLNER,  1962.  Economic 
Aspects  of  the  Pacific  Halibut  Industry,  Fishery  In- 
dustrial Research.  United  States  Department  of  Interior, 
Fish  and  Wildlife  Service,  Bureau  of  Commercial 
Fisheries,  U.S.  Government  Printing  Office,  Washington, 
D.C. 

PAULAHA,  DENNIS  E.,  1970.  A  General  Economic 
Model  for  Commercial  Fisheries  and  its  Application  to 
the  California  Anchovy  Fishery.  Unpublished  Ph.  D. 
Dissertation.  University  of  Washington. 

RICKER,  W.E..  1958.  Handbook  of  Computations  for 
Biological  Statistics  of  Fish  Populations.  Bulletin  of 
the  Fisheries  Research  Board  of  Canada,  No.  119, 
pp.  1-300. 

ROYCE,  W.,  D.  BE  VAN,  J.  CRUTCHFIELD,  G.  PAULIK, 
R.  FLETCHER,  1963.  Salmon  Gear  Limitation  in 
Northern  Washington  Waters.  University  of  Washington 
Publications  in  Fisheries,  New  Series,  Vol.  II,  No.  1. 


34 


APPENDIX  I:  PACIFIC  SALMON  SIMULATION  MODEL  COMPONENT 

The  proposed  simulation  model  will  treat  the  five  species  of  Pacific  salmon  in  the  North  American 
fisheries  as  five  separate  stocks: 


s, 

Chinook 

s, 

Chum 

Sa 

Pink 

s, 

Sockeye 

s. 

Coho 

The  location  sector  in  the  model  will  be  based  on  the  areas  for  which  statistical  information  is 
available  in  published  data  sources: 


L, 
L2 
L3 
U 
I,, 
L6 
U 
L8 
L9 
Lio 


Western  Alaska 

Central  Alaska 

Southeastern  Alaska 

Northern  B.C. 

Southern  B.C./Fraser  River 

Puget  Sound 

Washington  Coast 

Columbia  River 

Oregon  Coast 

California 


Stock/location  interaction  will  be  as  follows: 


Chinook      Chum      Pink      Sockeye      Coho 
S\  02  03  04  05 


Li  Western  Alaska 

L2  Central  Alaska 

L3  Southeastern  Alaska 

L4  Northern  B.C. 

L5  Southern  B.C./Fraser  River 

L6  Puget  Sound 

L7  Washington  Coast 

L8  Columbia  River 

L9  Oregon  Coast 

Lio  California 


X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

— 

X 

— 

X 

X 

— 

— 

— 

X 

and  these  interactions  reflect  the  distribution  of  species  in  the  actual  fishery.  Data  will  be  collected 
to  permit  segregation  of  stocks  into  age  groups,  with  age  specific  weights  and  spawner-return  curves 
(Ricker,  1958)  developed  for  each  stock/location.  In  effect,  this  scheme  results  in  a  total  of  45  separate 
stocks,  since  each  species  in  each  location  will  be  treated  separately  with  respect  to  spawner-return 
characteristics. 

Regulators  will  be  based  on  locations,  with  one  regulator  in  each  location. 

The  principal  types  of  fishing  gear  in  the  Pacific  salmon  fisheries  are  as  follows: 

Gill  nets,  drift 
Gill  nets,  anchor 
Seines 


35 


Troll  lines 

Reef  and  Pound  nets 

In  order  that  each  harvester  be  able  to  fish  for  each  species  with  each  type  of  gear,  it  is  necessary 
that  the  harvesters  be  denned  as  follows: 


Hi  Seine,  Western  Alaska  (Li) 

Ho  Seine,  Central  Alaska  (L2) 

H3  Seine,  Southeastern  Alaska  (L3) 

H4  Seine,  Northern  B.C.  (L4) 

H5  Seine,  Southern  B.C./Fraser  (L5) 

H6  Seine,  Puget  Sound  (L6) 

H7  Anchor  Gill  Net,  Western  Alaska  (Li) 

H8  Anchor  Gill  Net,  Central  Alaska  (L2) 

H9  Anchor  Gill  Net,  Southeastern  Alaska  (L3) 

H10  Anchor  Gill  Net,  Puget  Sound  (L6) 

Hn  Anchor  Gill  Net,  Wash.  Coast  (L7) 

H12  Anchor  Gill  Net,  Columbia  River  (L8) 

H13  Drift  Gill  Net,  Western  Alaska  (Li) 

Hi4  Drift  Gill  Net,  Central  Alaska  (L2) 

H15  Drift  Gill  Net,  Southeastern  Alaska  (L3) 

H16  Drift  Gill  Net,  Puget  Sound  (L6) 

Hi7  Drift  Gill  Net,  Wash.  Coast  (L7) 

His  Drift  Gill  Net,  Columbia  River  (L8) 

Hi9  Gill  Net,  Northern  B.C.  (L4) 

H20  Gill  Net,  Southern  B.C./Fraser  River  (L5) 

H2i  Troll,  Central  Alaska  (L2) 

H22  Troll,  Southeastern  Alaska  (L3) 

H23  Troll,  Northern  B.C.  (L4) 

H24  Troll,  Southern  B.C./Fraser  (L5) 

H25  Troll,  Puget  Sound  (L6) 

H26  Troll,  Wash.  Coast  (L7) 

H27  Troll,  Columbia  River  (L8) 

H28  Troll,  Oregon  Coast  (L9) 

H2g  Troll,  California  (Li0) 

H30  Reef  &  Pound  Nets,  Puget  Sound  (L6) 

The  above  table  indicates  that  there  is  no  fleet  for  a  particular  species  in  those  cases  where  the 
annual  catch  for  that  species  in  that  location  by  that  gear  type  is  less  than  one  percent  (1%)  of  the 
total  annual  catch  of  that  species  in  that  location  by  all  gear  types. 

There  will  be  one  processer  in  each  location,  with  processer  locations  defined  for  distance 
computation  purposes  as  follows: 

Pi  Bristol  Bay  (Western  Alaska) 

P2  Cook  Inlet  (Central  Alaska) 

P3  Yakutat  (Southeastern  Alaska) 

P4  Prince  Rupert  (Northern  B.C.) 

36 


Chinook 

Chum 

Pink 

Sockeye 

Co  ho 

Si 

s2 

s3 

s4 

S5 

X 

X 

X 



X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

— 

— 

— 

X 

X 

X 

X 

X 

— 

X 

X 

X 

— 

X 

X 

X 

— 

— 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

— 

— 

X 

X 

X 

— 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

— 

— 

— 

X 

X 

— 

— 

— 

X 

X 

— 

X 

X 

X 

X 

— 

X 

X 

X 

X 

— 

X 

— 

X 

X 

— 

X 

— 

X 

X 

— 

X 

— 

X 

X 

— 

X 

— 

X 

X 

— 

— 

— 

X 

X 

— 

X 

X 

X 

P5  Vancouver  (Southern  B.C./Fraser  River) 

P6  Seattle  (Puget  Sound) 

P7  Westport  (Washington  Coast) 

P8  Astoria  (Columbia  River) 

P9  Newport  (Oregon  Coast) 

Pio  San  Francisco  (California) 

Markets  will  be  synonymous  with  products,  with  demand  relationships  developed  for  each 
product  as  follows: 


I), 

Fresh/frozen 

I)., 

Salted  or  pickled 

D3 

Mild  cured 

I), 

Smoked  or  kippered 

D5 

Canned 

D6 

Roe  (cured) 

Products  will  be  produced  by  processers  as  follows: 


Chinook      Chum      Pink      Sockeye      Coho 
Si  S2  S3  04  05 


D2 
D3 
D5 
D6 


Pi        Bristol  Bay 

Di   Fresh /frozen 
Salted  or  pickled 
Mild  cured 
Canned 
Roe  (cured) 
P2        Cook  Inlet 
D, 
D2 
D5 
D6 
P3       Yakutat 
Di 
D3 
1), 
D6 
P4        Prince  Rupert 

P5        Vancouver 


X 

X 

X 

— 

X 

X 
X 
X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

— 

— 

— 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

— 

— 

— 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

P6        Seattle  (data  available  for  Washington  as  a  whole) 

P7        Westport 

D3 
D4 

I), 
D6 

P8        Astoria  (data  available  for  Oregon  as  a  whole) 

37 


X 

X 

X 

X 

X 

X 
X 
X 

— 

— 

— 

X 

X 

X 

X 

X 

X 

X 

X 

X 

X 

Chinook      Chum  Pink  Sockeye  Coho 

Si  S'2  S3  S4  S5 

P9       Newport 

D4                                                                                            X  —  —  —  — 

D5                                                                                            X  X  X  X  X 

P10      San  Francisco 

D,                                                                                            X  —  —  —  X 

D3                                                                                            X  —  —  —  X 

D4                                                                                            X  —  —  —  X 


38 


PRODUCTION  FUNCTIONS  AND  BIOECONOMIC 
MODELS:  RESEARCH  IMPLICATIONS 


Against  the  broad  background  of  these  four 
introductory  papers  we  can  proceed  to  some  of 
the  more  specific  research  which  will  constitute 
the  principal  inputs  into  the  broader  manage- 
ment process.  The  first  of  these  papers  relates 
the  results  of  an  extensive  effort  by  Carlson  to 
specify  production  functions  for  the  North 
Atlantic  groundfish  and  tropical  tuna  fisheries. 
In  each  case  the  research  is  designed  to  identify 
the  most  significant  determinants  of  vessel 
productivity,  with  some  of  the  investigation 
devoted  to  the  question  of  a  proper  measure 
of  productivity. 

Using  existing  data  series  on  the  area  and 
time  patterns  of  fishing  activity,  landings 
statistics  on  species,  quantity  and  value,  and 
other  sources  of  data  on  vessel  characteristics, 
specific  effort  combinations  are  related  to  produc- 
tivity. The  "best"  measure  of  productivity  was 
found  to  be  value  in  groundfish  and  a  weighted 
combination  of  species  landed  in  tuna. 

This  research  output  has  many  possible  uses, 
among  these  being  the  suggestion  of  the 
optimum  input  package  to  maximize  output 
and  the  development  of  a  fishing  power  index 
which  could  be  used  to  measure  effort,  a  critical 
input  into  those  types  of  management  plans  that 
require  the  administrator  to  develop  seasonal 
or  sharing  arrangements  based  on  the  fishing 
capabilities  of  the  fleet.  This  is  the  case  for  the 
Inter-American  Tropical  Tuna  Commission. 
Here  a  technique  of  measuring  fishing  power 
has  evolved  which  is  somewhat  different  from 
the  Carlson  approach.  Future  investigations 
will  determine  the  advisability  of  each  approach. 
Indeed,  if  differences  and  difficulties  cannot  be 
resolved,  this  may  have  some  effect  on  the  choice 
between  management  plans  which  require  this 
type  of  calculation  and  types  which  do  not. 

The  paper  by  Segura  relates  part  of  his  broad 
investigation  into  the  world  supply  and  demand 
for  fish  meal.  His  efforts  for  this  paper  have 
concentrated  on  a  measure  of  fishing  power  in 
the  Peruvian  anchoveta  fleet  for  the  purpose  of 
determining  the  optimum  harvest  level.  His 
focus  is  upon  the  role  of  technological  change 


as  this  relates  to  time  series  calculations  of 
effort  indices. 

In  his  paper,  Segura  points  out  the  differing 
results  which  will  be  forthcoming  if  you  use 
the  most  recent  years'  measure  of  yield-effort, 
the  index  of  vessel  productivity,  to  calculate 
changing  pressures  on  the  resource,  the  response 
of  the  resource  to  that  pressure,  and  use  these 
relationships  to  determine  an  optimum  catch 
quota  for  the  coming  year.  He  compares  these 
results  to  calculations  now  used  where  these 
interrelationships  are  all  derived  based  upon 
some  earlier  base  year.  The  results  are  sub- 
stantially different,  resulting  in  a  suggested 
catch  of  16.2  million  ton  trips  derived  via  the 
existing  method. 

The  work  done  by  Segura  relates  closely  to 
that  of  Carlson  in  that  a  method  of  cross- 
sectional  analysis  of  recent  years'  data  is  being 
developed  which  obviates  the  need  to  use 
standard  vessels  from  some  base  period,  supple- 
mented by  ad  hoc  measures  of  technological 
change.  These  considerations  are  in  addition  to 
the  question  of  diminishing  returns  as  intro- 
duced by  the  Carlson-Waugh-Bell  function. 

The  research  reported  by  Rich  is  an  extension 
of  a  generalized  model  to  be  applied  to  the 
Pacific  halibut  fishery.  The  purpose  is  to  evaluate 
possible  losses  which  may  have  resulted  in  the 
fishery  from  the  use  of  MSY  as  a  regulation  goal. 

Consistent  with  the  Carlson-Waugh-Bell 
exposition,  the  function  developed  incorporates 
short  run  diminishing  returns.  When  combined 
with  a  fish  growth  function  it  is  possible  to 
measure  the  long  run  externalities  associated 
with  this  alternative  specification  of  the  yield- 
effort  function. 

This  approach  is  the  antithesis  of  that  sug- 
gested by  Pontecorvo  in  that  it  is  explicitly 
structured  upon  the  classic  assumptions  of  full 
employment  and  complete  labor  mobility,  both 
in  the  short  run  and  the  long  run.  Political 
and  social  questions  are  definitely  excluded 
and  would  have  to  be  appended  on  an  ad  hoc 
basis  to  determine  if  there  was  any  cause  for 
modifying  the  constrained   results.   The  work 


39 


done  by  Rich  would  serve  as  but  one  component 
in  the  simulator  described  by  Crutchfield, 
albeit  possibly  the  dominant  component. 

Bell,  Carlson,  and  Waugh  focus  on  the  issue 
of  diminishing  returns  in  fisheries,  relaxing  a 
strong  assumption  of  fixed  proportionality 
utilized  by  most  writers  in  the  existing  literature. 

The  motivation  for  this  exercise  is  the  ap- 
preciation that  we  are  rapidly  approaching  total 
utilization  of  the  world's  fish  resources.  As  this 
point  is  approached,  demand  pressures  and 
considerations  of  maximum  efficiency  dictate 
the  need  to  make  maximum  use  of  these 
resources  consistent  with  any  overriding  con- 
servation objectives.  The  work  done  by  these 
authors,  though  preliminary,  suggests  that 
some  degree  of  diminishing  returns  can  be 
identified  for  the  fisheries  studied:  Chesapeake 
Bay  menhaden,  Atlantic  and  Gulf  blue  crab, 
Atlantic  longline  tuna,  Bering  Sea  king  crab, 
and  Cape  Flattery  sablefish. 

As  with  the  other  five  papers  in  this  section, 
this  paper  modifies  existing  biological  functions. 
The  modified  logistic  introduced  here  is  the 
author's  candidate  for  a  "better"  function, 
based  primarily  on  the  inclusion  of  diminishing 
returns  in  the  logistic  specifications.  As  with 
the  other  contributions  this  paper  suggests 
an  area  meriting  further  discussion  in  the  near 
future,  with  our  best  use  of  marine  food  re- 
sources being  the  stake. 

Thompson  continues  the  parade  of  alternative 
functions  with  his  concern  being  the  absence 
of  a  proper  dynamic  component  within  the 
prevalent  fisheries  models.  To  correct  this 
error  he  proposes  the  marriage  of  the  Schaefer 
model  and  the  Thompson-George  (TG)  produc- 
tion-investment model.  He  also  suggests  some 
alterations  in  the  Schaefer  model. 

The  TG  model  replicates  the  sequence  of 
investment-production  decisions  which  are  in- 
volved in  the  operation  of  the  individual  fishing 
firm  (vessel).  Pertinent  stocks  and  flows  are 
specified  with  elaborate  preconditions  for  entry, 
though  there  are  no  provisions  for  entry  within 
the  decision  period,  an  interesting  trait  in  light 
of  the  Johnson  fixed  asset  theory  as  referred 
to  by  Stevens  and  Mattox  subsequently.  By 
adjoining  this  model  to  the  Schaefer  biological 
fluctuation  we  have  a  bioeconomic  model  which 
is  uniquely  micro  in  character;  the  dynamics  of 
change  in  the  fishery  stock  (and  hence  fishing 


success)  will  be  reflected  in  the  investment 
decision  of  the  sole  owner  as  the  limiting  case, 
and  vice  versa. 

This  method  avoids  the  critical  use  of  static 
methods  prevalent  in  economic  literature.  In- 
herently, the  adjustment  mechanism  in  the 
individual  owner  also  facilitates  the  modification 
of  the  Schaefer  function  to  incorporate  decreas- 
ing returns  to  effort,  as  discussed  by  Bell, 
Carlson,  and  Waugh  and  by  Rich  and  increasing 
returns  to  scale.  Relaxation  of  the  sole  owner 
condition  further  amplifies  the  critical  nature 
of  these  alterations  and  within  the  confines  of 
standard  economic  assumptions  reaffirms  the 
desirability  of  limiting  entry  and  suggests  an 
additional  method  of  measuring  the  critical 
management  variables. 

The  final  author  in  this  section  addresses 
the  problem  of  multiple  species  fisheries  —  or 
combination  vessels.  In  this  regard  three  issues 
are  of  prime  importance  to  Adam.  The  first  of 
these  relates  to  the  existence  of  yield  curves  for 
fisheries.  Adam  views  most  of  these  curves  as 
average  curves,  pointing  out  that  for  many 
fisheries  this  average  curve  will  be  bounded  by 
upper  and  lower  curves  which  are  usually  the 
result  of  substantial  fluctuations  in  either 
effort  and/or  recruitment.  The  average  curve 
is  essentially  a  product  of  a  stable  fishery  where- 
as the  boundary  curves  are  the  result  of  a 
rapidly  growing  fishery.  In  his  opinion  we  do 
not  move  along  the  average  curve  as  a  fishery 
rapidly  develops.  We  move  from  one  curve  to 
another,  somewhat  erratically  as  the  fishery 
develops.  He  looks  to  the  economist,  via  a 
function  akin  to  Carlson,  where  effort  is  value- 
dependent,  to  indicate  what  effort  will  be  in 
subsequent  years,  as  the  fisherman's  response 
to  his  monetary  success  is  one  of  the  few 
reliable  variables  which  can  be  presented  to  a 
biologist  in  such  a  dynamic  situation. 

His  second  point  extends  this  argument  to 
multiple  species.  If  a  vessel  has  the  capability 
to  adjust  his  harvesting  pattern  based  upon 
conditions  in  the  fishery  or  the  market,  this 
would  preclude  estimation  based  solely  on 
biological  factors.  It  suggests  that  many  of 
these  calculations  must  be  made  instantaneously, 
at  that  time  each  year  when  a  fishery  is  being 
initiated.  It  suggests  also  that  this  must  be 
done  for  several  fisheries  simultaneously  if 
those  fisheries  are  interrelated.  For  the  North- 


40 


east  Atlantic  this  is  increasingly  the  case.  be  closely  examined,  however,  so  that  we  may 

Adams's    final    related    point    concerns    the  maximize  their  comparability  and/or  ascertain 

measurement  of  fishing  effort.  Simply  stated,  which  measure  would  be  most  appropriate  for 

he  concludes  that  there  is  no  single  measure  each  circumstance, 
which  can  unequivocably  serve  the  needs  of  all 
the  disciplines.  These  different  measures  should  A.  A.  S. 


41 


Cross  Section  Production  Functions  for  North 

Atlantic  Groundfish  and  Tropical 

Tuna  Seine  Fisheries 


Ernest  W.  Carlson1 
ABSTRACT 

This  paper  explores  the  use  of  cross  section  production  functions  to  estimate  the 
fishing  power  of  individual  vessels.  The  problems  addressed  are:  The  proper  measurement 
of  output;  the  measurement  of  technological  change,  and  the  effect  of  location,  crew 
size  and  important  vessel  characteristics. 

Regression  analysis  upon  data  from  the  North  Atlantic  groundfish  fishery  and  the 
tropical  tuna  seine  fishery  yielded  highly  significant  results.  Many  of  the  hypothesized 
relationships  are  measurable  and  stable  with  relatively  small  errors.  The  tests  indicate 
that:  there  are  better  measures  of  output  then  total  pounds;  fishing  time  is  measured 
better  using  days  absent  rather  than  days  fishing;  the  use  of  more  vessel  characteristics 
improves  explanatory  power;  crew  size  can  be  an  important  variable;  the  effects  of 
location  can  be  measured;  and  technological  change  can  be  measured. 

The  production  functions  measured  can  then  be  used  as  inputs  in  devising 
management  schemes. 


INTRODUCTION 

One  of  the  more  difficult  problems  in  the 
management  of  fisheries  has  been  the  measure- 
ment of  vessel  productivity.  If  the  vessels  in  a 
fleet  were  physically  homogeneous  and  utilized 
for  the  same  amount  of  time  and  if  no  learning 
took  place,  the  problem  of  measuring  productivi- 
ty indices  would  be  less  difficult.  The  problem 
does  exist,  though,  because  vessels  are  far  from 
homogeneous.  For  example,  a  typical  fleet  may 
have  vessels  that  are  10  or  more  times  larger 
than  the  smallest  vessels  in  a  fleet.  Obviously, 
under  such  conditions  there  will  be  serious 
errors  introduced  if  attempts  are  not  made  to 
measure  the  productivity  of  different  vessels. 

To  handle  this  and  related  problems,  econo- 
mists have  developed  techniques  of  measurement 
that  fall  into  a  general  category  called  production 
functions.  One  of  the  important  attributes  of 
using  a  production  function  is  that  it  allows 
the    simultaneous    measurement    of   as    many 


1    Economist,  Economic  Research  Laboratory,  National 
Marine  Fisheries  Service. 


parameters  of  fishing  power  as  may  be  thought 
to  be  important  in  its  determination.  According- 
ly, production  functions  were  estimated  using 
data  from  the  New  England  trawl  fleet  and  the 
tropical  tuna  seine  fleet.  Many  problems  were 
considered  in  arriving  at  a  "best"  production 
function  for  these  fisheries. 

THE  PRODUCTION  FUNCTION  FOR 
A  FISHERY 

The  basic  assumption  of  this  paper  is  that  a 
production  function  can  adequately  describe 
the  relationship  between  inputs  and  outputs  in 
a  fishery.  The  production  function  is  a  technical 
or  engineering  relation  between  inputs  and 
outputs  and  is  the  base  upon  which  the  economic 
theory  of  supply  is  built.  Since  it  is  an  engineer- 
ing relationship,  considerations  such  as  prices 
and  costs  are  not  relevant  to  the  production 
function  itself.  The  schedule  of  maximum  output 
for  given  inputs  is  the  production  function  we 
are  trying  to  measure. 

The  classical  production  function  for  the 
individual  firm  is  usually  presented  as  follows: 


42 


x  =f(l,k,t), 

where        x  =    output, 

I  —    labor, 

k  =    capital, 

t  =    natural  resources. 

Output  (x)  is  measured  as  the  flow  of  goods  and 
services  during  an  accounting  period.  The  input 
variables  (I,  k,  t)  are  the  various  kinds  and 
qualities  of  labor,  capital,  and  natural  resources 
that  go  into  producing  the  output.  It  is  assumed 
that  a  given  set  of  inputs  produces  as  much 
as  possible. 

The  estimation  of  the  parameters  of  the 
production  function  is  accomplished  by  running 
a  regression  upon  a  cross  section  of  fishing 
vessels.  A  cross  section  is  a  sample  of  the  vessels 
in  a  fishery  for  a  fixed  time  period.  The  para- 
meters estimated  from  the  cross  section  will  give 
the  marginal  contribution  to  output  of  each 
variable  being  used  to  explain  output. 

We  will  discuss  the  variables  that  will  be 
used  in  the  production  function  in  the  following 
section. 


Output  in  a  Fishery 

Most  systems  for  measuring  relative  vessel 
productivity  have,  ultimately,  related  output  to 
some  fishing  vessel  characteristic.  The  basic 
problem  with  this  is  that  output,  when  using 
commercial  landings  statistics,  is  a  very  complex 
concept.  Except  in  extremely  simple  fisheries, 
fishermen  do  not  ordinarily  attempt  to  maxi- 
mize pounds  of  fish  landed.  One  working 
hypothesis  is  that  in  all  fisheries,  the  fishermen 
attempt  to  maximize  their  profits.  This  is  not 
necessarily  the  same  as  maximizing  total  pounds 
of  fish  landed.  Using  total  pounds  as  a  measure 
of  output  would  be  an  acceptable  measure  of 
output  (1)  where  there  is  a  single  species 
fishery  or  (2)  if,  in  a  multispecies  fishery,  the 
prices  of  the  target  species  are  approximately 
the  same  and  the  species  are  equally  catchable. 
In  the  general  case,  these  conditions  are  not  met. 

How  do  the  fishermen  decide  where  to  go  and 
what  to  catch  when  there  are  multiple  species 
in  a  fishery?  Again,  the  answer  to  this  question 
is  difficult.  Let  us  consider  two  models  of 
behavior  that  might  help  answer  this  question. 
In  the  first  type  of  fishery,  the  vessel  captains 


take  into  account  the  species  that  are  available, 
the  grounds  where  they  are  available,  the  prices 
for  which  they  can  be  sold,  and  the  expected 
catch  rates  for  their  vessels  on  the  grounds. 
Integrating  all  this  information,  the  captain,  if 
he  is  a  profit  maximizer,  will  decide  to  go  to 
the  grounds  and  fish  for  the  species  which  pro- 
vide the  highest  net  profit.  His  decision  may  or 
may  not  be  to  fish  where  the  catch  rates  are 
highest  or  for  those  species  that  bring  the 
highest  prices. 

We  have  been  discussing  this  as  if  the  choice 
were  always  between  species.  The  choice  can 
also  be  made  within  a  species,  such  as  a  decision 
to  fish  on  local  grounds  rather  than  on  distant 
grounds  where  the  catch  rates  are  higher.  In 
this  case,  the  higher  catch  rates  may  not  offset 
the  extra  running  time  necessary. 

If  this  abbreviated  discussion  is  an  adequate 
description  of  how  fishermen  behave  in  one 
type  of  fishery,  then  it  follows  that  we  may  not 
be  able  to  estimate  relative  vessel  productivity 
with  total  pounds,  but  must  rely  on  some  higher 
order  measure  such  as  the  value  of  catch. 

Value  was  considered  by  Gulland  (1956)  as 
a  measure  of  output  and  rejected  because  of  the 
variability  of  prices.  A  large  part  of  the  vari- 
ability of  fish  prices  is  due  to  the  seasonal 
availability  of  the  fish  themselves  with  prices 
moving  inversely  to  availability.  We  can  lessen 
the  objections  to  value  at  least  partially  by 
using  annual  data  so  that  the  interseasonal 
effects  of  availability  average  out.  Another 
alternative  would  be  the  estimation  of  relative 
efficiency  on  a  quarterly  basis. 

The  second  type  of  fishery  is  one  where  the 
location  of  the  fish  by  species  is  generally 
known,  but  where  there  is  considerable  mixing 
of  single  species  schools  in  the  same  area.  If 
locating  any  school  has  a  low  probability  per 
unit  time,  the  fishermen  will  attempt  to  catch 
all  that  they  can  of  those  they  do  locate.  In  this 
case,  the  fish  will  be  joint  products  of  the  fishery. 
If  the  fish  are  equally  catchable  and  their  prices 
are  not  too  different,  then  total  pounds  could 
be  the  measure  of  output.  If  they  are  not  equally 
catchable,  it  would  take  more  fishing  power  to 
catch  one  than  the  other.  In  such  a  case,  we 
might  have  to  utilize  a  modified  estimation 
scheme  to  arrive  at  a  proper  weighting  for 
output.  One  such  scheme  will  be  discussed  under 
the  statistical  section  on  tuna. 


41] 


Inputs  in  a  Fishery  —  Fishing  Time 

The  abstract  production  function  refers  to 
outputs  and  inputs  per  unit  of  time.  The  unit  of 
time  is  undefined.  When  using  annual  vessel 
data,  we  have  to  note  the  fact  that  the  vessels 
are  not  utilized  for  the  same  amount  of  time 
and  standardize  for  this. 

In  the  simple  case,  an  economist  would  prefer 
to  use  days  absent  from  port  as  a  measure  of 
fishing  time  rather  than  days  fishing.  If  a 
fisherman  is  an  economic  maximizer,  he  will 
attempt,  ceteris  'paribus,  to  maximize  his  gross 
revenue  per  day  at  sea  and  will  plan  his  fishing 
strategies  accordingly.  Under  this  assumption, 
the  fisherman  may  or  may  not  fish  when  or 
where  his  expected  catch  is  higher. 

The  theory  is  not  clear  as  to  how  time  should 
enter  the  production  function.  Two  basic 
specifications  are  possible: 

(1)  x=Dccf(l,k,t),  or 

(2)  x=Dif(l,k,t) 

There  are  theoretical  reasons  that  could  justify 
the  use  of  either.  Equation  (1),  with  D°;  can  be 
justified  if  we  hypothesize  that  the  fishermen 
makes  trips  of  varying  length.  Therefore,  we 
would  want  to  find  the  marginal  contribution 
of  an  extra  day  at  sea.  Equation  (2),  with 
D1,  can  be  justified  if  we  hypothesize  that  all 
inputs  are  being  used  to  produce  output  all  the 
time,  so  that  the  relationship  is  strictly  linear. 
Experiments  were  run  initially  in  both  forms, 
but  the  second  form  was  abandoned  for  what 
may  have  been  specious  reasons.  If  further  work 
is  done  the  alternative  specification  will  be 
tested  more  fully. 

Capital  —  The  Vessel  Characteristic 
Variables 

The  abstract  production  function  has  a  vari- 
able called  capital.  This  represents  the  di- 
mensions of  the  equipment  being  utilized.  In 
fishing,  the  individual  firms  and  many  of  the 
characteristics  of  their  capital  are  identifiable 
and  measurable. 

Vessel  size  has  been  recognized  as  a  deter- 
minant of  catch  and  is  explicitly  recognized  in 
most  of  the  productivity  measures  in  use. 
Beverton  and  Holt  (1957)  related  gross  tonnage 


to  fishing  power,  and  the  Inter-American  Tropi- 
cal Tuna  Commission  (IATTC)  focuses  on  the 
capacity  of  a  vessel's  freezers  (Shimada  and 
Schaefer  1956). 

Other  researchers  have  noted  that  there  are 
other  measures  of  vessel  size  that  are  correlated 
with  output,  among  them  horsepower  and  length. 
Gulland  (1956)  and  Noetzel  and  Norton  (1969) 
experimented  with  production  functions  that 
included  both  tonnage  and  horsepower.  Their 
results  showed  that  these  variables  may  make 
an  independent  contribution  to  output.  In 
fisheries,  the  possibility  of  independent  con- 
tributions should  not  be  overlooked  because 
there  may  be  a  tendency  for  vessel  configurations 
to  be  changed  in  such  a  way  that  fishing  power 
is  increased.  This  happens  especially  with  horse- 
power relative  to  gross  tonnage  as  old  engines 
are  replaced  and  also  as  new  vessels  are  built. 

The  role  of  horsepower  in  the  trawl  fleet 
appears  to  be  that  the  larger  the  engine,  the 
larger  the  net  that  can  be  dragged,  the  faster 
the  net  can  be  dragged,  or  the  deeper  the  water 
that  can  be  fished.  In  this  type  of  fishery,  the 
profit-maximizing  skipper  will  adjust  his  net 
to  obtain  the  "best"  results.  Although  it  has 
been  noted  that  trawlers  do  not  often  use  the 
full  power  of  their  engines,  a  larger  engine 
increases  the  number  of  possibilities  a  skipper 
can  consider  when  deciding  where  to  fish  and 
what  to  fish  for. 

In  a  seine  fishery,  the  role  of  horsepower  is 
less  clear,  except  that,  ceteris  paribus,  higher 
horsepower  increases  the  "search  power"  of 
the  vessel.  A  better  measure  of  this  search 
power  than  horsepower  would  appear  to  be 
running  speed.  The  only  way  to  obtain  this 
information  is  by  interview  or  sea  trials. 

Hull  construction  is  an  identifiable  parameter 
of  a  vessel.  Throughout  the  U.  S.  fisheries,  there 
has  been  an  increasing  tendency  to  build  new 
vessels  of  steel  rather  than  wood,  in  spite  of  the 
extra  initial  cost.  One  would  presume,  then, 
that  there  are  lower  operating  costs  for  steel, 
or  that  it  is  more  "productive."  It  is  possible 
to  test  for  the  effect  on  productivity  of  a  wood 
hull  by  creating  a  dummy  variable  that  takes 
on  the  value  "one"  if  the  hull  is  wood  and 
"zero"  otherwise. 

The  last  capital  input  variable  that  was  con- 
sidered was  age  of  the  vessel.  Most  people  would 
consider  older  vessels  less  productive,  ceteris 


44 


paribus,  than  newer  vessels.  It  is  rather  simple 
to  test  this  hypothesis  by  including  in  the  tests 
the  age  of  the  vessels. 

Hence,  the  dimensions  of  the  capital  input 
will  be  measured  by  (1)  gross  tonnage,  (2)  horse- 
power, (3)  construction  materials,  and  (4)  age 
of  the  vessel. 

Labor  —  The  Crew 

Crew  size  could  also  be  tested  as  an  input 
variable  in  the  production  function.  It  seems 
reasonable  that  a  larger  crew  would  produce  a 
higher  output,  and  this  should  be  tested. 

One  need  not  work  in  fisheries  very  long 
before  he  is  made  cognizant  of  the  "good  captain 
hypothesis."  That  is,  the  catch  of  a  vessel  depends 
as  much  upon  the  managerial  skill  of  the  captain 
and  crew  as  it  does  upon  the  characteristics 
of  the  vessel.  As  such,  there  is  no  way  to  test 
this  hypothesis. 

One  might  attempt  to  test  the  good  captain 
hypothesis  by  using  the  years  of  schooling  or 
the  years  of  experience  of  the  captain  to  arrive 
at  a  proxy  for  his  skill.  One  may  suspect  on 
economic  grounds  that  the  best  captains  would 
gravitate  to  the  best  vessels  because  they  would 
be  able  to  buy  the  more  productive  vessels  or 
be  hired  away  from  the  poorer  vessels.  In  other 
words,  part  of  the  higher  output  of  a  larger 
vessel  may  not  be  due  to  its  hardware  but  to 
the  superior  men  running  it.  In  this  analysis 
we  are  restricted  to  crew  size  as  one  measurable 
variable. 

Location 

The  production  function  provides  for  the  dif- 
ferential productivity  that  could  be  due  to 
location  with  respect  to  the  fishing  grounds 
through  the  variable  called  land.  Vessels  from 
some  ports  could  have  higher  productivity  than 
vessels  from  other  ports  by  being  located  closer 
to  the  better  grounds.  Since  these  locations 
cannot  be  appropriated,  the  vessels  will  allocate 
themselves  between  ports  so  that  effects  on  net 
profits  will  be  dissipated.  It  is  possible  to  test 
whether  certain  locations  are  more  productive 
by  creating  dummy  variables  that  correspond 
to  home  ports.  If  their  coefficients  are  statistically 
significant,  then  a  location  may  be  either  more 
or  less  productive  than  the  average  location. 


Technological  Change 

One  of  the  major  problems  encountered  in  the 
management  of  fishing  power  has  been  the 
difficulty  in  adjusting  for  technological  change. 
Attempts  have  been  made  to  adjust  for  techno- 
logical change,  but  on  the  whole  they  have  been 
less  than  satisfactory. 

The  test  for  the  added  productivity  of  an 
innovation  should  be  done  when  the  fleet  is 
in  a  period  of  transition  from  the  use  of  the 
old  to  the  new  technique.  This  method  will 
hold  abundance  and  availability  constant  and 
therefore,  all  vessels  will  have  the  same  op- 
portunities. Bell  (1966)  used  a  dummy  variable 
to  measure  the  increased  productivity  due  to 
stern  trawling.  He  created  a  variable  that  was 
1  if  a  vessel  was  a  stern  trawler  and  0  if  it  was  a 
side  trawler.  The  coefficient  of  the  dummy 
variable  was  the  added  productivity  due  to 
stern  trawling. 

This  technique  can  be  used  to  test  the  added 
productivity  of  any  innovation,  for  example,  a 
new  electronic  instrument  or  the  use  of  spotter 
planes  or  maybe  even  the  use  of  a  radically  new 
technique  such  as  switching  from  bait  boats  to 
purse  seining.  The  added  productivity  of  a  new 
technique  would  thus  become  a  permanent 
attribute  of  the  vessels  even  after  it  was  no 
longer  possible  to  measure  the  contribution  of 
the  technique,  i.e.,  even  after  it  was  universally 
adopted. 

THE  DATA 

The  New  England  Trawl  Fishery 

The  National  Marine  Fisheries  Service 
(NMFS)  has  collected  comprehensive  data  on 
the  landings  of  the  New  England  trawl  fleet 
for  many  years.  The  data  consist  of  landings 
information  by  trip.  The  following  information 
is  noted  for  each  trip: 

1.  Official  number 

2.  Departure  date 

3.  Arrival  date 

4.  Number  of  days  fishing 

5.  Grounds  fished 

6.  Pounds  landed,  by  species 

7.  Price/pound  by  species 

The  data  are  stored  on  magnetic  tapes  and 
can  be  manipulated  with  a  digital  computer. 


45 


The  data  used  were  for  the  years  1964,  1965, 
and  1967.  The  data  were  aggregated  by  vessel 
for  the  whole  year.  For  each  vessel,  the  following 
information  was  produced: 

1.  Days  at  sea 

2.  Days  fishing 

3.  Total  trips 

4.  Days  at  sea  by  calendar  quarter 

5.  Days  fishing  by  calendar  quarter 

6.  Trips  to  major  areas:   offshore,  inshore, 
off  Canada 

7.  Pounds  caught,  by  major  species 

8.  Value,  by  major  species 

9.  Total  pounds  caught 

10.  Total  value 

This  information  was  augmented  by  the 
addition  of  information  from  the  Merchant 
Vessels  of  the  United  States  (1965),  including: 

11.  Gross  tons 

12.  Horsepower 

13.  Hull  construction 

14.  Year  built 

Information  from  National  Marine  Fisheries 
Service  files  was  added  on: 

15.  Crew  size 

16.  Home  port 

Vessels  with  total  landings  valued  at  less 
than  $10,000  were  excluded  from  the  sample; 
we  made  the  assumption  that  these  were  casual 
fishermen.  There  were  about  120  vessels  excluded 
per  year,  accounting  for  3%  of  New  England 
landings.  Otherwise,  no  editing  was  done; 
therefore,  the  sample  contains  all  trips,  in- 
cluding brokers.  Thus,  the  estimates  have  built 
into  them  all  conditions  that  vessels  from  this 
fleet  experience  on  the  North  Atlantic.  The  total 
sample  consisted  of  about  383  vessels  per  year 
or  1,149  vessel  years. 

The  Tropical  Tuna  Purse 
Seine  Fleet 

The  Inter-American  Tropical  Tuna  Commis- 
sion (IATTC)  kindly  let  us  transcribe  landings 
data  from  their  files  for  the  years  1966,1967, 
and   1968.  The  data  were  for  the  whole  year 


for  the  full-time  purse  seiners.  The  data  trans- 
cribed were  as  follows: 


1.  Official  number 

2.  Days  at  sea 

3.  Landings  by  species 

4.  Major  area  fished:  Atlantic  or  Pacific 

This  information  was  supplemented  by  the 
addition  of  information  from  the  Merchant 
Vessels  of  the  United  States  (1965)  including: 

5.  Gross  tons 

6.  Horsepower 

7.  Length 

8.  Year  built 

Finally,  the  following  information  was  added: 

9.  Capacity  (American  Tunaboat 
Association) 

10.  Crew  size  (NMFS  files) 

The  total  sample  consisted  of  89  vessels  per 
year  or  267  vessel  years.  The  data  were  divided 
into  two  periods:  (1)  when  there  was  unrestricted 
fishing  for  yellowfin  and  (2)  when  yellowfin  was 
restricted  to  15%  of  the  total  catch.  The  data 
from  the  restricted  season  were  not  used  in  the 
analysis  because  of  the  different  conditions 
following  the  season  closure. 

THE  STATISTICAL  RESULTS 
Overall  Results 

The  statistical  results  of  these  experiments 
are  quite  encouraging.  It  is  possible  to  explain 
very  high  variations  in  catch  with  a  minimum 
of  information.  In  the  tropical  tuna  fishery  we 
can  explain  approximately  70%  of  the  variation 
in  the  dependent  variable,  and  in  the  New 
England  trawl  fishery,  approximately  84%  . 

Tests  for  heteroscaedasticity  showed  that  it 
existed  in  the  linear  equations.  When  it  is 
present,  we  have  inefficient  estimators.  Log- 
arithmic transformation  of  the  variables  in  both 
fisheries  removed  this  problem.  Results  in  both 
forms  are  reported,  but  only  the  logarithmic 
results  are  suitable  for  analytical  work. 

Several    regression    experiments    were    run 


46 


using  a  single  year's  observations  in  both 
fisheries  on  the  same  variables.  The  results 
were  very  encouraging  in  that  there  was  a  high 
degree  of  stability  in  the  coefficients  and  their 
t  ratios.  These  stable  results  were  obtained  in 
fisheries  which,  if  anything,  are  notorious  for 
their  variability  in  almost  all  aspects:  biological, 
economic,  atmospheric,  and  oceanographic.  Some 
results  illustrating  this  stability  for  the  trawl 
fishery  are  shown  in  Appendix  1. 

The  New  England  Trawl  Fleet 

The  statistical  results  for  the  New  England 
trawl  fishery  were  very  good.  The  overall  "fit" 
of  the  data  in  the  equations  was  very  high, 
especially  when  one  considers  the  heterogeneity 
of  this  fleet.  The  equations  are  rich  in  informa- 
tion in  that  many  of  the  variables  about  which 
hypotheses  were  made  were  statistically  sig- 
nificant with  the  right  signs. 

Because  of  the  unclear  nature  of  variables 
discussed,  the  equations  were  run  using  the 
alternatives  for  the  same  variables  where  pos- 
sible. This  will  allow  direct  comparison  of  the 
results.  In  a  sense,  we  shall  permit  the  data 
to  decide  which  are  better  variables.  We  will 
briefly  run  through  the  results  according  to  the 
topics  covered  in  the  theoretical  section. 

The  following  general  production  function 
was  established  for  the  New  England  trawl 
fleet: 

(3)        0    =  /(FT,  GRT,  HP,  CR,  AGE,  C,  PT) 
where        O  =    output,  either  total  pounds 

or  total  value, 
FT       =    fishing    time,    either    days 

fished  or  days  absent, 
GRT    =    gross  registered  tonnage, 
HP       =    horsepower, 
CR       =    crew  size, 
AGE    =    age  of  the  vessel, 
C  =    construction,  1  if  wood,  0 

otherwise, 
PT       =    homeport  dummy  variables. 

The  equations  providing  the  best  results  are 
shown  in  Table  1.  These  equations  will  be 
discussed  below.  A  more  complete  set  of  regres- 
sions is  shown  in  Appendix  Table  1. 

The  tests  of  whether  total  value  or  total 
pounds  was  the  better  measure  of  output  in  this 


fishery  are  shown  in  Problems  1  through  4.  The 
measures  of  overall  fit  (R2)  are  lower  in  Problems 
1  and  2,  which  use  total  pounds  as  the  dependent 
variable  (0.40  and  0.54),  than  in  Problems  3  and 
4,  which  use  total  value  as  the  dependent 
variable  (0.83  and  0.83).  Thus,  the  fishermen 
appear  to  have  implicitly  taken  into  account 
expected  prices,  expected  catch  rates,  and 
steaming  time  to  the  grounds  and  made  deci- 
sions as  to  where  to  go  and  what  to  fish.  Hence, 
relative  total  revenue  appears  to  reflect  the 
fishing  power  of  New  England  vessels.  The 
more  fishing  power,  the  higher  revenues  are 
expected  to  be. 

The  most  powerful  explanatory  variables 
for  either  total  pounds  or  total  value  were  the 
fishing  time  variables.  That  is,  the  more  days 
fished  or  days  absent,  the  higher  the  total  value 
and  total  pounds.  On  the  basis  of  contributions 
to  the  overall  goodness  of  fit,  there  is  no  way 
to  choose  between  these  two  variables.  Our 
choice,  therefore,  will  have  to  rest  upon  their 
effects  on  other  variables  and  on  the  cost  of 
gathering  the  information. 

In  Problem  3,  using  total  value  as  the 
dependent  variable  and  days  fishing  as  the 
measure  of  fishing  time,  crew  size  becomes 
statistically  nonsignificant  and  negative.  In 
Problem  4,  when  days  absent  is  used,  crew  size 
becomes  statistically  significant  and  a  very 
powerful  explanatory  variable.  Days  fishing 
appears  to  be  a  less  desirable  measure  of  fishing 
time  in  that:  (1)  It  is  theoretically  inferior  on 
economic  grounds  as  discussed  previously; 
(2)  it  causes  other  important  variables  to  have 
the  wrong  sign;  (3)  it  costs  more  money  to 
collect  this  information;  and  (4)  it  is  probably 
more  subject  to  error. 

The  vessel  size  variables  used  were  gross 
registered  tonnage  (GRT)  and  horsepower  (HP). 
GRT  was  the  more  powerful  of  these  variables 
as  it  was  statistically  significant  in  all  equations 
and  explained  a  large  part  of  output.  HP  was 
not  as  powerful  a  variable  in  terms  of  its  partial 
correlation  coefficient.  However,  it  was  statisti- 
cally significant  when  total  value  was  the  de- 
pendent variable,  indicating  that  it  made  an 
independent  contribution  to  fishing  power. 

The  variable  that  indicated  the  age  of  a 
vessel  had  a  negative  coefficient  and  was  sta- 
tistically significant  in  most  cases.  There  are 
at   least   three   hypotheses   why   older   vessels 


47 


Table  1.  —  New  England  trawler  production  functions:  alternate  specifications. 


INDEPENDENT  VARIABLE 

Dependent  variable 

LOG  DAYS 

LOG  DAYS 

LOG 

LOG 

LOG 

LOG 

CONSTRUC- 

DUM 

DUM 

Y 

n  2 

ABSENT 

FISHED 

GRT1 

HP2 

CREW 

AGE 

TION3 

654 

674 

INT 

R 

F 

Problem  1 

Log  total  pounds  (All  years) 

Reg.  Coef. 

/.649 

.409 

.038 

-.410 

-.240 

-.138 

-.018 

-.084 

4.69 

.405 

98.70 

t  ratio 

18.300 

6.340 

.525 

5.160 

4.540 

3.780 

.776 

3.420 

Part.  Cor.  Coef.  s 

.477 

.184 

.016 

-.151 

-.133 

-.111 

-.022 

-TOO 

Problem  2 

Log  total  pounds  (All  years) 

Reg.  Coef. 

1.060 

.429 

.002 

-2.66 

-.207 

-.024 

.011 

-.059 

3.39 

.542 

170.28 

t  ratio 

27.800 

7.580 

.037 

4.040 

4.470 

.752 

.533 

2.750 

Part.  Cor.  Coef.  5 

.636 

.219 

.001 

-.119 

-.131 

-.022 

.015 

-.081 

Problem  3 

Log  total  value  (All  years) 

Reg.  Coef. 

.886 

.365 

.113 

-.002 

-.107 

-.043 

-.024 

.0006 

2.43 

.834 

724.34 

t  ratio 

47.900 

10.800 

2.980 

.062 

3.860 

2.280 

1.920 

.0500 

Part.  Cor.  Coef.  s 

.817 

.305 

.088 

-.001 

-.113 

-.067 

-.057 

.0010 

Problem  4 

Log  total  value  (All  years) 

Reg.  Coef. 

1.080 

.373 

.074 

.347 

-.129 

.095 

.023 

.010 

1.44 

.833 

718.97 

t  ratio 

47.600 

11.000 

1.940 

8.830 

4.660 

5.000 

1.790 

.855 

Part.  Cor.  Coef.  5 

.815 

.309 

.058 

.253 

-.136 

.146 

.053 

.025 

'Gross  registered  tonnage. 

2  Horsepower. 

Construction;  equals  one  if  wood,  zero  otherwise. 

4  Dummy  variables  for  year  of  observation. 

5  Partial  correlation  coefficient. 


may  be  less  productive:  (1)  Older  vessels  might 
tend  to  have  more  breakdowns  and  equipment 
that  was  not  in  the  best  working  order;  (2)  older 
vessels  might  have  poorer  working  conditions 
and  accommodations  and,  therefore,  attract  less 
able  crews;  (3)  older  vessels  may  embody  older 
technologies.  If  the  last  hypothesis  is  dominant, 
vessels  do  not  become  less  productive  as  they 
get  older,  rather  old  vessels  are  less  productive. 
This  would  have  different  implications  than 
the  first  hypothesis  when  fishing  power  factors 
are  computed. 

The  dummy  variable  created  for  hull  con- 
struction took  on  the  value  1  if  the  hull  was 
wood  and  0  if  steel.  The  results  using  this  vari- 
able were  mixed.  In  Problem  4,  using  total 
value  and  days  absent,  it  was  positive  and 
significant.  This  may  mean  that  ceteris  paribus 
wooden  hulls  are  25%  more  productive.2  There 
is    no    theoretical    reason    why    these    results 

2  The  antilog  of  1  is  10.  We  have  10. 095  which  equals 
1.25.  Therefore,  a  wooden  hull  is  25%  more  productive. 


should  be  obtained.  The  data  in  Appendix  Table 
3  show  that  the  large  vessels  in  the  fleet  are 
steel  and  the  small  ones  wood,  with  a  very  small 
overlap.  We  may  be  observing  an  upward 
adjustment  for  the  wood  vessels  because  they 
fish  many  fewer  days  during  the  most  productive 
portion  of  the  year. 

The  tests  for  locational  differences  in  produc- 
tivity were  made  by  creating  an  array  of  six 
dummy  variables,  one  for  each  of  the  major 
ports  in  New  England.  A  "one"  was  placed  in 
proper  location  in  the  array  corresponding  to  a 
vessel's  home  port  and  a  "zero"  in  all  the 
others.  Equations  showing  the  results  of  these 
tests  are  given  in  Appendix  Table  1.  In  the 
logarithmic  forms  of  the  equations,  there  are 
no  consistent  differences  between  ports  when 
total  value  is  the  dependent  variable,  the  ports 
designated  "Maine"  appear  to  catch  significantly 
more  and  "Boston"  significantly  less  (Problem 
10).  These  differences  appear  because  Maine 
specializes  in  low  value  species  and  Boston  in 


48 


high.  When  weighted  by  value,  these  differences 
disappear. 

On  the  basis  of  these  statistical  tests,  we 
conclude  that  the  best  specification  of  the  pro- 
duction function  for  the  New  England  ground- 
fish  fleet  is  shown  in  Problem  4,  where  total 
value  is  the  measure  of  output  and  days  absent 
is  the  measure  of  fishing  time.  Good  descriptions 
of  the  capital  variable  are  given  by  gross 
registered  tonnage,  horsepower,  vessel  age,  and 
construction  materials.  The  contribution  of 
labor  is  measurable  and  important. 

The  Tuna  Seine  Fleet 

In  fisheries  such  as  the  tropical  tuna  fishery, 
the  species  are,  in  the  jargon  of  the  economist, 
"joint  products."  That  is,  the  fishermen  take  as 
much  of  both  species  (yellowfin  and  skipjack) 
as  they  can  in  an  effort  to  fill  their  holds  as 
quickly  as  possible.  They  are  essentially  indis- 
criminate between  tunas  in  that  they  do  not 
appear  to  pass  up  any  that  they  sight  solely 
because  it  is  the  less  desirable  species,  although 
such  behavior  was  noted  up  to  about  1950 
(Shimada  and  Schaefer,  1956). 

According  to  IATTC  records,  the  probability 
of  a  successful  set  on  yellowfin  is  higher  than 
on  skipjack.  This  leads  one  to  hypothesize 
that  a  ton  of  skipjack  represents  in  some  way 
more  input  than  a  ton  of  yellowfin  because  it 
takes  more  work  to  catch  skipjack.  There  are  at 
least  two  techniques  that  might  be  used  in  this 
fishery  to  determine  a  weighting  system  for 
output.  One  technique  (which  is  not  used  here) 
is  canonical  regression  which  was  developed  by 
Hotelling  and  described  by  Tintner  (1952).  In 
a  sense,  it  is  a  search  technique  that  "weights" 
the  dependent  and  independent  variables  in 
such  a  way  that  the  sum  of  the  squares  of  the 
unexplained  variance  of  all  the  variables  is 
minimized.  The  second  technique3  is  to  sys- 
tematically try  different  weights  (whose  sum 
is  one)  for  the  dependent  variable  and  run  a 
series  of  regressions  using  a  common  set  of 
independent  variables.  The  regression  that 
maximizes  the  coefficient  of  determination  would 
have  the  weights,  which  are,  in  a  sense,  best. 


3  Suggested    by    Henri    Theil    during   a   discussion    of 
this  problem  with  the  author. 


The  following  regression  was  run  in  an  at- 
tempt to  arrive  at  the  best  weighting  system 
for  output: 

(4)  Q   =  f(D,  T,  CAPAC,  GRT,  ND,  PR, 

CR,AGE,HP) 
where        Q  =  (aY  +  |3S  +  5B)  and  (a  +  0  +  5)  =  1 

and  Y  is  tons  of  yellowfin  landed, 

S  is  tons  of  skipjack  landed, 
B  is  tons  of  bluefin  landed, 
D  is  days  at  sea  of  each  vessel, 
T  is  the  number  of  trips  of  each  vessel, 
CAPAC  is  the  capacity  of  each  vessel, 
GRT  is  the  gross  registered  tonnage, 
ND  is  a  dummy  for  new  design, 
PR  is  1  for  Puerto  Rico  home  port, 

zero  otherwise, 
CR  is  the  crew  size, 
AGE  is  the  age  of  the  vessel, 
HP  is  the  horsepower  of  each  vessel. 

The  results  of  this  experiment  are  shown  in 
Table  2,  where  the  left  hand  column  shows  the 
different  weights  applied  to  each  species.  The 
column  headings  are  for  each  year's  observations 
and  for  pooled  observations.  Tests  using  the 
H  statistic  show  that  the  observations  are  not 
random.  Weights  of  .3  for  yellowfin,  .4  for 
skipjack,  and  .3  for  bluefin  are  best.  This  fits  our 
a  priori  expectation  that  a  vessel  exhibited 
more  productivity  when  it  caught  a  ton  of 
skipjack  than  a  ton  of  yellowfin.  The  statistical 
results  indicate  that  a  vessel  does  one-third 
more  work  to  catch  a  ton  of  skipjack  than  a 
ton  of  yellowfin. 

The  above  experiment  presents  one  approach 
to  the  determination  of  output  in  a  fishery. 
Three  alternative  specifications  of  output  in 
the  tuna  fishery  were  used  in  estimating  the 
production  function.  These  specifications  were 
as  follows:  total  value,  total  pounds,  and 
weighted  total  pounds  using  the  weights 
determined  above. 

Selected  results  of  the  regression  experiments 
run  are  shown  in  Table  3  and  in  Appendix 
Table  2.  The  various  specifications  of  the 
dependent  variable  could  be  explained  with 
varying  degrees  of  precision.  As  expected, 
weighted  total  pounds  had  the  highest  coef- 
ficient of  determination,  followed  by  total 
pounds,  total  value,  skipjack  and  yellowfin,  in 
that  order.  The  actual  difference  between  co- 


41) 


Table  2.  —  Regression  results  using  various  weights  for 
tuna  species  holding  independent  variables  constant. 

Weights  of  yellowim.  1966         1967         1968         All_years 

skipjack,  and  bluefin  Rz  R2  R2  R2 


". 

1.. 

.2 

.559 

.332 

.697 

.486 

.6. 

.1. 

.3 

.573 

.351 

.701 

.505 

.6. 

.2. 

2 

.650 

.542 

.731 

.612 

.5. 

.1. 

.4 

.588 

.380 

.705 

.531 

.5. 

.3. 

.2 

.730 

.785 

.758 

.757 

.5. 

.2. 

.3 

.677 

.622 

.739 

.652 

.4. 

.1. 

.5 

.598 

.426 

.711 

.565 

.4. 

.4. 

2 

.772 

.873 

.775 

.779 

.376. 

.286. 

.344 

.756 

.837 

.767 

.763 

.4. 

.2. 

.4 

.703 

.711 

.748 

.698 

.4. 

.4. 

.3 

.756 

.837 

.767 

.760 

.3. 

■5, 

.2 

.770 

.884 

.778 

.776 

.3. 

2 

.5 

.707 

.790 

.757 

.740 

.3. 

4. 

.31 

.775 

.883 

.778 

.785 

.3, 

■  3, 

.4 

.764 

.868 

.774 

.783 

.2. 

•3, 

.5 

.723 

.875 

.774 

.775 

.2. 

.5, 

.3 

.744 

.877 

.769 

.757 

■  2, 

4. 

.4 

.745 

.879 

.774 

.769 

•2, 

■  2, 

.6 

.646 

.833 

.762 

.748 

-3, 

•  1, 

.6 

.584 

.494 

.715 

.603 

•  2, 

.1, 

.7 

.523 

.572 

.713 

.619 

'The  "Best"  solution. 

Source:    Economic   Research   Laboratory,   National   Marine 


efficients  of  determination  in  the  weighted 
total  pounds  equation  and  the  total  pounds 
equation  is  not  statistically  significant  (0.70 
vs.  0.68). 

The  total  pounds  variable  has,  of  course, 
almost  the  same  weights  (Vb,  Vs,  V3)  as  the 
weighted  output  variable  so  that,  ultimately, 
it  may  be  of  marginal  significance  to  distinguish 
between  them  in  this  fishery.  Nevertheless, 
we  cannot  know  this  before  further  experiments 
are  conducted. 

Total  value  as  a  dependent  variable  is  inferior 
to  total  pounds.  This  tends  to  confirm  our 
hypothesis  that  yellowfin  and  skipjack  are  joint 
products  in  this  fishery.  The  weight  of  skipjack 
in  total  value  is  less  than  the  weight  for 
yellowfin  and  bluefin.4  Therefore,  it  appears 
that  the  amount  for  which  skipjack  can  be  sold 
is  not  reflected  in  the  extra  work  done  in 
catching  it,  at  least  relative  to  yellowfin  and 
bluefin. 

The  best  production  functions  for  the  tuna 
fishery  are  shown  in  Table  3.  The  only  fishing 
time   variable    available   for   this   fishery   was 


Fisheries  Service.  1970. 


4  The    relative   price    weights    are    .286    for   skipjack, 
.376  for  yellowfin,  and  .344  for  bluefin. 


Table  3.  —  Tuna  purse  seine  production  function:  alternate  specifications. 


INDEPENDENT  VARIABLE 


Dependent  variable  LOG  CAPACITY      LOG  DAYS  LOGH.P.1         66  DUM         67  DUM  YINT.       R 


Problem  1 


Log  total  value 

Reg.  coef. 

.365 

.310 

.368 

.067 

.044 

t  ratio 

5.14 

3.32 

4.66 

2.08 

2.21 

Part.  Cor.  Coef. 

.303 

.201 

.277 

.128 

.136 

Problem  2 

Log  total  pounds 

Reg.  Coef. 

.438 

.373 

.339 

-.024 

.049 

t  ratio 

7.39 

4.79 

5.15 

.914 

2.94 

Part  Cor.  Coef. 

.416 

.284 

.304 

-.056 

.179 

Problem  3 

Weighted  total  pounds 
Reg.  Coef. 
t  ratio 
Part  Cor.  Coef. 


.520 
8.41 
.462 


.416 
5.12 
.302 


.328 

4.77 
.283 


026 

.065 

946 

3.71 

058 

.224 

.196 


.587 


76.17 


.453 


.680 


113.84 


.168 


.704 


127.07 


1  Horsepower 

Source:  Economic  Research  Laboratory,  National  Marine  Fisheries  Service,  1970. 


50 


days  absent  so  that  alternative  specifications 
of  the  equations  could  not  be  run.  Days  absent, 
however,  was  not  as  important  a  variable  in 
this  fishery  as  in  the  trawl  fishery.  The  reason 
for  this  may, be  that  there  is  a  basic  difference 
in  the  way  the  vessels  in  these  fisheries  operate. 
The  trawl  fishery  is  a  wetfish  fishery  so  that  the 
vessels  are  constrained  by  time  when  they  go 
to  sea,  whereas  the  tuna  boats  are  freezers, 
and  they  stay  at  sea  until  their  holds  are  filled; 
hence,  there  is  a  different  connotation  to  the 
fishing  time  variable. 

The  vessel  size  variables  used  in  the  final 
equation  were  capacity  and  horsepower.  Capacity 
was  the  more  important  of  these  variables.  This 
indicates  that  the  industry  is  justified  in  using 
capacity  as  an  index  of  a  vessel's  fishing  power. 
Several  tests  were  run  with  gross  tonnage  in 
place  of  capacity,  but  the  results  were  not  as 
good,  although  they  were  still  meaningful. 

Horsepower  makes  an  important  independent 
contribution  to  explanation  of  output.  The  con- 
tribution of  horsepower  to  the  increase  in  the 
coefficient  of  determination,  though  small  at 
any  point  in  time,  may  be  important  in  the 
maintenance  of  an  effort  series  as  the  com- 
position of  a  fleet  changes. 

Tests  were  run  using  crew  size  but  results 
were  poor,  presumably  because  there  is  such 
small  variation  of  crew  in  this  fleet  (12-14 
men).  In  addition,  crew  size  is  defined  by  custom 
and  union  contract  according  to  the  capacity 
of  a  vessel,  hence  crew  size  does  not  give 
additional  information. 

The  tuna  fleet  has  two  main  bases:  Puerto 
Rico  and  southern  California.  To  test  whether 
vessels  located  in  Puerto  Rico  were  more  pro- 
ductive, a  dummy  variable  was  created  that  took 
the  value  one  if  a  vessel's  home  port  was  Puerto 
Rico  and  was  zero  otherwise.  The  results  were 
generally  positive  but  not  statistically  signifi- 
cant. This  indicates  that  the  fleet's  shift  toward 
Puerto  Rico  is  because  of  reasons  other  than 
catching  more  fish  (see  Appendix  Table  2). 

Tests  to  see  if  the  age  of  the  vessels  could 
explain  some  of  the  variation  in  output  generally 
showed  that  older  vessels  were  less  productive 
in  the  linear  forms  of  the  equations.  When  the 
logarithmic  transformations  were  made,  the 
age  variable  became  nonsignificant;  hence,  it 
is  not  included  in  the  final  equations. 

The  original   purse   seine   fleet  consisted   of 


vessels  converted  from  either  military  craft 
or  bait  boats.  There  has  been  a  major  expansion 
of  this  fleet  since  1963  with  vessels  designed 
specifically  for  purse  seining.  To  see  if  these 
vessels  were  superior  in  a  way  that  could  not 
be  accounted  for  by  either  horsepower  or  capaci- 
ty, a  dummy  variable  was  created  that  took  the 
value  one  if  a  vessel  were  built  after  1962  and 
zero  if  built  before  1963.  It  was  hoped  that 
this  would  pick  up  technological  change.  The 
results  using  this  were  generally  positive  and 
sometimes  statistically  significant,  but  the 
dummy  variable  is  not  included  in  the  final 
equations  because  it  was  not  statistically  sig- 
nificant in  them. 

We  conclude  that  for  the  tuna  fishery  the 
best  production  function  is  given  by  Table  3, 
Problem  3,  where  weighted  total  pounds  is  the 
dependent  variable,  days  absent  is  the  measure 
of  fishing  time,  and  capacity  and  horsepower 
are  measures  of  the  capital  used. 


CONCLUSION 

The  basic  assumption  underlying  this  work 
is  that  a  production  function  can  adequately 
describe  the  productivity  of  vessels.  The  stability 
of  the  estimates  arrived  at  using  this  technique 
rely  most  upon  the  constant  patterns  of  economic 
behavior.  The  coefficients  would  have  to  be 
re-estimated  if  the  ratio  of  days  absent  to  days 
fishing  changed  significantly  in  a  fleet,  or  if 
the  form  of  regulation  changed  the  pattern  of 
fishing.  Pattern  changes  are  undoubtedly  taking 
place  in  the  tuna  fishery  where  the  quota  system 
of  regulation  makes  it  imperative  for  vessels 
to  leave  the  home  port  the  day  the  season  opens 
and  to  fish  as  intensively  as  possible.  This  makes 
vessel  utilization  in  the  first  part  of  the  year 
much  higher  than  it  has  been  historically  or 
would  be  without  the  quota  regulation  system. 
It  has  probably  had  the  effect  of  changing  the 
effective  productivity  markedly  by  putting  a 
premium  upon  running  speed. 

Once  an  estimating  equation  has  been  deter- 
mined suitable,  it  should  be  used  as  long  as 
possible,  say  up  to  10  years  to  provide  continuity. 
Checks  should  be  made  periodically  to  see  if 
the  equation  being  used  is  still  appropriate. 

The  technique  of  using  dummy  variables  to 


51 


measure  technological  change  can  be  a  very 
powerful  means  of  keeping  productivity  indices 
up  to  date.  Any  new  device,  strategy,  or  vessel 
design  can  be  tested  for  its  ability  to  increase 
productivity  as  it  is  being  introduced  and 
therefore,  can  be  permanently  built  into  the 
vessel  productivity  indices. 

One  of  the  more  important  attributes  of  these 
production  functions  is  that  they  provide  a 
simple  way  to  test  whether  information  being 
gathered  is  relevant  to  the  task  at  hand.  For 
example,  fishing  days  are  collected  in  New 
England.  Upon  further  testing  it  may  be 
decided  that  this  information  is  not  worth 
its  cost. 

The  technique  can  also  provide  a  way  to 
handle  some  of  the  causes  of  secular  changes  in 
the  productivity  of  a  fleet.  For  example,  in 
both  of  the  fleets  considered,  both  vessel  size 
(GRT  and  capacity)  and  horsepower  made 
significant  contributions  to  the  determination 
of  productivity.  Thus,  as  new  vessels  are  added 
to  a  fleet,  their  productivity  can  be  estimated 
even  though  they  have  larger  engines  relative 
to  vessel  size  than  other  vessels  in  their  size 
class.  It  is  also  possible  to  keep  estimates  of 
productivity  current  as  the  engines  of  old 
vessels  are  replaced  or  upgraded  and  changes 
in  crew  size  are  made. 


LITERATURE  CITED 

BELL.  F.  W.  1966.  The  Economics  of  the  New  England 
Fishing  Industry:  The  Role  of  Technological  Change 
and  Government  Aid,  Research  Report  to  the  Federal 
Reserve  Bank  of  Boston,  No.  31.  Boston:  Federal 
Reserve  Bank  of  Boston.  216  pp. 

BEVERTON,  R.  J.  H.,  and  S.  J.  HOLT.  1957.  On  the 
Dynamics  of  Exploited  Fish  Populations.  Ministry  of 
Agriculture,  Fish  and  Food  (U.  K.),  Fishery  Investiga- 
tions, Ser.  II  (19):  533  pp. 

GULLAND,  J.  A.  1956.  On  the  Fishing  Effort  in  English 
Demersal  Fisheries.  Ministry  of  Agriculture,  Fish  and 
Food  (U.  K.),  Fishery  Investigations,  Ser.  II  (20): 
44  pp. 

Merchant  Vessels  of  the  United  States.  1965.  U.  S. 
Department  of  the  Treasury.  Washington,  D.  C:  U.  S. 
Government  Printing  Office. 

NOETZEL,  B.  G.,  and  V.  J.  NORTON.  1969.  Costs  and 
Earnings  in  the  Boston  Large-Trawler  Fleet.  Economics 
of  Marine  Resources,  University  of  Rhode  Island 
Agricultural  Experiment  Station,  Bulletin  400. 

SHIMADA,  B.,  and  M.  B.  SCHAEFER.  1956.  A  Study 
of  Changes  in  Fishing  Effort,  Abundance,  and  Yield 
for  Yellowfin  and  Skipjack  Tuna  in  the  Eastern  Tropical 
Pacific  Ocean.  Inter-American  Tropical  Tuna  Commission, 
Bulletin  1(7):  348-469. 


TINTNER,    G.    1952.    Econometrics. 
Wiley  and  Sons. 


New    York:    John 


52 


Appendix  Table  1    —  New  England  production  function. 


INDEPENDENT  VARIABLE 

Dependent  variable 

DAYS  ABSENT 

DAYS  FISHING 

CRT 

i 

CREW 

YEAR  BUILT 

H  P 

Reg  Coef. 

t  val 

Reg.  Coef. 

f  val. 

Reg.  Coef. 

(sal 

Reg.  Coef. 

(val. 

Reg  Coef 

(val 

Reg  Coef 

(val. 

f'<    <htr"i    ! 

lot.il  pounds  64 

2677. 

2.47 

6334. 

6.19 

•33074. 

1.42 

10899. 

2.64 

1140 

2  58 

Total  pounds  65 

2215. 

2.02 

6282. 

6.13 

-30615. 

1.31 

7291 

1.75 

1256 

2.79 

Total  pounds  67 

■150.6 

.13 

6198. 

6.87 

-33566. 

2.70 

11076. 

2,95 

607.6 

1  59 

Pooled  total  pounds 

2498. 

4.15 

6395. 

11.41 

■38441. 

3.83 

8561. 

3.73 

896.8 

3.67 

Log  total  pounds  64 

.4914 

8.87 

.9850 

11  59 

-  1912 

1.62 

.2665 

2.18 

.7188 

12.76 

Log  total  pounds  65 

.2530 

4.48 

9640 

12.56 

-.0500 

.47 

.2510 

2.30 

6430 

12.38 

Log  total  pounds  67 

.1032 

1.85 

.9800 

11.84 

.0206 

.18 

.5560 

4.67 

91111 

15.78 

Problem  2 

Total  pounds  64 

5687. 

8.01 

5415. 

5.74 

■42049. 

2.26 

6252. 

1.62 

951  6 

2.31 

Total  pounds  65 

5455. 

7.28 

5375. 

5.62 

^4 1627 

2 

22 

3234. 

.83 

1081. 

2.54 

Total  pounds  67 

5037. 

6.40 

5153. 

5.94 

-45400. 

4 

07 

5828. 

1.63 

372.8 

1.02 

Pooled  total  pounds 

5796. 

13.87 

5365. 

10.18 

-44789. 

5 

24 

4395. 

2.05 

718.9 

3.15 

Log  total  pounds  64 

.8527 

11.08 

.7662 

8.95 

-.1051 

98 

.2256 

1.93 

.4669 

7.18 

Log  total  pounds  65 

.5070 

6.00 

.8320 

10.45 

-.0200 

20 

.2170 

2.02 

.5050 

8.39 

Log  total  pounds  67 

.3263 

4.07 

.8940 

10.55 

-.0330 

32 

.5140 

4.41 

7750 

11.43 

Problem  3 

Total  value  64 

KK9    ' 

24.77 

214  2 

6.32 

1939. 

2.51 

1284 

.94 

103.7 

7.08 

Total  value  65 

884.3 

22.61 

200.5 

5.47 

2082. 

2.50 

41.75 

28 

117  7 

7.31 

Total  value  67 

728.2 

21.30 

204.4 

7.43 

-380.6 

1.00 

8299 

73 

72.24 

6.20 

Pooled  total  value 

889.1 

4  2  ss 

223.2 

11.47 

416.2 

1.19 

26.93 

34 

94  ss 

11  19 

Log  total  value  64 

.9603 

25.20 

.5459 

9.34 

.0497 

6  1 

1779 

2 

12 

.0927 

2.39 

Log  total  value  65 

.7880 

23.95 

.4990 

1118 

1714 

2.75 

.2060 

3 

23 

.0830 

2.75 

Log  total  value  67 

.6848 

22.52 

.5530 

12.22 

.045 

.73 

.2620 

4 

03 

2730 

8.66 

Problem  4 

Total  value  64 

566.1 

19.78 

30.36 

km 

8706. 

11.61 

148.3 

.95 

89.94 

5.40 

Total  value  65 

582.1 

18.24 

11    16 

.27 

8778. 

11.00 

119.1 

.71 

103.6 

5.73 

Total  value  67 

538.5 

21.31 

154.0 

5.52 

1324. 

3.69 

90.76 

79 

77.16 

6.62 

Pooled  total  value 

607.8 

34.90 

100.4 

4.57 

4476. 

12.58 

69.60 

.78 

96.03 

10.03 

Log  total  value  64 

1.349 

23.08 

.2402 

3.69 

.3465 

4.26 

.1633 

1.84 

2122 

4.29 

Log  total  value  65 

1.140 

21.38 

.2290 

4.53 

.4190 

6.69 

.2017 

2.95- 

-.1560 

4.08 

Log  total  value  67 

.9900 

21.81 

.3930 

8.19 

.2200 

3.65 

.2860 

4.34 

.0360 

.94 

Problem  5 

Total  value  64 

896.5 

24.43 

206.7 

5.95 

1759. 

2.21 

133.1 

.97 

101.7 

6.88 

Total  value  65 

8940 

22.44 

190.9 

5.10 

1854. 

2.17 

48.61 

.33 

114.3 

7.01 

Total  value  67 

730.7 

21.13 

201.1 

7.04 

-426.9 

I    MS 

80.81 

.71 

71,4 

6.06 

Pooled  total  value 

896.1 

42.24 

214.3 

10.69 

262.3 

.73 

28.70 

.36 

92.45 

10.78 

Log  total  value  64 

.9690 

25.22 

.5154 

8.38 

.0369 

.45 

.1855 

2.21 

.082 

2.09 

Log  total  value  65 

.7960 

23.92 

.4750 

si 

.1600 

2.57 

2090 

3.29 

.0740 

2.40 

Log  total  value  67 

.6877 

22.36 

.5450 

11  58 

.0390 

.63 

.2600 

4.00 

.2710 

8.54 

Problem  6 

1    .i.il  *  dur  (.4 

563.6 

19.92 

61.4 

1.58 

9046 

12.06 

114  0 

.74 

96.96 

5.84 

Total  value  65 

581.0 

18.35 

36.43 

.88 

9040. 

11.33 

87.10 

.52 

111.2 

6.12 

Total  value  67 

535.6 

21.18 

165.5 

5.74 

1460 

3.96 

94.66 

.83 

79.59 

6.78 

Pooled  total  value 

605.1 

34.94 

122  5 

5.44 

4729. 

13.16 

55.37 

.62 

101.4 

10.64 

Log  total  value  64 

1.352 

23.26 

.2846 

4.24 

.3536 

4.37 

.1469 

1.66 

-.2040 

4.15 

Log  total  value  65 

1.140 

21.61 

.2700 

5.17 

.4240 

6.83 

.1890 

2.79 

-.1460 

3.83 

Log  total  value  67 

.9890 

21.91 

.4200 

8.56 

.2330 

3.87 

.2880 

4.39 

.0390 

1  04 

Problem  7 

Total  value  64 

921  1 

24.25 

198.5 

5.49 

2203. 

2.47 

63.16 

.45 

102.1 

6.86 

Total  value  65 

927.3 

22.57 

187.0 

4.84 

2102. 

2.20 

-21.7 

.14 

116.0 

7.10 

Total  value  67 

865.3 

22.92 

168.0 

6.05 

-279.3 

.72 

-.4334 

.003 

68.96 

6.15 

Pooled  total  value 

951.57 

43.00 

200.8 

9.92 

502.3 

1.33 

-25.40 

.32 

92.60 

1092 

Log  total  value  64 

.9626 

23.87 

.4611 

6.80 

.2064 

2.24 

.1467 

1  'i 

TOO 

2.29 

Log  total  value  65 

8039 

22.93 

.4610 

8.90 

.2760 

3.91 

.1610 

2.53 

09526 

2.79 

Log  total  value  67 

.7550 

21.93 

.4660 

9.35 

.1969 

2.95 

1749 

2.73 

2.557 

7.24 

Problem  8 

Total  value  64 

561.1 

19.55 

63.58 

1.56 

8686.7 

9.68 

132  7 

.85 

100.4 

5.98 

Total  value  65 

578.9 

17  91 

40.95 

.93 

8806.1 

9  12 

7642 

.45 

115.3 

6.24 

Total  value  67 

554.6 

21.05 

162.5 

5.55 

1308. 

3.33 

143.3 

1.23 

79.43 

6.78 

Pooled  total  value 

606.3 

34.49 

130.1 

5.61 

4157. 

10.30 

96.70 

1.07 

103.4 

10.79 

Log  total  value  64 

1.344 

21.93 

.2518 

3.39 

34 1  3 

3.59 

.1953 

2.20 

-.1900 

3.43 

Log  total  value  65 

1.130 

19.79 

.2830 

4.78 

4050 

5.40 

.2000 

2.90 

■  1240 

2.80 

Log  total  value  67 

1.006 

19.51 

.3920 

7.14 

2880 

4.12 

.2780 

4.12 

.0340 

.75 

Problem  9 

Total  pounds  64 

3982. 

3.53 

.6973 

4.44 

-10062. 

.38 

10111 

245 

754.2 

1.71 

Total  pounds  65 

3546. 

3.09 

4981. 

4.62 

-17193. 

.65 

7576. 

1.80 

867.6 

1.90 

Total  pounds  67 

2983. 

2.32 

5251. 

5.56 

-26509. 

2.02 

8165 

2  14 

512.3 

1    14 

Pooled  total  pounds 

4386. 

6.90 

5180.6 

8.92 

-27597. 

2.55 

7471 

3.26 

639  2 

2.63 

Log  total  pounds  64 

.5389 

9.57 

.6974 

7.37 

.2183 

1.70 

.2400 

2.03 

1,1.111,9 

10.85 

Log  total  pounds  65 

.2827 

4.82 

.7771 

8.96 

.3120 

2.64 

.2310 

2.17 

.6550 

11.47 

Log  total  pounds  67 

1420 

2.21 

.8430 

9.06 

.2820 

2  27 

.4879 

4.08 

9330 

14.17 

fr,,blem   f" 

Total  pounds  64 

5955. 

8.47 

3774. 

3.78 

-10501 

.48 

6796. 

1  77 

613.4 

1.49 

Total  pounds  65 

5828. 

7.80 

3893. 

3  84 

-5828. 

.71 

4626 

1.18 

731  9 

1.71 

Total  pounds  67 

6698 

8.50 

3959 

4.52 

-26853. 

2.28 

3868 

1   1  1 

371.3 

1.06 

Pooled  total  pounds 

6402 

15.50 

4062. 

7  45 

■2.16  14. 

2.49 

4586 

2  16 

528  1 

2.34 

Log  total  pounds  64 

8643 

10  94 

5303 

5.53 

24291 

1  98 

.25445 

2  22 

4233 

5  93 

Log  total  pounds  65 

4KI4 

5  52 

.6810 

7  52 

3260 

2.84 

.2320 

2  21 

.5340 

7  90 

Log  total  pounds  67 

3600 

404 

.7540 

7.95 

2370 

1.96 

.4810 

4.13 

7800 

9  86 

:Hor> 


registered  tonnage 


53 


Appendix  Table  1.  —  New  England  production  function.  —  (Continued., 


INDEPENDENT  VARIABLE 

. _ — _ 

Dependent  variable 

CONSTRUCT.3 

MAINE 

4 

GLOUCESTER4 

BOSTON4 

NEW  BEDFORD4 

RHODE  ISLAND4 

Reg.  Coef. 

fval. 

Reg.  Coef. 

rval. 

Reg.  Coef. 

rval. 

^eg.  Coef. 

rval. 

*eg.  Coef. 

rval. 

Reg.  Coef. 

rval. 

Y  INT 

R* 

R1 

F 

Problem  1 

Total  pounds  64 

-90900. 

.368 

362 

44.37 

Total  pounds  65 

74900. 

.387 

381 

46.096 

Total  pounds  67 

11300. 

.285 

277 

29.541 

Pooled  total  pounds 

38300. 

.339 

337 

117.5 

Log  total  pounds  64 

1.83 

.650 

646 

141.2 

Log  total  pounds  65 

2.35 

.592 

588 

105.85 

Log  total  pounds  67 

1.52 

.568 

563 

101.06 

Problem  2 

Total  pounds  64 

-355000 

.451 

445 

62.48 

Total  pounds  65 

-222000. 

.460 

453 

61.915 

Total  pounds  67 

-209000 

.354 

347 

42122 

Pooled  total  pounds 

-261000 

.426- 

424 

169.7 

Log  total  pounds  64 

1.70 

-681 

677 

162.0 

Log  total  pounds  65 

2.22 

.608 

604 

113.29 

Log  total  pounds  67 

1.49 

.578 

528 

107.10 

Problem  3 

Total  value  64 

48000 

.877 

876 

543.1 

Total  value  65 

46200 

.871 

870 

494.17 

Total  value  67 

-18600 

.814 

813 

338.33 

Pooled  total  value 

-35000 

.846 

845 

1260 

Log  total  value  64 

1.56 

.845 

843 

415.0 

Log  total  value  65 

1.86 

.876 

878 

516.19 

Log  total  value  67 

1.63 

.845 

843 

419.96 

Problem  4 

Total  value  64 

-82900 

.842 

840 

404.6 

Total  value  65 

-85500 

.838 

846 

378.53 

Total  value  67 

48100 

.814 

812 

336.77 

Pooled  total  value 

-70700 

.808 

807 

960.3 

Log  total  value  64 

1.28 

.828 

826 

365.3 

Log  total  value  65 

1.54 

.858 

857 

442.78 

Log  total  value  67 

1.26 

.839 

838 

402.3 

Problem  5 

Total  value  64 

-5419 

.96 

41800 

.878 

876 

452.6 

Total  value  65 

-7455 

1.24 

-37700 

.872 

870 

412.89 

Total  value  67 

-2020 

.44 

-16200 

.815 

812 

281.3 

Pooled  total  value 

■5982 

1.81 

-28000 

.847 

846 

1053. 

Log  total  value  64 

-.0581 

1.58 

1.67 

.846 

844 

347.6 

Log  total  value  65 

-.0420 

1.56 

1.94 

.877 

875 

432.17 

Log  total  value  67 

-.0180 

.66 

1.67 

.845 

843 

349.59 

Problem  6 

Total  value  64 

19489 

3.16 

-105000 

.846 

844 

346.8 

Total  value  65 

17085 

2.61 

-104000 

.841 

839 

321.68 

Total  value  67 

7184 

1  -f, 

-56600 

.815 

813 

282.14 

Pooled  total  value 

14256 

3.95 

-87100 

.810 

809 

813.1 

Log  total  value  64 

.0923 

2.40 

1.12 

.830 

828 

309.3 

Log  total  value  65 

.0780 

2.74 

1.40 

.861 

859 

376.95 

Log  total  value  67 

.0670 

2.35 

1.14 

.842 

840 

340.43 

Problem  7 

Total  value  64 

-690.9 

.12 

11319. 

1.55 

-808.7 

.12 

4635 

.62 

1  1  74 

.17 

11044 

1.52 

51000 

.881 

878 

251.3 

Total  value  65 

-1323 

.21 

12207 

1.55 

-853.1 

1  1 

8115 

1.00 

1470 

20 

14986 

1.91 

49700. 

.876 

873 

232.05 

Total  value  67 

-382.1 

.09 

3708 

.61 

-5542 

.94 

-7683 

1.23 

19939 

3 

42 

7313.1 

1.19 

-13300. 

.839 

835 

178.81 

Pooled  total  value 

-1253 

.45 

8604 

1.99 

-2798 

.67 

2738 

.62 

7520 

1 

86 

11149 

2.57 

-34700. 

.854 

852 

605.9 

Log  total  value  64 

-.0529 

1.36 

.1055 

1.95 

.0044 

.02 

.0535 

.97 

0251 

50 

.0363 

.67 

1.67 

.853 

849 

197.7 

Log  total  value  65 

-.0310 

1  09 

.0790 

1.96 

-.0007 

.002 

-.0280 

.68 

0320 

87 

.0820 

2.02 

1.88 

.884 

881 

248.8 

Log  total  value  67 

-.0164 

.57 

.0110 

.25 

-.0960 

.67 

.1620 

3.54 

.1060 

2 

54 

.0320 

.74 

1.827 

.862 

858 

214.64 

Problem  8 

Total  value  64 

18171 

2.79 

-7228 

K8 

-7694 

.98 

6246 

75 

2402 

.32 

-14404 

1.77 

-96600. 

.848 

844 

190.2 

Total  value  65 

18068 

2.59 

-5969 

.68 

-8943 

1.04 

4459 

.47 

.3492 

.42 

-9282 

1.05 

-99100. 

.842 

838 

174.48 

Total  value  67 

7666. 

1.64 

-14675 

2.31 

-8656 

L39 

7234 

1.10 

16873 

2.77 

-9428 

1.45 

48900. 

.822 

818 

159.45 

Pooled  total  value 

14241 

3.78 

-12459 

2.56 

-8300 

1.75 

2039 

41 

8035 

1.75 

-12687 

2.59 

-78900. 

.812 

811 

449.3 

Log  total  value  64 

.0581 

1.43 

-.0034 

.06 

.0268 

.49 

.0283 

.49 

0046 

.09 

-.1293 

2.30 

1.15 

.838 

834 

175.9 

Log  total  value  65 

.0600 

1.93 

-.0160 

.39 

.0037 

.09 

.0200 

.45 

0120 

.30 

-.0510 

1.17 

1  40 

.863 

859 

206.2 

Log  total  value  67 

.0634 

2.08 

-.0260 

.57 

-.0370 

.81 

.0740 

1.53 

.0560 

1.28 

-.0220 

.49 

1.19 

.843 

839 

185.77 

Problem  9 

Total  pounds  64 

-339024 

1.96 

562668 

2.60 

351898 

1.69 

59089 

.27 

29683 

.15 

358050 

1.66 

51500. 

.408 

392 

23.4 

Total  pounds  65 

-380585 

2.19 

479210 

2.18 

332023 

1.59 

103437 

.45 

51156 

.30 

355627 

1.62 

189000- 

.418 

401 

23.393 

Total  pounds  67 

-102455 

.67 

85775 

.42 

-124072 

.62 

354221 

1.67 

409081 

2.06 

215326 

1.03 

257000. 

.330 

313 

16.979 

Pooled  total  pounds 

-272358 

2  8  1 

386757 

3.12 

181225 

1.51 

52889 

42 

135014 

1.16 

321650 

2.58 

196000. 

.378 

372 

62.83 

Log  total  pounds  64 

-.0678 

1.25 

.2383 

3.16 

.0319 

44 

.1261 

1.63 

.0587 

.84 

.0658 

.87 

2.15 

.694 

686 

77.14 

Log  total  pounds  65 

-.0280 

.60 

.1960 

291 

.0079 

.12 

.1300 

1.86 

.0190 

.31 

.0190 

.28 

2.41 

.638 

628 

57.558 

Log  total  pounds  67 

-.0560 

of, 

-.0330 

.42 

-.1400 

1  75 

.3240 

3.80 

.1670 

2.13 

-.1470 

1.81 

1   Sll 

.598 

587 

51.203 

Problem  10 

Total  pounds  64 

-265530. 

1.66 

475561. 

2.37 

319786. 

1.65 

(7985. 

.23 

45782. 

.25 

205331. 

1.03 

-312000. 

.487 

473 

32.25 

Total  pounds  65 

-298694 

1  85 

429528. 

2  10 

302920. 

1.53 

S2596. 

.39 

20693. 

1  1 

230513. 

1.13 

-187000. 

489 

475 

31.19 

Total  pounds  67 

-106493. 

.76 

-52962. 

28 

-314526 

1.69 

532042. 

2  71 

655419. 

3.60 

50766. 

.26 

105000. 

430 

415 

25.934 

Pooled  total  pounds 

-209963. 

2  3  7 

278594. 

2.43 

103667. 

.93 

118571. 

101 

250792. 

2.33 

159885. 

1.38 

-130000 

965 

460 

89.87 

Log  total  pounds  64 

-.0017 

.03 

1834 

2.53 

.0584 

.82 

0926 

1.23 

.0722 

1.07 

-.01951 

.27 

1.88 

711 

704 

83.81 

Log  total  pounds  65 

.0050 

.12 

.1670 

2  52 

.0170 

.27 

.1160 

1.66 

.0290 

.47 

-.0240 

.37 

2.22 

645 

635 

59.29 

Log  total  pounds  67 

-.0330 

.64 

-.0390 

.50 

-.1300 

1  65 

.3040 

3.62 

.1910 

2.50 

-.1570 

1.96 

1.72 

610 

599 

53.770 

3£quak  1  if  wooden  vessel.  0  otherwise. 
4  Equals  1  if  vessel's  homeport  Puerto  Rk 


54 


Appendix  Table  2.  —Alternative  specifications  of  production  functions  for  vessels  in  the  eastern  tropical  Pacific  tuna  fishery,  1966,  1967,  1968. 


Dependent  variable 


INDEPENDENT  VARIABLE 


CAPACITY    A^T    HORSEPOWER     CRT'     CREW    ™f*J2°     Jgg    gfjfj     DUM  66<    DUM  67«      VINT  F 


Problem  1 


Total  value 

Linear  - 

Reg.  Coef.s 

.407 

.594 

.124 

f  ratio 

7.590 

.285 

4.820 

Log 

Reg.  Coef.5 

.365 

.310 

.368 

t  ratio 

5.140 

3.320 

4.660 

Problem  2 

Total  pound 

i 

Linear  - 

Reg.  Coef.5 

3.840 

5.620 

.751 

f  ratio 

1.030 

3.890 

4.180 

Log 

Reg.  Coef.5 

.438 

.373 

.339 

t  ratio 

7.390 

4.790 

5.150 

Problem  3 

Weighted  total  pounds 


Linear  -    Reg.  Coef.5 

4.810 

6.740 

.605 

f ratio 

11.800 

4.270 

3.080 

Log       -    Reg.  Coef.5 

.520 

.416 

.328 

t  ratio 

8.410 

5.120 

4.770 

Problem  4 

Total  pounds 

Linear  -   Reg.  Coef.5 

f ratio 
Log       -    Reg.  Coef.5 

t  ratio 


3.930 

5.780 

179.000 

10.700 

3.960 

2.880 

.410 

.061 

.242 

1.67 

4.870 

5.860 

3.330 

2.87 

37.300  -  6.920    36.890  96.28  .643 

1.350  2.440 

.067  .044  -    .196   76.17  .587 

2.080  2.210 


582.000  202.000  -  660.400  121.20  .694 

2.750  1.030 

.024  .049      .453  113.80  .680 

.914  2.940 


255.000  732.000  -  116.300  132.00  .712 

1.180  3.500 

.026  .065      .168  127.00  .704 

.946  3.710 


414.000 

209.000 

540.000 

1.980 

1.060 

2.810 

.010 

•       .037 

.051 

.304 

1.290 

2.870 

-1735.000  102.00  .697 


.559  70.74  .649 


Problem  5 
Weighted  total  pounds 
Linear  -    Reg.  Coef.5 

t  ratio 
Log      -    Reg.  Coef.s 
t  ratio 


4.440 

6.300 

8.030 

3.940 

.448 

5.030 

.217  1.230  -283.00  370.00        13.30    -143.000  -668.000  -479.000     1856.000     69.68  .648 

.829  1.750  1.54  1.97          1.22           .458  3.290  3.890 

.065  .317  1.56  -        .010  -        .039  6.580    -          .762      75.87    .653 

.586  4.111  2.53  .287  1.300  3.530 


'Gross  registered  tonnage. 

2 Equals  one  if  vessel's  home  port  is  Puerto  Rico,  zero  otherwise. 

3Equals  one  if  vessel  was  built  after  1962,  zero  otherwise. 

4 Dummy  variables  for  year  of  observation. 

5 Regression  coefficient. 


Appendix  Table  3.  -  New  England  trawl  fleet:  average  vessel  data  by  tonnage  class,  1964,  1965,  1967. 


Number  of 
observations 

GRT 

Days 
absent 

Days 
fishing 

Trips 

Horsepower 

Year 
built 

Crew 

Construction 
(percent  wood) 

Measures  of 

output 

Tonnage  class 

Thousands  of 
pounds 

Total 

value 
($1000) 

0-50 

492 

30 

118 

4S 

87 

163 

42 

3.6 

98 

808 

37 

51-100 

354 

7(1 

149 

89 

36 

253 

43 

5.9 

93 

1086 

83 

101-150 

147 

120 

162 

104 

24 

349 

44 

7.9 

88 

1225 

118 

151-200 

57 

170 

168 

96 

20 

479 

44 

8.6 

24 

1142 

114 

201-250 

33 

229 

235 

155 

24 

604 

45 

14.4 

0 

2672 

242 

251-300 

15 

271 

224 

152 

2  3 

630 

38 

13.7 

0 

2591 

253 

301-400 

15 

313 

235 

141 

17 

623 

36 

9.0 

0 

4942 

191 

400 

6 

495 

221 

126 

24 

503 

44 

12.7 

0 

3439 

260 

55 


Appendix  Table  4.  -  Tropical  tuna  seine  fleet:  average  vessel  data  by  tonnage  class,  1966,  1967,  1968. 


Number  of 
observations 

Capacity 

Days 

at  sea 

GRT 

Horsepower 

Year 
built 

Measures  of  ot 

tput 

Capacity  class 

Total 

value 

($1000) 

Total  in 

thousands  of 

pounds 

Weighted 

total  in 

thousands  of 

pounds 

100-199 

47 

173 

152 

210 

508 

46 

236 

1504 

1388 

200-299 

83 

251 

168 

370 

731 

48 

292 

2542 

2401 

300-399 

62 

346 

172 

421 

908 

51 

360 

2550 

2461 

400-499 

24 

453 

182 

482 

1100 

50 

389 

2765 

2766 

500-599 

19 

537 

162 

619 

1281 

56 

523 

3749 

3719 

600-699 

5 

650 

133 

673 

1649 

59 

448 

3166 

3319 

700-799 

4 

793 

180 

856 

1589 

63 

817 

6447 

6946 

800-899 

6 

811 

191 

804 

1600 

64 

781 

6016 

6479 

900-999 

12 

924 

161 

793 

1850 

53 

637 

5092 

5492 

K 

5 

1067 

171 

855 

1600 

43 

687 

5751 

6454 

56 


Optimal  Fishing  Effort  in  the  Peruvian  Anchoveta  Fishery 


Edilberto  L.  Segura1 


ABSTRACT 

This  paper  introduces  a  new  approach  to  measuring  technical  change,  increased 
skills  of  the  skipper  and  the  fishermen,  water  temperature,  etc.,  to  obtain  a  better  measure 
of  fishing  effort  and  therefore  a  revised  estimate  of  the  optimum  quantity  to  be  landed. 
The  revised  technique  used  adjusts  the  level  of  landings  to  an  index  rather  than  the 
level  of  fishing  effort,  indicating  the  level  of  landings  that  would  have  resulted  in 
previous  periods  if  the  current  landings/effort  relationship  is  used. 

The  revised  yield/effort  relationship  which  results  yields  16.2  million  ton-trips  as 
the  optimal  fishing  effort,  as  opposed  to  the  23  million  ton-trips  which  were  obtained 
without  this  measure  of  technical  change. 


INTRODUCTION 

During  the  last  decade  the  Peruvian  fishing 
industry  has  become  one  of  the  most  important 
elements  of  the  Peruvian  economy.  In  1969 
Peruvian  exports  offish  meal  and  fish  oil  reached 
U.  S.  $195  million,  or  30%  of  total  Peruvian 
foreign  exchange  earnings  during  that  year. 
Almost  all  fish  meal  and  fish  oil  production  has 
utilized  "anchoveta"  {Engraulis  ringens)  as  raw 
material.  Total  landings  of  anchoveta  have 
increased  from  1.9  million  metric  tons  in  1959  to 
8.9  million  metric  tons  in  1969.  This  increase  in 
landings  represents  an  average  annual  rate  of 
growth  of  18% . 

In  recent  years,  due  to  the  rapid  expansion 
of  the  industry,  its  importance  to  the  Peruvian 
economy,  and  the  size  of  the  landings,  several 
studies  have  been  made  to  determine  the  maxi- 
mum sustainable  yield  of  the  Peruvian  fish 
stock  and  the  optimal  level  of  fishing  effort 
(Boerema  et  al.,  1965;  Schaefer,  1967,  1970; 
Gulland,  1968).  Although  these  studies  contain 
extensive  discussions  of  fishing  effort,  there 
remain  some  doubts  about  the  adequacy  of  the 
measures  used  to  evaluate  fishing  effort.  As  a 
result,  the  estimation  of  the  optimal  level  of 
fishing  effort  has  been  biased.  In  this  paper  I 
attempt  to  estimate  the  optimal  level  of  fishing 
effort  taking  into  consideration  the  effect  of 
input  variables  not  previously  included,  such  as 


technological  change,  increased  skills  of  skippers 
and  fishermen,  water  temperature,  etc. 

CONCEPTUAL  ISSUES 

In  a  bioeconomic  model,  fishing  effort  is  an 
index  or  proxy  for  several  inputs  that  partici- 
pate in  the  fishery,  including  capital,  labor, 
management,  technological  change,  and  other 
variables.  Although  the  fishing  effort  index 
might  vary  for  different  fisheries,  it  can  be 
generalized  as  being  the  product  of  fishing  time 
(number  of  days  in  grounds,  number  of  trips 
made,  number  of  hours  fished,  etc.)  multiplied 
by  some  measure  of  fishing  power  (gross  tonnage, 
length,  engine  horsepower,  etc.).  This  measure 
should  be  a  proxy  for  capital  and  labor.  The 
resulting  measure  of  fishing  effort  should  be 
corrected  by  such  factors  as  technological 
change  (introduction  of  power  block,  echo 
sounder,  steel  vessels,  etc.),  changes  in  manageri- 
al and  fisherman  skill,  and  other  variables  that 
represent  changes  in  fishing  power.  To  determine 
the  optimal  fishing  effort  in  the  fishery  and  the 
maximum  sustainable  yield,  most  of  the  studies 
of  the  Peruvian  stock  have  utilized  the  following 
Schaefer  production  function: 


1   Ph.D.  Candidate,  Columbia  University  and  Economist 
of  Bailey,  Tondu,  Warwick  and  Company,  Ltd.,  New  York. 


(1)  CIE   =   a 

-  bE  or,  C  =  aE-  bE2 

Where: 

C 

=    Total  landings  of  anchoveta 

E 

=    Fishing  effort 

a,b 

=    Parameters 

57 


Schaefer  (1967)  used  as  a  measure  of  fishing 
effort  the  average  number  of  boats  during  the 
period,  adjusted  for  changes  in  the  size  composi- 
tion of  the  fleet.  In  1970,  Schaefer,  recognizing 
that  this  measure  could  generate  some  bias, 
utilized  as  a  measure  of  fishing  effort  the  number 
of  trips  made  by  the  fleet,  times  the  average 
vessel  capacity  (ton-trips).  This  unit  was  also 
utilized  by  Boerema  et  al.  (1965).  However,  all 
these  studies  ignored  the  effect  of  technological 
change  and  increased  fisherman  skills  on  the 
level  of  landings.  This  neglect  arose  from  the 
difficulty  in  quantifying  these  variables. 
Although  for  several  advanced  fisheries  these 
two  variables  can  indeed  be  ignored,  such 
neglect  is  questionable  in  the  Peruvian  anchoveta 
fishery.  The  importance  of  such  variables  was 
recognized  by  Gulland  (1968). 

The  size  of  the  Peruvian  fleet  has  increased 
from  462  units  in  1959  to  1,064  in  1962  and  to 
1,836  in  1964.  After  1964,  fleet  size  began  to 
decline,  reaching  1,308  units  in  1969.  From 
these  figures  it  is  clear  that  up  to  1964  a  large 
percentage  of  the  skippers  and  fishermen  were 
fishing  for  the  first  time.  However,  after  1964, 
with  the  reduction  of  the  fleet  size,  only  the 
most  efficient  skippers  remained  in  the  fishery. 
This  situation  and  the  experience  gained  by  the 
fishermen  after  several  years  of  operations,  have 
served  to  increase  the  average  skill  of  the 
fisherman. 

In  addition  to  increased  labor  skills,  during 
the  last  decade  several  technological  innovations, 
such  as  power  block,  echo  sounder,  steel  vessels, 
and  pumps  for  transferring  the  fish  from  the 
net  to  the  hold,  have  been  gradually  introduced 
into  the  fleet.  In  1969,  92%  of  the  fleet  had  at 
least  three  of  these  items  of  gear,  as  opposed 
to  79%  two  years  before.  If  a  measure  of  fishing 
effort  omits  the  effect  of  increased  labor  efficiency 
and  technological  innovations,  then  the  most 
recent  estimates  of  fishing  effort  will  be  biased. 
The  estimation  of  the  optimal  fishing  effort 
will  also  be  biased. 

The  type  of  bias  that  will  be  introduced  by 
omitting  the  increased  efficiency  of  the  fleet  can 
be  deduced  from  Figure  1. 

In  Figure  1,  if  the  efficiency  of  the  fleet 
increased  during  periods  1  to  3,  the  observed 
data  for  catch  and  effort  will  produce  curve  A. 
However,  the  relationship  of  catch  to  effort  in 
terms  of  efficiency  in  year  base  "0"  is  given  by 


curve  B.  The  effect  of  ignoring  increased  ef- 
ficiency would  be  to  underestimate  the  most 
recent  measures  of  fishing  effort.  If  the  observa- 
tions of  fishing  effort,  unadjusted  by  efficiency, 
are  consistent  from  year  to  year,  they  still  will 
give  a  correct  measure  of  the  maximum  sus- 
tainable yield,  as  it  is  shown  in  Figure  1. 
However,  the  determination  of  the  optimal  level 
of  fishing  effort,  in  terms  of  some  constant  level 
of  efficiency,  will  be  biased.  Usually,  one  is 
interested  in  obtaining  the  optimal  level  of 
fishing  effort  in  terms  of  efficiency  during  the 
current  period.  This  relationship  is  given  in 
Figure  2,  where  "period  3"  is  the  current  period. 

Since  vessels  are  more  efficient  during  period 
3,  to  obtain  the  maximum  sustainable  yield 
C-2,  the  industry  will  require  a  smaller  effort 
in  terms  of  number  of  ton-trips  than  the  effort 
used  in  period  2.  In  fact,  instead  of  requiring 
an  effort  E2,  the  industry  will  require  only  an 
effort  E'2,  considering  the  higher  efficiency  of 
vessels  in  period  3.  It  is  obvious  that  to  obtain 
an  unbiased  optimal  level  of  fishing  effort  at 
current  efficiency,  it  will  be  necessary  to  adjust 
the  index  for  fishing  effort  to  reflect  technological 
change,  changes  in  fishermen  skills,  and  other 
variables. 

Although  the  construction  of  an  index  for 
fishing  effort  that  includes  technological  change 
and  other  such  variables  is  the  ideal  method  to 
determine  an  unbiased  level  of  optimal  fishing 
effort,  usually  it  is  not  easy  to  construct  such 
an  index.  This  is  because  several  of  the  above- 
mentioned  variables  are  difficult  to  quantify. 
When  this  is  the  case,  an  alternative  approach 
has  to  be  devised. 

The  alternative  approach  that  is  used  in  this 
paper  is  to  adjust  the  level  of  landings  obtained, 
rather  than  the  level  of  fishing  effort,  for  changes 
in  efficiency.  That  is,  given  the  observed  un- 
adjusted fishing  efforts  and  the  landings  in 
several  periods,  the  problem  is  to  obtain  a  catch- 
to-effort  relationship  that  will  show  that  level 
of  landings  that  would  have  been  obtained  in 
the  several  periods  if  vessels  of  efficiency  of  the 
current  period  would  have  been  used.  This 
adjusted  curve  and  the  actual  observed  curve 
are  shown  in  Figure  3.  The  optimal  level  of 
fishing  effort  in  terms  of  vessels  of  current 
efficiency  (E*2  in  the  figure)  will  be  obtained  by 
maximizing  catch  in  curve  A. 

The  difference  between  this  approach  and  the 


58 


MSY 


"~     Curve   Bi   Catch-effort 
relationship   in   terms   of 
efficiency  of  period  0. 


Curve  Aj 
Catch-effort 
relationship 
as   observed 


Figure  1. —  Biased  estimate  of  fishing  effort  due  to  an  underestimate  of  increased  efficiency. 


Catch 


"2  2 

Level  of  effort  required  Actual  level  of 

to  obtain  the  catch  of  period  2  effort  observed 

using  vessels  of  effic.  3.  in  period  2. 


Effort 
(ton-trips) 


Figure  2.  —  Optimal  fishing  effort  based  on  current  efficiency  levels. 

59 


MSY 


Actual  level 
of  landings 
observed  in 
period  2 


Level  of 
landings 
that  would 
have  been 
obtained  in 
period  2,  if 
vessels  had 
efficiency 
of  period  4. 


Effort 


(ton-trips) 


Figure  3.  —  Actual  and  adjusted  catch  effort  curves. 


first  one  is  that  in  the  first  approach  the  fishing 
effort  is  adjusted  for  efficiency  changes  and 
catch  remained  at  the  observed  levels;  in  the 
second  approach  the  catch  is  adjusted  for  ef- 
ficiency changes  and  the  fishing  effort  remains 
at  the  level  observed.  The  second  approach  has 
the  advantage  that  it  can  be  more  easily  handled 
with  statistical  techniques.  It  should  be  noted 
that  curve  A  in  Figure  3  gives  the  level  of 
landings  that  would  have  been  obtained  in  past 
periods  if  vessels  at  that  time  had  the  efficiency 
of  the  current  period.  Actually,  this  curve  has 
not  been  observed;  and  the  maximum  of  the 
curve,  although  it  indicates  the  optimal  level 
of  fishing  effort  in  terms  of  current  efficiency, 
will  not  give  the  maximum  sustainable  yield 
of  the  stock.  The  maximum  sustainable  yield 
will  in  fact  be  given  by  curve  B,  as  it  was 
shown  in  Figure  1. 

Since  curve  A  in  Figure  3  is  actually  the 
relationship  of  effort  to  catch  keeping  all  other 
variables  (including  efficiency)  constant,  the 
multiple  regression  technique  can  be  applied. 
In  fact,  the  statistical  meaning  of  a  partial 
regression  coefficient  is  that  it  measures  the 
effect  of  the  independent  variable  on  the  de- 
pendent one,  keeping  all  other  variables  constant. 


The  use  of  the  regression  analysis  to  obtain  the 
optimal  fishing  effort  is  presented  below. 

The  logistic  model  as  presented  by  Schaefer 
(1957)  and  reproduced  in  equation  (1)  is  a 
stochastic  rather  than  an  exact  relationship: 

(2)  C  =  aE  -  bE2  +  e 
Where  "e"  is  an  error  term. 

In  this  model,  if  the  measure  of  effort  used 
were  a  proxy  for  all  the  several  inputs  utilized 
when  fishing  and  affecting  catch,  then  the  error 
term  "e"  should  be  randomly  distributed.  That 
is,  no  other  input  variable,  when  added  to 
equation  (2),  should  be  statistically  significant 
in  explaining  changes  in  the  level  of  catch.  In 
fact,  if  no  variables  have  been  omitted  in 
equation  (2)  (all  of  these  are  represented  in  the 
proxy  fishing  effort),  then  no  sign  of  auto- 
correlation of  the  error  term  should  exist.  If 
this  is  the  case,  one  could  conclude  that  the 
measure  of  fishing  effort  used  is  adequate  and 
that  it  can  be  reliably  used  to  estimate  both  the 
maximum  sustainable  yield  and  the  optimum 
level  of  effort. 

We  can  further  test  if  the  measure  of  fishing 
effort  is  adequate  by  introducing  into  equation 
(2)  input  variables  such  as  technological  change 


(JO 


and   crew   size.    If  we   did   this   the   following 
equation  would  result: 

(3)   C  =  axE  -  bxE2  +  cL  +  dT  +  e 
Where: 

C  —  Total  landings 
E  =  Fishing  effort 
L   =  Labor  employed  or  crew  size 
T  =  Technological  change  expressed 
as  T  =  1  in  period  1,  T  =  2  in 
period  2,  T  =  3  in  period  3,  etc. 

If  the  coefficients  of  "L"  and  "T"  are  statisti- 
cally significant  (as  given  by  their  t-values), 
it  means  that  the  measure  of  fishing  effort  used, 
"E,"  did  not  adequately  include  the  effect  of 
these  variables  on  catch.  Consequently,  the  use 
of  equation  (2)  alone  would  produce  biased 
estimators  of  the  coefficients  "a"  and  "6"  of 
"E"  and  "E2,"  respectively.  In  this  case  we 
can  either  correct  the  measure  of  fishing  effort 
used  (which  is  the  first  procedure  indicated  in 
Figure  1)  or  we  can  isolate  the  effect  of  other 
variables  on  catch  using  a  multiple  regression 
equation  that  would  include  these  variables 
(which  is  equivalent  to  the  second  approach 
indicated  in  Figure  3). 

If  the  second  approach  is  used,  technological 
innovation  and  crew  size  must  be  kept  at  a  fixed 
level  in  equation  (3).  Usually  this  would  be  at 
the  current  levels.  After  this  is  done  we  can 
obtain  the  true  value  of  fishing  effort  by  maxi- 
mizing catch  in  equation  (3).  Keeping  the  effect 
of  "T"  and  "L"  on  catch  at  some  constant  level 
K,  equation  (3)  would  become 


C 


axE   -    bxE2    +    K 


or 

(4)     (C   -    A')    =   aiE   -    bxE2 

Which  is  the  model  as  developed  by  Schaefer 
(1957)  after  the  effect  of  technological  change 
and  crew  size  is  removed.  The  optimal  level 
of  fishing  effort,  at  constant  vessel  capacity 
and  crew  size,  that  will  maximize  catch  is 
given  by  equating  zero  to  the  first  derivative 
of  equation  (4)  as  follows: 


d(C-K) 
de 


a,  -2  6,  E  =  0 


or 


(5)  Optimal  fishing  effort  =  E*  = 


ax 


2  6, 


STATISTICAL  RESULTS 

Using  the  data  presented  in  Table  1,  several 
regressions  were  made  to  test  for  the  adequacy 
of  the  measures  of  fishing  effort  available  to  us. 
In  Table  1,  total  landings  is  defined  as  the  catch 
by  the  fishermen  in  thousands  of  pounds.  The 
unit  used  for  fishing  effort  is  the  number  of 
trips  made  times  the  average  vessel  capacity. 
Data  on  fishing  effort  was  compiled  by  the 
Instituto  del  Mar  del  Peru,  and  it  is  supposed 
to  be  adjusted  for  the  effect  of  closed  seasons, 
strikes,  and  for  some  changes  in  gear  efficiency. 
Other  variables  included  in  the  analysis  are  the 
number  of  fishermen  employed  in  the  industry, 
the  size  of  the  bird  population  (which  is  supposed 
to  be  an  important  element  in  fishing  mortality), 
and  veda  (closed)  seasons. 

As  has  been  recognized  by  Gulland  (1968) 
and  by  Schaefer  (1967),  because  of  the  rapid 
growth  of  the  Peruvian  fishery,  it  has  not 
remained  in  steady  state  equilibrium  in  every 
year.  Under  these  circumstances,  the  use  of  a 
relationship  of  catch  to  effort  will  produce  too 
high  an  estimate  of  steady  state  abundance 
and  catch  for  a  given  fishing  effort.  One  way  to 
correct  this  situation  is  to  use  the  "Gulland 
Method"  (Gulland,  1961)  in  which  the  total 
landings  are  related  to  the  average  effort  exist- 
ing during  the  life  span  of  a  fish  in  the  fishery, 
which  is  approximately  two  years.  This  method 
has  been  used  in  this  paper. 

Schaefer  (1970)  used  the  same  data  presented 
in  Table  1  to  estimate  the  maximum  sustainable 
yield  of  the  stock  and  the  optimum  level  of  fishing 
effort.  I  have  added  observations  for  the  year 
1968-1969.  The  regression  equivalent  to  the 
one  used  by  Schaefer  in  1970  is  as  follows: 

(6)     C   =   0.7769  E   -    0.1706  E2 
(8.6)  (-3.8) 

Coefficient  of  Determination  {R2)  =  0.84 
Durbin-Watson  Statistic  (D-W)  =  0.7 
Standard  Error  of  Estimate  (SEE)  =  813 
Figures  in  parentheses  are  £ -values. 

Equation  (6)  is  useful  for  finding  the  maximum 
sustainable  yield  of  the  fishery.  The  estimated 
MSY  is  given  at  8.8  million  metric  tons.  This 
value  is  very  close  to  the  value  of  8.5  million 
metric  tons  obtained  by  Schaefer  (1970).  By 
observing  the  data  of  total  landings  in  Table  1 
we  cannot  appreciate  the  danger  of  overfishing 


61 


Table    1.    —    Catch    and    Effort   Data   for   the   Peruvian 
Anchoveta  Fishery,  1960-1969. 


Catch  by 

Fishing 

Adult 

fishermen 

effort 

Number  of 

bird 

Catch  per 

Fishing 

103  metric- 

103  ton- 

fishermen 

population 

unit  of 

year 

tons 

trips 

employed 

103 

effort 

(1) 

(2) 

(3) 

(4) 

(5) 

1960-61 

3,934 

6,367 

8,800 

12,000 

0.551 

1961-62 

5,502 

8,131 

11,750 

17,000 

.603 

1962-63 

6,907 

11,788 

19,100 

18,000 

.478 

1963-64 

8,006 

17,866 

20,100 

15,000 

.376 

1964-65 

8,037 

21,329 

18,900 

17,300 

.376 

1965-66 

8,096 

22,058 

19,000 

4,300 

.356 

1966-67 

8,242 

20,845 

17,800 

4,800 

.435 

1967-68 

9,818 

19,874 

17,500 

4,500 

.472 

1968-69 

10,088 

22,350 

19,600 

5,000 

.421 

Source:  (1),  (2),  (4):   Years  1960-1968,  from  Schaefer  (1970) 

Year    1968-1969,  from  Instituto  del  Mar  del 
Peru,  Resumen  General  dela  Pesqueria, 
Lima,  1970. 
(3):    From    Sociedad    Nacional   de    Pesqueria,   unpublished 
materials. 


in  the  Peruvian  stock,  since  landings  have 
increased  throughout  the  period.  However,  by 
analyzing  data  for  calendar  years  up  to  1969 
a  different  picture  of  the  situation  is  observed. 
During  recent  years  annual  landings  have  been 
as  follows: 


Year 
1961 
1962 
1963 
1964 
1965 
1966 
1967 
1968 
1969 


Million  Metric  Tons 
4.58 
6.28 
6.42 
8.80 
7.23 
8.53 
9.82 
10.44 
8.95 


It  is  clear  from  these  data  that  landings  will 
not  continue  to  increase  at  the  rates  experienced 
in  the  past,  and  that  we  can  only  expect  to 
see  fluctuations  in  landings  around  the  MSY, 
if  fishing  effort  is  kept  under  control. 

The  result  given  by  equation  (6)  as  to  the 
optimal  level  of  fishing  effort  is  less  than 
satisfactory.  The  value  given  by  this  equation, 
and  which  is  close  to  that  obtained  by  Schaefer 
(1970),  is  23  million  ton-trips.  Observing  the 
data  in  Table  1  we  can  see  that  this  value  of 
fishing  effort  has  not  been  obtained  up  to  now. 
This  result  is  very  unrealistic  since  it  says  that 
the  Peruvian  fishery  has  actually  surpassed 
the  MSY  but  has  not  yet  reached  the  optimum 


level  of  fishing  effort.  However,  from  the 
discussion  in  the  first  part  of  this  paper,  it 
seems  that  the  measure  of  fishing  effort  used 
is  inadequate. 

In  equation  (6)  we  can  see  that  the  value  of 
0.7  for  the  Durbin-Watson  statistics  indicated 
that  there  is  a  strong  autocorrelation  of  the 
error  term.  This  level  of  autocorrelation  is  an 
indication  that  important  variables  have  been 
omitted  from  the  equation.  Using  the  procedure 
indicated  above,  several  input  variables  will  be 
introduced  in  equation  (6),  in  order  to  determine 
their  significance  and  the  bias  of  the  estimation 
of  fishing  effort.  Some  of  the  regressions  that 
were  run  are  the  following: 


(7)  C     =    0.7022  E    -   0.2167  E2      R2  =  0.97 

(15.7)  (-9.5)    D-W  -  1.8 

+     541.0  T     SEE  =  382 

(5.2) 

(8)  C    =    0.5225  E    -   0.1722  E2      R2  =  0.98 

(2.7)  (-3.4)    D-W  =  2.2 

+      561. IT      +    0.0884  L    SEE  =  384 

(5.1)  (1.0) 

(9)  C    =     0.499  E     -  0.1556  E2       R2  =  0.98 

(4.2)  (-4.0)    D-W  =  2.8 
+      733.7  T     +     903.0  V     SEE  =  325 

(5.2)  (1.8) 


62 


(10)  C    =    0.4977  E  -   0.1539  E2      R2  =  0.98 

(4.3)  (-4.0)  D-W  =  3.0 
+  690.9  T        +        6.43  5  SEE  =  322 

(5.7)  (1.8) 

(11)  C    =    0.6584  E  -   0.2129  E2      R2  =  0.98 

(13.5)  (-10.3)  D-W  =  2.5 

+     582.2  T  +     0.215  °C    SEE  =  360 

(5.9)  (1.6) 


Where: 


C 
E 

T 


L 
B 

V 


°c    = 


=  Total  landings 

=  Fishing  effort 

=  Technological  change,  labor 
skills  (1961,  T  =  1;  1962, 
T  =  2;  1963,  T  =  3;  etc.) 

=  Labor  employed  in  the  fishery 

=  Adult  bird  population 

=  Dummy  variable:  closed  sea- 
son V  —  0;  open  season  V  =  1 
Temperature  of  water  in 
Trujillo,  Peru 


Due  to  the  fact  that  the  theoretical  Schaefer 
model  does  not  include  a  constant  term,  the 
estimations  of  the  t-values  of  the  coefficients 
presented  above  are  biased  upwards.  However, 
in  regressions  having  the  constant  term  in  it, 
it  happens  that  this  constant  term  is  not 
significant  in  any  regression  (f-value  around 
0.2).  The  difference  between  coefficients  of 
regressions  with  and  without  the  constant  term 
is  not  significant,  since  in  all  cases  this  dif- 
ference is  less  than  0.4  standard  deviations  of 
the  coefficients. 

In  all  regressions  having  the  constant  term, 
the  variable  technological  change  (T)  is  sta- 
tistically significant  at  the  1%  level  of  signifi- 
cance. In  the  equations  presented  above,  even 
though  the  ^-values  are  biased  upwards,  the 
variables  labor  size  (L),  veda  seasons  (V),  bird 
population  (B),  and  temperature  (°C)  are  not 
statistically  significant.  However,  the  impor- 
tance of  technological  change  (T)  alone  is  such 
that  its  introduction  into  equation  (7)  is  suf- 
ficient to  improve  substantially  the  coefficient 
of  determination  of  the  equation  from  0.84  in 
equation  (6)  to  0.97  in  equation  (7).  Also  the 
Durbin-Watson  statistics  (1.8)  are  now  in  the 
acceptable  range  (1.6-2.4). 

Using  expression  (5)  on  page  61  we  can 
obtain  the  optimal  level  of  fishing  effort  in  terms 
of  the  efficiency  of  1969  vessels.  Equation  (7) 


gives  16.2  million  ton-trips  as  the  optimal  level 
of  fishing  effort.  Equations  (8)  to  (11)  give  the 
following  values  for  optimal  effort  in  terms  of 
million  ton-trips:  15.2,  16.0,  16.0,  and  15.0, 
respectively.  All  these  estimates  are  in  close 
agreement,  but  differ  markedly  from  the  value 
of  23  million  ton-trips  obtained  by  Schaefer 
(1970),  and  from  equation  (6).  However,  because 
of  the  statistical  significance  of  the  variable  "T" 
in  equation  (7),  the  high  autocorrelation  in 
equation  (6),  and  the  theoretical  appeal  of  the 
procedure,  it  seems  that  the  value  of  16.2 
million  ton-trips  is  closest  to  the  true  optimal 
level  of  fishing  effort.  Also,  this  value  makes 
more  sense  in  terms  of  the  data  presented  in 
Table  1.  In  this  table  we  can  see  that  in  1962- 
1963,  with  vessels  of  less  efficiency  than  those 
existing  today,  11.8  million  ton-trips  produced 
6.9  million  metric  tons  of  landings.  A  simple 
extrapolation  would  indicate  that  8.8  million 
tons  of  fish  could  be  landed  by  18.3  million  ton- 
trips  of  vessels  with  1963  efficiency  levels. 

CONCLUSIONS 

The  method  presented  here  appears  useful 
in  obtaining  an  unbiased  estimation  of  the 
optimal  level  of  fishing  effort  in  a  fishery.  It 
adequately  considers  the  effect  of  several  signifi- 
cant inputs  that  cannot  be  directly  introduced 
into  the  traditional  measure  of  fishing  effort. 
Using  this  procedure,  the  optimal  level  of  fishing 
effort  in  the  Peruvian  fishery  is  16.2  million 
ton-trips,  or  only  68%  of  the  level  of  effort  used 
in  Peru  in  1968-1969.  This  result  has  clear 
implications  for  the  management  of  the  Peruvian 
fishing  industry. 

LITERATURE  CITED 

BOEREMA,  L.  K.,  et  al,  1961.  Report  on  the  Effects  of 
Fishing  on  the  Peruvian  Stock  of  Anchovy.  FAO 
Fisheries  Technical  Paper  Number  55.  Rome. 

GULLAND,  J.  A.,  1961.  Fishing  and  the  Stock  of  Fish 
in  Iceland.  Ministry  of  Agriculture,  Fisheries  and  Food. 
United  Kingdom. 

GULLAND,  J.  A.,  1968.  Report  on  the  Population 
Dynamics  of  the  Peruvian  Anchoveta.  FAO  Fisheries 
Technical  Paper  Number  72.  Rome. 

Instituto  del  Mar  del  Peru,  Lima.  1970.  Resumen  General 
de  la  Pesqueria,  1970. 


63 


SCHAEFER.  M.  B.,   1957.  A  Study  of  the  Dynamics  of  the  Anchoveta,   Engraulis   Ringens,   off  Peru.   Instituto 

the    Fishery    for    Yellownn    Tuna    in    Eastern    Tropical  del  Mar  del  Peru,  Bulletin  L-5. 

Pacific  Ocean.  InterAmerican  Tropical  Tuna  Commission 

Bulletin  Number  2.  SCHAEFER,   M.   B.,   1970.   Men,   Birds,  and   Anchovies 

in  the  Peru  Current-Dynamic  Interactions.  Transactions 
of  the  American  Fisheries   Society,   Volume   99,   No.   3. 

SCHAEFER.  M.  B.,   1967.  Dynamics  of  the  Fishery  for  pp.  461-467. 


64 


Natural  Resources  and  External  Economics 

Regulation  of  the  Pacific 

Halibut  Fishery 


Jack  Rich1 


ABSTRACT 

In  a  static,  long  run  competitive  equilibrium  framework,  a  catch  function  allowing 
for  short  run  diminishing  returns  is  combined  with  a  fish  growth  function  developed  by 
Pella  and  Tomlinson  which  facilitates  the  derivation  of  an  expression  for  the  long  run 
marginal  cost  of  "effort"  in  a  common  property  resource  such  as  a  fishery.  This 
expression  takes  into  account  both  "congestion"  and  "growth"  costs.  The  diagramatic 
technique  of  Crutchfield  and  Zellner  is  modified  to  take  account  of  these  externalities. 
The  modified  Crutchfield-Zellner  diagrams  are  used  to  illustrate  the  potential  economic 
losses  from  maximum  sustainable  yield  regulation  or  other  nonoptimal  output. 


INTRODUCTION 

The  task  of  the  International  Pacific  Halibut 
Commission,  as  established  by  treaty  between 
the  United  States  and  Canada,  is  to  regulate 
the  Pacific  Halibut  Fishery  at  maximum  sus- 
tainable yield  (MSY).  The  purpose  of  this  paper 
is  to  develop  a  model  which  will  permit  the 
estimation  of  the  economic  losses  which  may  be 
associated  with  MSY  regulation  or  other  non- 
optimal  output  levels.  The  model  has  certain 
inherent  limitations.  It  is  static,  deterministic, 
partial  equilibrium,  and  ignores  income  dis- 
tribution and  second-best  effects.  Still,  it  may 
be  useful  in  analyzing  a  fishery  not  much 
affected  by  others,  such  as  the  Pacific  halibut 
fishery,  and  in  focusing  attention  on  the  potential 
magnitude  of  economic  losses  resulting  from 
the  present  type  of  regulation  and  from  a 
decentralized,  unregulated  fishery,  although, 
at  least  at  present,  it  does  not  provide  an 
answer  to  the  problem  of  how  long  run  equi- 
librium is  to  be  attained. 

THEORETICAL  FOUNDATION 

The  starting  point  for  the  current  model  is 
the  Crutchfield-Zellner  model  (1962).  Modifica- 


1   Department  of  Economics,  Oregon  State  University, 
Corvallis,  Oregon. 


tions  to  this  model  are  made  which  are  designed 
explicitly  to  account  for  technological  exter- 
nalities resulting  from  the  common  property 
nature  of  the  fishery,  several  of  the  modifications 
having  been  developed  by  Smith  (1969),  Carlson 
(1969),  Bell  (1969),  and  Worcester  (1969), 
among  others.  The  present  paper  develops  a 
framework  for  the  estimation  of  the  rent  and 
consumer  surplus  losses  (conventionally  defined) 
resulting  from  MSY  regulation  or  other  non- 
optimal  output  in  the  static  framework  out- 
lined above. 

Figure  1  depicts  the  Crutchfield-Zellner 
model.  Growth  of  the  fish  stock  biomass  as  a 
function  of  stock  size  is  illustrated  in  Part  A, 
and  has  the  typical  characteristics.  The  de- 
centralized, competitive  supply  and  demand  for 
fish  are  illustrated  in  Part  B,  where  the  in- 
dividual "S"  curves  are  "short  run"  supply 
curves  for  fish  and  show  how  the  amount 
supplied  varies  with  prices,  increases  in  quantity 
resulting  from  additional  units  of  "effort" 
entering  the  fishery  at  higher  prices.  Decreases 
in  fish  stock,  such  as  from  OC  to  OB,  result  in 
an  upward  shift  of  the  S  curves,  from  S-OC 
to  S-OB;  hence,  with  fewer  fish  exposed  to  the 
gear,  the  costs  of  catching  any  given  quantity 
of  fish  are  increased.  The  curve  XX  "traces 
out  the  locus  of  points  on  each  of  these  supply 
curves  which  are  sustainable;  that  is,  where 
the  catch  at  the  corresponding  population  will 


65 


$/lb. 


Part  B:   Demand  and 
Competitive  Decentral- 
ized Supply  of  Fish 


M.-.Y 


(lbs.    per  Unit   Time) 


Figure  1.  —  Industry  demand  together  with  competitive  supply  of  fish. 


leave     population     (biomass)     constant     over 
time"  (Crutchfield  and  Zellner,  1962). 

Since  the  individual  competitive  fisherman 
has  no  control  over  the  size  of  the  fleet  or  the 
stock  of  fish,  these  factors  do  not  enter  the 
decision  making  process  of  the  individual  fisher- 
man, although  they  do  enter  the  cost  function. 
Thus  there  are  technological  externalities 
associated  with  a  fishery  —  a  "congestion" 
cost,  reflecting  the  decreasing  catch  per  unit 
effort  from  a  given  stock  of  fish  as  more  vessels 
enter  the  fishery,  and  a  "growth"  cost,  reflecting 
the  decreased  catch  per  unit  effort  by  a  given 
number  of  units  of  effort  from  a  reduced  biomass 
of  fish,  and  represented  by  the  upward  shifting 
of  the  S  curves  as  the  stock  offish  is  reduced. 


The  curve  XX  is  thus  a  long  run  average 
cost  curve.  A  regulatory  agency  which  has  as 
its  purpose  the  maximization  of  the  net  eco- 
nomic benefits  of  a  fishery  will  have  to  take 
account  of  the  technological  externalities  in- 
herent in  a  common  property  resource,  such 
as  a  fishery. 

Figure  2  adds  Long  Run  Marginal  Cost 
(including  congestion  and  growth  costs)  to  the 
Crutchfield-Zellner  model.  The  LRMC  curve 
is  the  sum  of  the  marginal  congestion  and 
marginal  growth  cost  curves,  and  is  asymptotic 
to  MSY  since,  as  sustainable  yield  harvest 
increases,  equilibrium  fish  biomass  decreases 
(from  its  maximum  level  WmaK  )  until  eventu- 
ally a  further  increase  in  effort  results  in  a 


66 


$/lb 


m  n 


(lbs.   per  Unit   Time) 


Figure    2.   —   Longrun  marginal  cost  added  to  Crutchfield-Zellner 

model. 


zero  increase  in  sustainable  yield.  That  is,  the 
marginal  physical  product  of  another  unit  of 
effort  (in  terms  of  sustainable  yield)  is  zero. 
This  occurs  when  sustainable  yield  is  maximum, 
and  at  this  point  the  cost  of  an  additional  pound 
of  fish  (in  terms  of  sustainable  yield)  is  infinity 
(Carlson,  1969).  In  the  static  framework  of 
this  model,  the  economic  benefits  from  the 
fishery  are  maximized  when  price  is  set  equal 
to  long-run  marginal  cost,  including  congestion 
and  growth  costs  —  that  is,  where  the  extra 
costs  of  an  additional  pound  of  fish  are  just 
equal  to  what  consumers  are  willing  to  pay  for 
that  additional  pound. 

Assuming  a  normal  downward  sloping 
demand  for  fish,  long  run  equilibrium  under  a 
regulatory  agency  which  sets  price  equal  to 
marginal  cost  can  be  determined,  and  this 
equilibrium  can  be  compared  with  that  for  an 
unregulated,  competitive  regime,  and  with 
MSY  regulation. 

Under  a  decentralized,  competitive  regime, 
the  fishery  will  be  in  long  run  equilibrium  where 
the  long  run  average  cost  curve  (including 
normal  returns)  is  equal  to  price  —  point  A  in 


Figure  3  —  with  catch  X0  and  price  P0.  But,  as 
noted  by  Carlson  (1969,  p.  20),  "the  cost  ...  of 
(harvesting)  an  additional  unit  of  fish  [X0B] 
at  this  level  is  in  excess  of  what  consumers  are 
willing  to  pay  for  it"  [X,A].  Since  LRMC  is 
always  above  XX,  a  competitive  fishery  always 
operates  in  long  run  equilibrium  at  a  non- 
optimal  output,  with  too  small  a  stock  of  fish, 
although  the  harvest  may  be  larger  than 
(Figure  3),  equal  to  (Figure  4),  or  smaller 
than  (Figure  5)  the  optimum  level. 

Under  the  present  assumptions  (including 
instantaneous  transfer  of  resources  to  their 
next  best  alternative  use,  and  that  demand 
accurately  reflects  consumer  preferences),  the 
"social"  or  "welfare"  loss  of  a  decentralized 
as  compared  to  an  optimally  regulated  fishery 
is  the  area  ABE  of  Figure  3  —  the  excess  of 
the  extra  cost  above  what  consumers  are  willing 
to  pay  for  the  extra  production  of  fish 
Xo  —  Yi,  beyond  the  level  Xi.  ABE  is  also  the 
extra  value,  above  the  gain  in  consumer  surplus 
PiEAPo  the  resources  used  to  produce  the 
extra  fish  X0  —  Xi  could  have  produced  had 
they  been  used  in  their  next  best  alternative 


67 


PRICE 
$/lb. 


(lbs.  per  Unit  Time) 


Figure  3.  —  "Deadweight"  loss  (Area  ABE). 


(lbs.    per  Unit   Time) 


Figure  4.  —  Identical  competitive  and  regulated  output  rent 
loss  only  (Area  (P0  -  C0)  X0). 


68 


Price 


(lbs.    per    Unit    Time) 


Figure  5.  —  Comparison  of  equilibria:  competitive  output  lower 
than  regulated  output.  (If,  with  demand  as  given,  output 
is  restricted  to  MSY,  the  welfare  loss  is  area  HJKL  plus 
the  shaded  area  above  the  demand  curve  and  to  the  right 
of  LRMS  curve.) 


use.  In  Figures  4  and  5,  a  rent  loss  (PiAGCi) 
is  also  included.  In  Figure  4,  the  entire  loss 
consists  of  rent.  That  is,  output  under  decentral- 
ization and  optimal  regulation  are  identical. 
However,  that  output  would  be  produced  with 
a  much  larger  stock  of  fish,  and  hence  lower 
costs,  under  optimal  regulation  than  under  a 
regime  of  decentralization.  Thus,  all  the  extra 
units  of  effort  used  to  produce  output  X0  are 
"wasted,"  and  could  better  have  been  used  in 
other  industires. 

A  MEASUREMENT  MODEL 

The  derivation  of  marginal  congestion  and 
growth  costs  can  be  expressed  mathematically. 
This  will  permit  estimation  of  the  production 
function,  once  specific  growth  and  catch  func- 
tions are  determined.  With  the  addition  of 
costs  and  demand,  estimation  of  the  welfare 
losses  discussed  above  may  be  achieved. 

Summarizing  all  inputs  under  the  umbrella 
term  "effort"  (E),  catch  (X)  is  a  function  of 
effort  and  the  stock  of  fish  ( W) : 

(1)  X   =  f(E,  W). 


Effort,  catch,  and  stock  can  all  be  expressed  in 
terms  of  the  long  run  equilibrium  catch,  X, 
which  will  give  us  an  expression  in  terms  of 
long  run  equilibrium  catch  alone: 

(2)  E(X)    =   g(X,  W(X)), 

where  E(X)  is  the  effort  associated  with  a  long 
run  equilibrium  catch  of  X,  and  W(X)  is  the 
stock  of  fish  consistent  with  a  sustainable  catch 
of  X  —  i.e.,  one  such  that  dW/dt  =  X.  Since 
cost  is  a  function  of  effort,  we  have,  for  long 
run  equilibrium, 

(3)  C   =    C(X,  W(X)). 

From  (3),  we  can  obtain  marginal  congestion 
cost,  marginal  growth  cost,  and  long  run 
marginal  cost: 


(4)  MCC  =  dc/dx  =   Cx 

(5)  MGC  =  dc/dw  ■  ^  =  Cw  -^ 

ax  dx 


(6)   LRMC  =  MCC  +  MGC  =  Cx+Cw 


dw 
dx 


69 


Equations  (7)  —  (16)  summarize  the  model 
developed  by  J.  J.  Pella  and  P.  K.  Tomlinson 
for  the  Inter-American  Tropical  Tuna  Com- 
mission (1969)  (hereafter  called  the  TC  model). 
The  TC  biological  model  results  in  a  curve 
relating  growth  of  population  to  population 
size.  It  resembles  models  previously  used  by 
the  International  Pacific  Halibut  Commission 
(Southward,  1968)  although  it  is  in  terms  better 
suited  for  economic  analysis.  The  biological 
portion  of  the  TC  model  will  be  used  in  what 
follows  for  an  unexploited  fishery.  In  the  dis- 
cussion of  an  exploited  fishery  modifications 
will  have  to  be  made  to  take  account  of  the 
congestion  phenomenon,  and  this  will  be  achieved 
by  use  of  the  Carlson  "engineering"  function 
for  a  fishery  (1969). 

In  the  TC  model  the  growth  of  the  fish  stock 
is 

(7)  dWt/dt  =  HW"'  -KWt 

where  H,  K,  and  m  are  constants.  Limiting 
population  to  some  absolute  maximum  Wmax  , 
and  integrating  (7)  yields  the  population  at  any 
time  t: 


\  —m  \  —  m  l  —  mi 

(8)    Wt  =  [Wmax   -(Wmax   -Wo        ) 


x  e 


K(\  -m  )t  ,1 


where  W0  is  the  population  at  time  zero,  and 

(9)  Wmax  =(K/H)l/(m~1} 

Further,  Wmsy ,  the  stock  which  yields  the 
maximum  sustainable  yield,  can  be  expressed 

as 

(10)  Wmsy  =  (K/mH)lKm~l) 

The  TC  model  for  an  exploited  fishery  hy- 
pothesizes a  constant  "catchability  coefficient," 
q,  which  is  the  fraction  of  the  population  caught 
by  a  standard  unit  of  fishing  effort  per  unit  of 
time.  The  model  assumes  that  the  instantaneous 
catch  rate,  dXJdt,  can  be  expressed  as: 

(11)  dXt/dt  =  qf,W,, 

where  fi  is  the  number  of  units  of  effort  applied 
to  the  fishery  at  time  t.  It  is  the  assumption  that 
qf  varies  in  the  same  proportion  as  q  or  /  that 
must  be  modified  to  take  account  of  the  con- 
gestion externality.  The  TC  model  implies  a 
constant  short  run  marginal  physical  product 


of  effort,  and  hence  a  constant  short  run  mar- 
ginal cost  of  fish,  at  least  until  the  stock  of 
fish  is  exhausted.  This  assumption  does  not  hold 
for  the  Pacific  halibut  fishery,  and  may  not  hold 
for  any  fishery.  However,  maintaining  the  TC 
assumptions  for  the  moment,  (7)  for  an  un- 
exploited fishery  becomes 

(12)  dWt/dt  =  HW,m  -KWt-qftW, 

for  an  exploited  fishery. 

With  effort  constant  in  the  time  interval  (0,  t), 
and  excluding  those  cases  in  which  the  stock  of 
fish  is  fished  to  extinction,  integration  of  (12) 
yields 


(13)   W, 


\    H 


-(*  +  <?/)(!■ 


m)t 


[k  +  qf 
1/1  -m 


wn  -m)) 


Eliminating  the  time  variable,  and  considering 
only  those  populations  that  have  adjusted  to 
the  given  constant  level  of  effort  (i.e.,  as  t  ap- 
proaches infinity),  we  have 


(14) 


W-i3£jfi 


1  lm 


Biological  equilibrium  when  catch  (X)  is  equal 
to  growth  of  the  fish  stock  is 

(15)   X  =  HW'"  -KW=  qfW. 

From  (15)  we  can  now  express  biological  equi- 
librium catch  as  a  function  of  effort: 


(16)  X  =  qf( 


Qf+k 
H 


i  lm 


To  take  account  of  congestion  externalities 
the  Carlson  "engineering"  function  will  be  used. 
Let  k  be  the  fraction  of  a  stock  of  fish  caught 
by  the  first  unit  of  effort  applied  to  the  fishery; 
assume  that  two  units  of  effort  catch  not  2k 
of  the  original  stock,  but  only  k  +  k(\  —  k)  of  the 
initial  stock.  That  is,  each  unit  of  effort  catches 
a  fraction  k  of  the  stock  remaining  after  all 
previous  units  of  effort  have  been  applied  to 
the  fishery.  For  N  units  of  effort  the  fraction, 
F,  of  a  fish  stock  caught  is 

d7)  f=  i-n-k)x 

where  total  catch  is 


70 


(is)  x  =  a  -a-k)  )w 

or,  writing  W  in  terms  of  N 
(19)^-U-(l-»,']|1-(1-^'     +  V 


That  is,  whenever  we  find  qf  in  equation  (16) 
we  replace  it  with  equation  (17). 

Restricting  ourselves  to  equilibrium  values 
(that  is,  where  catch  is  equal  to  growth), 
differentiation  of  (18)  with  respect  to  N  yields 
the  marginal  physical  product  of  effort  in  long 
run  equilibrium: 

(20)  dX/dN  =  -(l-kfln(l-k)W  +  [1  - 
(l-kf]dW/dN 


The  first  expression  on  the  right  of  (20)  is  the 
short  run  marginal  physical  product  of  another 
unit  of  effort,  and  is  always  positive  for  any 
positive  W,  and  declines  as  N  increases,  thus 
illustrating  short  run  diminishing  returns.  The 
second  expression  on  the  right  of  (20)  shows 
the  effect  on  long  run  equilibrium  catch  of 
another  pound  of  fish  stock,  and  is  equal  to  the 
percentage  of  the  stock  caught  by  N  units  of 
effort  multiplied  by  the  change  in  equilibrium 
stock  resulting  from  a  marginal  change  in 
equilibrium  effort.  Thus,  (20)  includes  both 
congestion  and  growth  externalities. 

Solving  explicitly  for  dW  I  DN  and  rearrang- 
ing terms  can  also  yield  the  expression  for 
dX  I  dN  in  terms  of  W,  N,  and  the  parameters 
k,  m,  and  H: 


(21)  dX/dN  = -[(1— fe)    ln(l-fe) 


-l-(l-fe) 
(m— 1)H 


.N 


w2-m  +W}. 


If  we  assume  that  the  cost  per  unit  of  effort 
is  some  constant  A,  then  the  marginal  cost  per 
pound  of  fish  under  biological  equilibrium  con- 
ditions, and  including  both  congestion  and 
growth  externalities  (the  long  run  marginal 
cost,  LRMC)  is 

(22)   LRMC  =  A KdX/dN). 

Estimation  of  the  parameters  H,  K,  m,  and 
k,  together  with  data  on  demand  and  the  cost 
of  effort  can  be  used  to  estimate  long  run 
equilibrium  catch  and  the  welfare  losses  in  any 
year  associated  with  MSY  regulation  or  other 
nonoptimal  output. 


LITERATURE  CITED 

BELL,  FREDERICK  W.  1969.  Estimation  of  the  Econ- 
nomic  Benefits  to  Fishermen,  Vessels,  and  Society  from 
Limited  Entry  to  the  Inshore  U.  S.  Northern  Lobster 
Fishery,  National  Marine  Fisheries  Service,  Working- 
Paper  Number  36. 

CARLSON,  ERNEST  W.  1969.  Bio-Economic  Model  of 
a  Fishery.  National  Marine  Fisheries  Service,  Working 
Paper  Number  12. 

CRUTCHFIELD,  JAMES  and  ARNOLD  ZELLNER. 
1962.  Economic  Aspects  of  the  Pacific  Halibut  Fishery, 
Fishery  Industrial  Research,  1(1). 

PELLA,  J.  J.  and  P.  K.  TOMLINSON.  1969.  A  Gen- 
eralized Stock  Production  Model,  Inter-American  Tropical 
Tuna  Commission,  Bulletin  13(3). 

SMITH,  VERNON.  1968.  Economics  of  Production  from 
Natural  Resources,  American  Economic  Review,  58(3): 
409-431. 

SOUTHWARD,  G.  MORRIS.  1968.  A  Simulation  of 
Management  Strategies  in  the  Pacific  Halibut  Fishery, 
International  Pacific  Halibut  Commission,  Report 
Number  47. 

WORCESTER,  DEAN  A.,  JR.  1969.  Pecuniary  and 
Technological  Externality,  Factor  Rents,  and  Social 
Cost,  American  Economic  Review,  59(5): 873-885. 


71 


Production  from  the  Sea 


Frederick  W.  Bell,  Ernest  W.  Carlson, 
and  Frederick  V.  Waugh1 


ABSTRACT 

The  sea  constitutes  a  common  property  resource  which  causes  factor  productivity  to 
be  heavily  influenced  by  technological  externalities.  The  sea  is  also  subject  to  the  spectre 
of  Malthusian  scarcity  since  man  cannot  manipulate  the  ocean  environment  (Barnett  and 
Morse,  1963).  We  estimated  the  parameters  using  ordinary  least  squares  of  the  dynamic 
Schaefer  production  model  of  the  intervention  of  man  into  the  oceanic  ecosystem.  A 
second  production  model  for  the  sea  to  specify  diminishing  returns  to  capital  and  labor 
for  any  fixed  biomass  was  developed.  The  parameters  of  the  latter  model  were  estimated 
by  a  computer  search  technique.  The  results  indicate  that  the  industry  production 
function  for  marine  life  is  subject  to  diminishing  physical  returns  to  capital  and  labor.  For 
the  cases  considered  in  this  study  it  also  appears  that  the  parabolic  yield  function 
developed  by  Schaefer,  assuming  constant  returns  to  factors  inputs,  is  not  as  realistic 
as  a  production  function  with  diminishing  returns  to  inputs  with  a  given  biomass. 


INTRODUCTION 

After  explaining  the  principle  of  diminishing 
returns  in  agriculture,  that  great  economist, 
Alfred  Marshall  (1920,  p.  166)  wrote: 

As  to  the  sea,  opinions  differ.  Its  volume  is  vast, 
and  fish  are  very  prolific;  and  some  think  that  a 
practically  unlimited  supply  can  be  drawn  from  the 
sea  by  man  without  appreciably  affecting  the  numbers 
that  remain  there;  or  in  other  words,  that  the  law  of 
diminishing  returns  scarcely  applies  at  all  to  sea- 
fisheries;  while  others  think  that  experience  shows  a 
falling-off  in  the  productiveness  of  those  fisheries 
that  have  been  vigorously  worked,  especially  by  steam 
trawlers.  The  question  is  important,  for  the  future 
population  of  the  world  will  be  appreciably  affected 
as  regards  both  quantity  and  quality,  by  the  available 
supply  offish. 

We  have  waited  50  years  to  answer  Marshall's 
question.  We  must  not  wait  much  longer.  The 
world's  population  will  double  by  the  year  2000. 
What  will  happen  to  the  production,  prices, 
and  consumption  offish  (Bell  et  al.,  manuscript)? 

As  in  Marshall's  day,  some  doubtless  still 
think  that  the  future  supply  offish  is  practically 


1  The  authors  are  respectively  Chief  and  Economist, 
Economic  Research  Laboratory,  National  Marine 
Fisheries  Service,  and  Professor,  Department  of  Agri- 
cultural Economics,  University  of  Maryland.  The  ideas 
expressed  in  this  article  do  not  necessarily  reflect  the 
official  position  of  the  National  Oceanic  and  Atmospheric 
Administration  (NOAA). 


unlimited.  But  those  biologists  and  economists 
who  are  studying  fisheries  doubt  this.  They 
know  that  some  species  offish  have  already  been 
"overfished";  that  is,  increased  inputs  of  capital 
and  labor  have  actually  reduced  yields.  Examples 
are  menhaden  and  haddock  in  the  Atlantic 
fisheries.  Biologists  have  found  that  the  catches 
of  eastern  tropical  Pacific  yellowfin  tuna  and  of 
northeastern  Pacific  halibut  have  reached  their 
"maximum  sustainable  yields."  International 
controls  have  been  found  necessary  to  prevent 
depletion  of  the  aforementioned  species. 

Of  course,  these  are  only  a  few  of  the  many 
species  of  commercial  fish.  But  we  doubt  if  any 
fishery  biologist  today  would  be  among  those 
who  Marshall  said,  "...  think  that  a  practically 
unlimited  supply  can  be  drawn  from  the  sea." 
To  be  sure,  the  sea  is  vast,  but  Ryther  (1969),  a 
prominent  biologist,  says: 

The  open  sea —  90%  of  the  ocean  and  nearly  three- 
fourths  of  the  earth's  surface  —  is  esentially  a 
biological  desert.  It  produces  a  negligible  fraction  of 
the  world's  fish  catch  at  present  and  has  little  or  no 
potential  for  yielding  more  in  the  future. 

Upwelling  regions,  totaling  no  more  than  about 
one-tenth  of  1%  of  the  ocean  surface  (an  area  roughly 
the  size  of  California)  produce  about  half  the  world's 
fish  supply.  The  other  half  is  produced  in  coastal 
waters  and  the  few  offshore  regions  of  comparably 
high  fertility. 

We  could  cite  many  other  fishery  biologists 


72 


to  indicate  that  the  potential  supply  of  fish 
from  the  sea  is  limited.  But,  even  if  there  were 
no  fixed  limit  to  fish  production,  we  believe  that 
diminishing  returns  would  apply  to  fisheries  at 
least  as  much  as  to  agriculture;  perhaps  more. 
This  has  important  implications  to  public 
policies,  as  Marshall  noted.  Hence,  the  purpose 
of  this  article  is  to  explore  the  production 
function  for  the  sea. 


DIMINISHING  RETURNS  OF  FISHERIES 

Marshall's  (1920,  p.  150)  first  statement  of 
the  law  of  diminishing  returns  in  agriculture 
was: 

An  increase  in  capital  and  labour  applied  in  the 
cultivation  of  land  causes  in  general  a  less  than 
proportionate  increase  in  the  amount  of  produce 
raised,  unless  it  happens  to  coincide  with  an  im- 
provement in  the  arts  of  agriculture. 

In  the  case  of  fisheries,  indices  of  capital  and 
labor  inputs  are  known  as  "effort."  Diminishing 
returns  from  fishing  means  (paraphrasing 
Marshall)  that  an  increase  in  effort  results  in 
less  than  a  proportionate  increase  in  the  yield 
of  fish,  assuming  no  change  in  technology. 
Thus,  if  effort  were  doubled,  the  yield  would  be 
less  than  doubled. 

But  if  we  are  to  manage  the  world's  fisheries 
well,  we  need  more  than  general  comments 
about  diminishing  returns  —  we  need  usable 
estimates  of  the  effort-yield  functions  for  the 
major  species  of  fish.  Schaefer  (1954)  wrote  a 
pioneering  paper  on  the  theory  and  measurement 
of  such  functions.  In  recent  years,  many  bi- 
ologists have  added  to  the  theory  in  this  area, 
and  have  presented  important  statistical  veri- 
fications and  measurements  (Pella  and  Tomlin- 
son,  1969;  Fox,  1970). 

The  necessary  theory  is  in  two  parts:  (1)  the 
theory  of  biological  growth,  and  (2)  the  theory 
of  yield  from  a  given  biomass. 

Theory  of  Biological  Growth 

First,  consider  biological  growth  —  for 
example,  the  growth  of  "biomass"  or  the  total 
weight  of  marketable  fish.  Schaefer  (1954), 
hypothesized  that  if  there  were  no  fishing,  the 
growth  curve  of  the  biomass  would  look  some- 


biomass 


Figure  1.  —  Growth  with  no  fishing. 


thing  like  that  shown  in  Figure  1.  The  species, 
in  each  region,  would  tend  to  approach  some 
maximum  biomass,  M.  Here  natural  mortality 
would  just  offset  recruitment  (from  young  stock) 
and  growth  in  body  size. 

A   curve  commonly  used   to  represent  such 
growth  is  the  logistic,2 


(1)  m, 


M. 


1  +  be 


where  mt  is  the  biomass  at  time  t,  M  is  the 
potential  maximum  biomass,  e  is  the  base  of 
natural  logarithms,  t  is  time,  and  a  and  b  are 
parameters.  (We  shall  generally  measure  time 
in  years.)  Davis  (1941)  discussed  the  proper- 
ties of  this  curve  in  detail,  and  gave  many 
references  to  its  uses  in  biology  and  in  the  study 
of  growth  of  human  populations.  Its  derivative 
is: 


2  Most  work  using  the  logistic  has  been  done  with 
numbers  in  populations,  here  we  are  applying  it  to  the 
total  weight  of  the  population.  Tomlinson  and  Pella 
(1969)  have  suggested  that  the  following  function  be 
used  to  approximate  biological  growth: 


'  HP'"  It)  -KP(t). 


When  m  =  2,  the  growth  function  becomes  the  well- 
known  logistic  or  as  used  by  Gulland,  an  autocatalytic 
equation.  Fox  (1970)  has  suggested  a  Gompertz  function 
to  approximate  biological  growth. 


73 


(2)   dmr/dt=  am 


■(i-jt)- 


So    the    proportional    rate    growth    (with    no 

fishing)  is: 

dmt    _  n  /-,     mt\ 
The  second  derivative  of  (1)  is: 


d  ~mt        2 


(i-SO^) 


Maximum  absolute  growth  occurs  when  (4) 
equals  zero;  that  is,  when  mt  =  V2M  (when 
current  biomass  is  one-half  the  potential 
maximum).  At  that  point,  equation  (2)  shows 
that  the  maximum  growth,  dmt/dt  =  aM/4. 

Suppose  aM/4  were  taken  from  the  biomass 
each  year  by  fishermen:  each  year,  the  biomass 
would  grow  by  aM/4;  biological  growth  would 
just  offset  the  amount  taken  by  fishermen;  and 
there  would  be  a  steady-state  equilibrium. 

The  Theory  of  Yield 
from  a  Given  Biomass 

We  now  consider  how  yield  responds  to 
effort  when  we  abstract  from  changes  in  biomass. 
Schaefer  (1954)  made  the  simple  assumption 
that  the  catch  mt  would  be  proportional  to 
effort,  k  is  the  constant  of  proportionality,  and 
xt  is  effort: 

(5)  y,/m,  =  kxt. 

Schaefer  assumed  that,  with  a  given  biomass, 
there  would  be  constant  returns  to  effort;  dou- 
bling the  effort  would  double  the  yield,  tripling 
the  effort  would  triple  the  yield  —  and  so  on. 
As  a  first  approximation,  this  may  be  adequate 
in  many  cases  within  the  observed  range  of  the 
data.  Schaefer  and  others  have  used  it  to  make 
many  important  estimates  of  maximum  sus- 
tainable yield;  and  as  a  basis  for  economic 
controls. 

But  we  think  that  a  more  realistic  catch 
function  is: 

(6)  y,/m,  =  (l-zXr), 

with  0  <z  <  1,  and  with  m,  fixed. 

The  rationale  of  (6)  was  explained  by  Carlson 
(1969).  Briefly,  assume  that  the  original  biomass 
is  m,  and  that  one  unit  of  effort  will  catch  pmr, 


leaving  (1— p)mT\  assume  that  the  next  unit 
of  effort  will  catch  the  same  proportion  of  the 
remaining  biomass  —  that  is,  it  will  catch 
p(l—p)mr,leaving(l—p)2mr.  The  same  reasoning 
shows  that  n  units  of  effort  will  catch  (1  m  ,)  . 
In  equation  (6),  we  simply  let  z  =  1— p.  We 
believe  that  on  an  a  priori  basis  (6)  is  more 
realistic  than  is  (5).  But  probably  there  is  no 
magic  mathematical  formula  that  is  exactly 
right  for  all  species  and  for  all  amounts  of  effort. 

yield,   y 


Figure  2.  —  Two  yield  functions.  (Based  upon  equations 
5  and  6,  assuming  that  one  unit  of  effort  yields  one- 
half  of  the  existing  biomass.) 


Figure  2  compares  the  growth  functions 
represented  by  equations  (5)  and  (6).  Each 
assumes  that  one-half  the  existing  biomass  was 
caught  with  one  unit  of  effort  in  some  base 
period.  (The  units  are  arbitrary.  We  find  it 
desirable  to  "normalize"  both  yield  and  effort  by 
dividing  by  the  base-period  data.)  Note  that 
equation  (5)  would  indicate  that  the  entire 
biomass  would  be  caught  with  two  units  of 
effort.  But  equation  (6)  would  indicate  that  if 
effort  were  increased  indefinitely,  the  existing 
biomass  would  be  approached  as  a  limit,  but 
never  quite  reached.  Within  the  observed  range 
of  historical  data,  it  may  not  be  easy  to  choose 
between  the  two  curves  in  Figure  2.  But  they 
give  far  different  results  when  they  are  ex- 
trapolated to  estimate  the  effects  of  large  in- 
creases in  effort.  This  is  especially  critical  where 
one  must  make  forecasts  of  the  likely  effect  of 
the  expansion  in  fishing  effort. 


71 


STATIONARY  STATE  EQUILIBRIUM 

The    stationary    state   equilibrium    is    found 
bj'  letting  annual  yield  equal  annual  growth : 


where  y  is  the  equilibrium  yield  and  mt  is  the 
corresponding  biomass.  Thus,  Schaefer  let: 

(8)  kxr  =  a{l-^L) 
Solved  for  mt 

(9)  mt=M(l-k-f) 

and  got  the  equilibrium  yield  as  a  function  of 
effort: 

(10)  yr  =  rhtkxt  =  Mkxt  (l  -  -^)  ■ 

This  is  a  simple  quadratic.  To  estimate  it 
from  statistical  data  using  ordinary  least- 
squares,  we  write: 


(11)  yt=Axt-Bxi 

where  A  =  Mk  and  B  = 


Mk' 


The  graph  of  (11)  is  shown  in  Figure  3.  Note 
that  while  Schaefer  assumed  constant  returns 
from  a  fixed  biomass,  his  curve  of  equilibrium 
yield    indicates    decreasing    returns.    In    fact, 


equilibrium  yield 


Figure    3. 


effort 


Equilibrium    yield-effort    with    constant 
returns. 


average  yield  per  unit  of  effort  is  easily  seen 
to  be  (by  dividing  (9)  by  x), 

(12)  yt/xt  =A  -Bxt  . 

If  we  use  (6),  instead  of  (5)  as  an  estimate 
the    response   of  yield   to   effort   with   a   fixed 
biomass,  we  have: 

(ia)i-<"-.(i-Sf). 

Solving  for  m,  we  find : 

(14)  mr=M[l-(l-~)]- 

So  the  steady-state  equilibrium  yield  is: 

(15)  y,=mt(l-zXt)  = 

m[(i-/-)-±(w;')1 

that  is, 

(16)  yt  =  C(l-zXf)-D(l-zXtf 

where  C  =  M  and  D  =  Mia. 

This  is  not  as  easy  to  fit  statistically  as  is  the 
Schaefer  function  (11).  It  can  be  handled  without 
undue  difficulty  on  a  computer  by  a  "search 
method,"  trying  a  series  of  values  for  z;  in  each 
case  computing  R2,  the  Durbin-Watson  statistic 
(D-W),  and  the  t  values  of  the  two  regression 
coefficients;  then  by  interpolation  we  find  the 
"best"  fit. 

Equations  (15)  and  (16)  indicate  decreasing 
returns  to  effort.  Their  graph  is  like  that  in 
Figure  4.  In  this  case  —  which  we  think  is 
more  realistic  —  we  get  diminishing  returns 
for  two  reasons: 

1.  Because  annual  growth  declines  as  the 
fish  population  increases,  and 

2.  Because  the  yield-per-unit-of -effort  de- 
clines with  effort;  that  is,  doubling  the  effort 
will  result  in  less  than  doubling  the  yield,  even 
with  a  fixed  biomass.  The  net  result  is  a  much 
flatter  curve  after  MSY  is  reached. 

STOCK  ADJUSTMENT  MODEL 

So  far,  we  have  considered  only  the  steady- 
state  equilibrium.  This  assumes  that  full  adjust- 


75 


equilibrium  yield 


MSY 


Figure  4. 


effort 


Equilibrium  yield-effort  with  diminishing 
returns. 


ment  is  made  instantaneously,  thus  the  present 
catch  is  a  function  of  the  present  effort  only. 
This  may  give  a  satisfactory  approximation  for 
some  species.  But  in  other  species,  several  time 
periods  may  be  required  to  establish  a  new 
equilibrium.  In  such  cases,  current  yields  are 
affected  not  only  by  current  effort,  but  also  by 
the  efforts  of  several  past  periods. 

That  is,  annual  observations  on  catch  and 
effort  do  not  represent  equilibrium  observations. 
To  remedy  this  situation,  biologists  have  sug- 
gested various  adjustments  to  the  data  (Appen- 
dix I). 

In  reality,  the  observed  catch  in  any  given 
year  may  be  the  result  of  effort  expended  in 
previous  periods;  i.e.,  the  observed  catch  is 
some  kind  of  weighted  average  of  catch  produced 
by  fishing  effort  in  previous  periods.  The  Gulland 
procedure  employs  a  similar  assumption  in  that 
it  assumes  that  this  year's  observed  catch  is 
parabolically  related  to  a  simple  average  of 
previous  effort.  An  alternative  specification  of 
the  yield  effort-relation  for  many  stocks  of  fish 
may  take  the  following  form  (assuming  for 
example  a  logistic  and  constant  returns  equi- 
librium relation): 


(17)  yt  -  axt  —  bxt  +  ax  xt-\ 

+  .  .  .€t  ■ 


2 


Let  us  now  make  the  classic  assumptions  about 
the  disturbances,  et,  of  constant  variance  and 
zero  covariance. 

Although  (17)  is  a  general  specification  of  the 
yield-effort  relationship,  its  estimation  presents 


obvious  difficulties.  Since  our  sample  will  be 
finite  in  size,  the  infinite  set  of  lagged  regressors 
must  be  terminated  at  some  point.  Also,  there 
is  likely  to  be  colinearity  among  the  successive 
regressors. 

One  way  of  solving  the  problem  is  to  hypothe- 
size that  the  coefficients  on  the  lagged  variables 
diminish  in  size  as  the  time  period  is  more 
distant  from  the  present  observation  on  catch. 
Put  differently,  let  us  hypothesize  that  the 
coefficients  on  successive  x's  decline  systemati- 
cally as  we  go  further  back  in  time.  This  was 
suggested  by  Fisher  (1925);  more  recently  it  has 
been  revived  and  extended  by  Koyck  (1954)  and 
by  Nerlove  (1958).  We  shall  call  this  a  Koyck 
specification.  Koyck  hypothesized  that  a  useful 
approximation  would  be  that  the  coefficients 
of  (17)  decline  geometrically: 

(18)   ak  =  a\k    (k  =  0,  1,  .  .  .  )  and 


(19)   bk  =b\      (fc-0, 1,  ...). 


(17)  may  be  rewritten  as  the  following: 3 


(20)  yt  =  axt  —  bxt  +  Xaxt- 1 

—  Xbxt- 1  +  .  .  .  et  . 

If  we  lag  (20)  by  one  period  and  multiply  by  A. 
we  obtain 


7 

(21)  Xyt_  i  =  Xaxt_  ^  —  Xbxt_  ] 

2  2         2 

+  X  axt_  2  —X  bxt_  2  +  .  .  .  Xet _  i 


Now,  subtract  (21)  from  (20)  and  rewrite: 

(22)  y,  =  ax,  —bx,  +  Xyt-  i  +  e, 
where 

(23)  et  =  et  —  Xet-\   . 


3  Equation  (20)  may  be  interpreted  to  mean  that 
observed  catch  depends  on  this  year's  effort  (a  common 
assumption  used  by  many  population  dynamicists)  plus 
effort  expended  in  previous  periods.  This  is  merely  a 
hypothesis  that  can  be  tested  empirically. 


76 


Equation  (22)  may  be  estimated  using  ordinary 
least-squares.4 

Nerlove  provides  an  alternative  theory  to 
justify  (22).  Suppose  that  xt  determines v,*  ,  the 
"equilibrium  value"  of  catch, 

(24)  yt*  =  axr  -  bx] , 

but  that  the  adjustment  to  the  equilibrium  value 
in  one  period  is  only  gradual  (i.e.,  not  complete): 

(25)  yt—yt-\  =  b(yt*  —yt-i) 

where  0  <  6  <  1  is  the  coefficient  of  adjustment. 
Inserting  (24)  into  (25)  and  rewriting  gives 
the  same  form  as  (22): 

(26)  yt  =  abxt  —  bhx]  +  (1—5)  yt_  i 
where  (1—5)  =  X. 

Using  (26)  or  (22),  we  may  also  compute 
how  many  periods  it  takes  one-half  the  gap  to 
be  filled.  If  yt-  i  is  in  equilibrium,  then  the  gap 
at  period  t(Gt)  is  equal  to  the  following: 

(27)  (yt*-yt-i)  =  Gt. 

Each  period  a  constant  percentage  of  the  re- 
maining gap  is  filled;  so  that  at  time  t  +  k 
the  remaining  gap  is 

(28)  Gt+k  =Gt  (1-5  f  . 

If  K  =  0,  (29)  indicates  that  all  the  gap  remains 
to  be  filled.  When  will  one-half  of  the  initial 
gap  be  filled?  This  may  be  found  by  substituting 
y2Gt  for  Gt+K  ,  or 

(29)  Gf  (1-5)*'  =1/2. 
Hence, 


4  In  essence,  a  researcher  attempting  to  estimate  the 
parameters  of  the  yield  function  can  ran  the  following 
regressions:   y,  -  ax,  -  bx?  ■, 


x,  —  bx, 

+  Xy,- 

or  y,  = 

fc+x,. 

i  +... 

♦  «,  .1 

L 

n  +  t 

j 

t["  + 

t,    i  * 

..♦«,.  „-], 

L 

n  * 

i       J 

where  (n  +  1)  is  the  number  of  years  the  fish  are  in 
the  fishery.  The  latter  is  the  Gulland  technique  where 
the  first  two  specifications  are  with  and  without  the 
Koyck  formulation  respectively.  Equation  (20)  may  be 
specified  as  the  following: 

(y/x)t  =  a  —  bxt  —  \bxt_  !  -  ...  -X  bxt_k. 

With  this  form,  the  final  estimating  equation  will  have 
(yjx)  as  a  lagged  independent  variable. 


(30)  (1-6)A  =i/2, 
or 

(31)  K  =  log  1/2  *  log  (1-5)  =     log  2 

log(i^5) 

K  is  the  "half-life";  that  is,  the  number  of 
periods  required  to  cut  the  gap  in  half.  In  2K 
years,  the  gap  will  be  reduced  to  V4  ;  in  3A' 
years  to  Vb  .  .  .  and  so  on.  It  would  never  com- 
pletely disappear.  In  theory,  K  should  be 
related  to  the  following  biological  factors: 

(1)  Fertility  of  the  species  (i.e.,  number  of 
eggs  laid  and  reaching  full  term); 

(2)  Rate  of  growth  of  the  species  (i.e.,  how 
many  periods  it  takes  to  reach  maturity).  K 
should  be  large  for  relatively  unfertile  and 
slowly  growing  species  and  small  for  very 
fertile  and  rapidly  growing  species. 

In  sum,  we  are  interested  in  eight  estimating 
equations.  First,  a  group  of  four  equations  based 
upon  the  assumption  of  constant  returns  from 
a  fixed  biomass;  these  are  all  designated  LCR 
(logistic  constant  returns).  LCRa  is  the  static 
function  with  total  yield,  yti  dependent.  LCRb 
is  the  same  with  average  yield  per  unit  of 
effort,  yt/xt,  dependent.  Then  LCRaS  and 
LCRbS  are  lagged  or  stock  adjustment  models. 
This  gives  us  four  functions.  There  are  four 
more  (designated  LDRa,  LDRb,  LDRaS,  and 
LDRbS)  based  upon  the  assumption  of  decreas- 
ing returns  from  a  fixed  biomass.  Finally,  we 
have  included  an  estimate  of  the  parameters  of 
LCRa  using  the  Gulland  technique  for  adjusting 
the  effort  series.5 


RESULTS  OF  THE  ANALYSES 

In  order  to  illustrate  the  applicability  of  our 
theoretical  yield  functions,  we  selected  five 
species  for  consideration:  (1)  Chesapeake  Bay 
menhaden;  (2)  Atlantic  and  Gulf  blue  crab; 
(3)  Atlantic  longline  tuna;  (4)  Soviet  and 
Japanese  king  crab  fishery  in  the  eastern 
Bering  Sea;  and  (5)  Cape  Flattery  sablefish. 


5  For  the  five  fisheries  studied  (below)  the  fish  are 
in  the  fishery  about  two  years.  Therefore,  a  two-year 
moving  average  of  effort  was  computed. 


77 


Chesapeake  Bay  Menhaden 

Table  1  shows  the  empirical  results  for  this 
fishery.  Based  upon  the  R2  criterion,  LDRa 
represented  the  "best"  function  where  total 
catch  was  used  as  an  independent  variable. 
There  is  no  doubt  from  the  statistical  analysis 
that  the  Schaefer  function  (LCRa)  is  definitely 
inferior  when  compared  to  the  LDRa  model  in 
its  ability  to  describe  the  catch-effort  relation 
in  the  menhaden  fishery.  No  evidence  of  auto- 
correlation was  detected  in  the  LDRa  function. 
As  shown  by  LDRaS,  there  seems  to  be  no 
stock  adjustment  effect  as  the  coefficient  on  the 
lag  variable  is  not  statistically  significant. 
Among  all  the  functions,  LDRb  shows  the  best 
fit  when  catch  per  unit  of  effort  is  used  as  the 
dependent  variable.  From  a  theoretical  point  of 
view,  there  should  be  no  difference  between  the 
"a"  and  "b"  functions.  However,  the  statistical 
estimation  procedure  does  yield  two  estimators 
for  each  parameter.  LDRa  and  LDRb  do  yield 
similar  estimates  of  y*  and  a;*.  Also,  the 
Gulland  -LCRb  equation  yielded  very  similar 
estimates  of  y*  and  x*  as  the  LCRb  (unadjusted 
data).  Further  the  choice  between  the  "a"  and 
"b"  functions  should  be  made  on  the  basis  of 
just  what  one  wants  to  predict  —  catch  or  catch 
per  unit  of  effort.  The  LCR  and  LDRa  functions 
are  shown  in  Figure  5.  It  should  be  noted  that 
in  equation  (16)  M  =  A.  Thus,  the  LDRa 
equation  estimated  by  least-squares  will  also 
yield  the  maximum  biomass  without  fishing. 
That  is,  MSY  =  M/4  =  158.7  thousand  tons.  M 
is  therefore  equal  to  634.8  thousand  tons.  The 
logistic  function  can  be  directly  computed  since 
a  =  AIB  and  a  =  1.1512,  and  if  t  -  0  at  the 
point  of  maximum  growth,  then 


m, 


M_  634.8 
2     1  +be 


0  ,  so  b  =  1,  or 


m,  = 


634.8  thousand  tons 
1  +  e-i.isi2f 


This  is  one  additional  advantage  of  the  LDRa 
over  the  LCRa  function. 

Atlantic  and  Gulf  Blue  Crab 

Table  2  shows  the  empirical  results  for  this 
fishery.  Based  upon  the  R2  criterion,  it  would 


seem  that  we  have  little  basis  on  which  to 
choose  between  the  LCRaS  and  the  LDRaS 
models,  each  having  an  R2  of  0.94.  Both  show  a 
strong  stock  adjustment  effect.  The  half-life 
for  the  adjustment  process  was  0.57  years.  In 
this  case,  the  data  cannot  adequately  distinguish 
between  the  two  functions.  The  MSY  ranges 
from  129.6  million  pounds  in  the  LCRaS  model 
to  189.0  million  pounds  in  the  LDRaS  model. 
The  autocorrelation  test  for  the  two  functions  is 
inconclusive.  Hence,  the  choice  between  the 
functions  must  be  made  on  a  priori  grounds. 
Since  the  LDRaS  model  seems  more  plausible 
on  a  priori  grounds,  it  would  seem  that  this 
function  should  be  selected  for  fishery  manage- 
ment purposes.  As  the  fishery  expands,  addition- 
al data  will  be  generated  to  verify  the  existence 
of  one  or  the  other  function.  This  general 
prescription  will  probably  apply  to  many 
fisheries  where  data  are  only  available  in  the 
upward  expansion  phase  (i.e.,  catch  is  below 
MSY).  Finally,  as  with  Chesapeake  Bay  men- 
haden, there  seems  to  be  little  difference  between 
Gulland  LCRb  and  LCRb  unadjusted.  Figure  6 
shows  the  two  functions  discussed  above. 

Atlantic  Longline  Tuna 

Table  3  shows  the  results  for  the  Atlantic 
longline  tuna  fishery.  On  the  basis  of  R2,  the 
LDRa  model  is  superior  in  predicting  changes 
in  catch  in  response  to  effort.  The  stock  adjust- 
ment coefficient  was  not  statistically  significant. 
The  autocorrelation  test  is  inconclusive  for 
LDRa.  The  MSY  for  the  LDRa  function  is  106.7 
thousand  metric  tons  with  140.1  million  hooks 
of  effort.  Notice  that  the  MSY's  associated  with 
the  LCRa  and  LDRa  functions  are  not  appre- 
ciably different;  however,  the  number  of  hooks 
necessary  to  harvest  MSY  is  vastly  different. 
This  is  due  to  the  flatness  of  the  function 
generated  by  the  LDRa  model.  The  Gulland- 
LCRb  gives  a  much  higher  estimate  of  y*  and 
a  lower  estimate  of  x*  than  the  unadjusted 
LCRb.  Figure  7  shows  the  LCRa  and  LDRa 
functions. 

Bering  Sea  King  Crab 

Table  4  shows  the  results  for  the  Bering  Sea 
king  crab  fishery.  On  the  basis  of  R2,  the  LDRa 
model  is  the  best  in  "explaining"  the  catch-effort 


78 


182.6 


158.7 


LDR„ 


LCR„ 


o.oo     50.00     loo.oo    150.00     200.00 

EFFORT:   NUMBER  OF  UNITS 


250.00 


rf 


•     •" 


0.00  50.00  100.00  150.00  200.00 

EFFORT: NUMBER  OF  UNITS 


250.00 


Figure  5.  —  Chesapeake  Bay  menhaden  fishery,  1946-68:  Catch,  effort, 
and  catch  per  unit  of  effort. 


79 


r 


M 


I 


"u  ~* 

<—  D 
X  => 

q 

no 

■tf 

oo 

■* 

■* 

— 

NO 

rt 

00 

00 

»N 

in 

d 

oo 

00 

00 

OC 

o 

(N 

-H 

r-j 

-^ 

ON 

ON 

ON 

o- 

3| 

c 

o 

00 

C-) 

m 

oo 

in 

m 

c 

OC 

N 

rr 

00 

NO 

oo 

00 

q 

IO 

ON 

t-- 

o 

NO 

r-4 

*t 

VI  *> 

CN 

<N 

en 

r-' 

_' 

oo' 

r^" 

oo' 

fs 

^    0 

CN 

00 

~m 

oo 

m 

m 

NO 

in 

NC 

{N 

•— * 

CM 

<•■ i 

~* 

-H 

•-H 

•^ 

. 

o 

!s<2 

VO 

ON 

< 

< 

00 

V") 

Z 

Z 

"3 

^ 

u 

o 

2 

£ 

< 

00 

> 

> 

o 

O 

<N 

m 

NO 

r- 

r~) 

. — 1 

C-" 

t 

"* 

Q 

m 

q 

•H 

oo 

1; 

ON 

rn 

00 

<^~ 

C-) 

t-; 

o> 

C- 

■* 

00 

m 

(^1 

m 

rs 

Qi 

(N 

r^ 

■* 

p» 

NO 

00 

NO 

OC 

O) 

— 

I 

' 

>, 

~- 

>, 

II 

X 

4t 

r- 

+ 

-1 

r^j 

q 

q 

1 

OJ 

r-l ' 

X 

X 

Cn] 

>>  Ix 

o 

00 

m 

X 
in 
Cn! 

X 

X 

in 

X 

o 
m 

X 

m 

V0 

00 

fNj 

01 

tt 

On 

ON 

i 

E 

1 

tS     ^j 

•* 
■* 

t— 

© 

ro 

2 

NO 
<S 
CnI 

in 

ON 

in 

a 

+ 

m 

ON 

OO 

— 

ON 

X 

+ 

NO 

. 

m 

00 

C3 

cs 

(S 

m 

TT 

i — 

rn 

'-» 

NO 

m 

X 

^r 

in 

o 

o 

X 

nf 

■^; 

E 

•<* 

q 

X 
m 

ON 

oo 

r— 

m 

00 

rn 

X 

o 

CN, 

x 

w 

M 

i 

X 

1 

m 

1 

X 

in 

ro 

X 

X 

o 

" 

i 

1 

X 

1 

' 

■* 

— 

00 

*-• 

c— 

^H 

(^1 

00 

T 

(N 

^H 

x 

00 

00 

no 

in 

On 

o 

cs 

r^ 

rn 

f>l 

in 

-i- 

+ 

m 

ON 

00 

ON 

■<t 

M 

NO 

X 

m 
■* 

oo 

ii 

NO 
ON 

m 

ii 

"* 

-1- 

II 

~ 

II 

n 

II 

II 

_  - 

II 

_ 

- 

>, 

>. 

>TTx 

>, 

>TTx 

>. 

>.l  X 

>>l . 

_, 

^ 

ai 

5 

U 

Vi 

n 

as 

O0 

O0 

oo 

-= 

_i 

-3 

71 

as 

tA 

as 

aS 

2 

(5 

ai 

e 

o 

U 

u 

o 

Q 

Q 

Q 

Q 

•  T* 

ii 

_J 

_i 

_J 

_i 

_) 

_) 

_J 

_J 

o  2 


c  as 

a; 

3 

> 

-    h 

a> 

5  B 
—  c 

fc^    SO 

0)  — ' 

R  x, 

3    2 

-c  >, 

E  b 

*J    3 

x   « 

ca  T3 

E  c 

C    3 

2  o 

it    « 

JC 

a.  w 

^"S 

.52  ^J 

-a  2 

C 

-  c 

C    a) 

T3  — 

O  c 

to  > 

-    tiljO 

o  — 

c 

HI 

s^ 

0 

,.  -o 

*-    2    3 

o   Jj 

'  1 

c  c<*- 

n.2 

.a  -o 

c 

3 

u  «  o 

»H     >   <*- 

T3    C 

^ 

bug 

C    3 

u 

'I 

in  ^   3 

ST  s 

■^ 

c  o  2 

5  sf 

«  8 -a 

Q  8 

■J 

o  « 

G 

E 

DC 

W5     cd 

0) 

> 

c 

ifl 

W  <u 

3 

E 

-r 

C  JS 

a-O'S 

On 

o  ■" 

= 

3 
•T1 

5S  as 

3:  £ 

o 

O 

B 

TJ 

CO    5*   tfl 

C     M 

Ul 

«n    ^    W 

u 

£  E 

"-» 

'J 

C/5 

c  aSI 

u 

3   to 

Q  U 

Q 

II 

n   ii 

E 

< 

> 

oo  tt  — .      oo 


E  2 

o  J3 

iota 
•a  H 

X  — 

'—   o 
0  " 


5S& 

i  NO    & 


X  J2     CO 


—    ?5    n  .. 

'S  e  B ' 

o  o  d 


0   > 


5S1 

-I  " 

<  >  ■— 


ii    ii    ii    ii    ii 
as  as 


80 


100.00  180.00  280.00  3U0.00 

EFFORT:   POTS   (X103) 


1 

U20.00 


500.00 


100.0 


180.00 


200.00 


300.00  1+20.00 


500.00 


EFFORT  :POTS    (XLO   ) 


Figure  6.  —  Atlantic  and  Gulf  blue  crab  pot  fishery,  1950-67:  Catch,  effort 
and  catch  per  unit  of  effort. 


81 


2 

-. 


2 
2 


3 


C 


< 
I 


s 

© 

o 

CT- 

I/O 

i/o 

00 

o 

o 

o 

c 

r*i 

ro 

r*N 

no 

00 

NO 

■<* 

NO 

■£  o. 

o 

i- 

cm 

VO 

^H 

CN 

>/->_ 

■rr 

or 

ts 

r~ 

^T 

NO 

NO 

fN 

r~J 

.  c 

cs 

c-i 

CI 

r— 

ON 

"* 

r-i 

r-4 

oo 

X       . 

NO 

r- 

NO 

NO 

NO 

00 

VO 

Tf 

CN 

03    O 

■— ' 

(N 

m 

iA 

o 

O 

no 

ON 

00 

o 

* 

r*". 

o 

J3 

r> 

NO 

o 

00 

NO 

r^-, 

l^) 

>■  — 

r-~ 

!-H 

CN 

vq 

ri 

CI 

1/1 

O 

23  s 
S  o 

fN 

I/O 

fsf 

ON 

r-* 

"*" 

u-> 

ON 

r-f 

ts 

r*l 

ON 

<N 

r- 

■* 

NO 

00 

o 

x: 

Oi 

[_ 

0 

tw 

3 

r- 

o 

NO 

00 

J^c 

uo 

oo 

u-> 

t~- 

03 

X 

W9 

<D 

O 

Z 

> 

> 

< 

> 

< 

> 

> 

CO 

00 

*— 

C— 

o 

r- 

<N 

n 

00 

o 

~^ 

ON 

i/o 

00 

~*_ 

ON 

wo 

c- 

Q 

rl 

ro 

ON 

■<* 

o 

m 

■* 

o 

BS 

lo 

ON 

"Of 

ON 

C~ 

ON 

lO 

ON 

r— 

1i  *• 

t 

3* 

I" 

to 

ro 

r- 

■* 

■ — ' 

ON 

tJ- 

n 

cs 

+ 

^_ 

+ 

oi 

,— « 

cs 

t^l 

_£- 

1 

CN 

X 

00 

X 

■-1 

■-H 

<j 

""*< 

X 

X 

♦J 

x 

X 

oo 

00 

r 

X 

oj 

+ 

>> 

>, 

I/O 

ON 

r 

x" 

1 

03 

X 

NO 

f» 

.' 

i— « 

' — ' 

E 

' — ' 

r) 

</0 

1 

— 

• — * 

* 

r- 

>/0 

o 

T_H 

NO 

% 

r- 

to 

o 

ON 

fN) 

ON 

o 

oi 

*— i 

OJ 

wo 

n 

■* 

cn 

lO 

uo 

m 

— 

(N 

rt 

NO 

00 

O 

03 

r 

w 

+ 

+ 

<~l 

r- 

00 

3 

X 

X 

CM 

X 

1 

1 

£ 

ro 

r- 

, 1 

i/o 

1 

j^* 

^-^ 

u 

II 

on 

ro 

00 

[ — 
oi 

X 

X 
00 

-X 

X 

X 
oo 

- 

1 

I/O 

1 

00 

1 

r^l 

I 

ON 

1 

no 

1 

CN 

1 

o 

X 

ON 

1 

** 

c— 

X 

o 

CN 

ON 

oo 

Tt- 

ON 

T 

cs 

tn 

•<* 

r-- 

ON 

I/") 

/ 

Tf 

oj 

r- 

on 

00 

l/^ 

+ 

r ) 

1    ' 

C*- 

o 

1/0 

^f 

r-i 

II 

ON 

oo 

II 

«* 

u") 

~~; 

CO 

X 

— ' 

^r 

,*? 

fN) 

II 

"3- 

II 

II 

X 

II 

X 

II 

^^" 

II 

^~ 

>v 

~5 

>.|X 

>^_x 

>. 

>> 

s^ 

>* 

— 

XI 

^ 

0* 

o 

u 

-3 

03 

tt5 

C/3 

03 

B< 

00 

Pi 

03 

0C 

X) 

00 
03 

oo 

as 

c 
3 

*0 

u 

u 

U 

U 

Q 

a 

a 

2 

r3 

-J 

-J 

-J 

-J 

J 

j 

—i 

_J 

c 

§  2 

a  e 


2  | 

>.    3 

-g   S 


•a 


c  > 
O    OJ 

Is 

^  c 

■e,a 


a  a 

-§   E 

O      w 


.SP  a 


E^ 


T3 
1> 


SB    S' 

2  SN 
oo  c  oa 

ii  ii    n 

00  *  — i 


.2P-2 


E  E 

^3    OJ 

,3  c 


c-     U 

^>   « 

<u  "" 

•£  — 

03  TO 

B   3 

rv  w5 
""  U 

o^    u 
C 

C    03 

■2  a 

03    > 

o 

2  s 

3X> 


K  o 


M  C    u 

°  °  5  u 


o 


u 


03 

3  2 

C    3 


3  ^~ 


u 


Q  S 

o  1 

*3  c 

4->  o.  > 

0  *'  o 

«°° 

^S  = 
C  ed  « 

II? 

QWQ 
<     D> 


T3 

U 

■d 

^ 

u 

— 

o 

0) 

a. 

0^ 
> 

0> 

C 

trj 

0 

03  T3 

■o  g 


c  o 

O  uo 

• -  ON 

c  — 
300 


XJ   - 

.2  S3 


2».a     - 


ii  ii  n  ii 

JU  Q  03 


82 


108.5 


106.7 


LDR 


0.00 


1;0.00      80.00 


120.00      160.00 


EFFORT:  HOOKS  (XLO  ) 


200.00 


1 

Uo.oo 


0.00 


80.00 


120.00 


160.00 


200.00 


EFFORT:  HOOKS  (XLO  ) 


Figure  7.  —  Atlantic  tuna  longline  fishery,   1956-67:   Catch,  effort, 
and  catch  per  unit  of  effort. 


83 


- 
a 

Taj 

a 
z 
- 


CO 
0! 


Q  -* 

r- 

t 

r~ 

in 

00 

o 

o 

M- 

c:  o 

~- 

c~t 

NO 

r~i 

fN| 

■"3- 

in 

CO 

^f 

V,   0 

IT) 

in 

00 

■* 

r— 

* 

vq 

Tf 

°.c 

00 

<N 

© 

(nJ 

■*" 

o 

ON 

cs 

(S 

S§ 

■* 

oo 

NO 

00 

NO 

■* 

00 

in 

ON 

OJ 

si 

~ 

j 

^1 

C/3    c 

cc 

vq 

r- 

r- 

r-i 

r- 

m 

r^ 

Tf 

o 

r^ 

CO 

oo" 

c-4 

r-i 

NO 

r- 

00 

NO 

CN 

o 

o 

i — i 

** 

o 

ON 

o 

ON 

~* 

*~ < 

— 

— ■! 

*— 

^^ 

•— 1 

o 

«g 

— i 

< 

C<1 

< 

"3 

vq 

Z 

r^, 

Z 

JS 

— 

fl 

o 

Z 

cs 

£ 

1/1 

On 

00 

r- 

C-- 

"9- 

rj 

r~ 

r- 

Q 

NO 

<N 

O 

m 

o 

"* 

r-4 

t*~ 

o 

as 

NO 

in 

c-) 

on 

■* 

00 

'O 

ON 

mj 

r- 

r- 

00 

r~ 

00 

00 

ON 

00 

ON 

r~) 

n 

"^i 

t 

r-4 

x 

n 

rl 

X 

>> 

X 

o 

in 

X 

X 

X 
in 

+ 

in 

IT) 

1 

in 

•3- 

*- 

~ 

in 

ON 

i 

1 

•* 

1 

X 

X 

V, 

o 

| 

CO 

+ 

X 

>>    1  X 

% 

in 
in 

00 

00 

o 

r- 

rn 

00 

NO 
ON 

00 

"rt 

O 

ci 

* 

00 
ON 

C*"l 

1 

1 

-' 

-Si 

ON 

X 

^_, 

1 

, — . 

1 

' 

.13 

m 

© 

^H 

X 

oo 

-w 

^ 

*-• 

^ 

^ 

a- 

ON 

in 

X 

**rf 

X 

4_i 

«^ 

c 

c^ 

r- 

*t 

ON 

X 

in 

^ 

X 

X 

w 

II 

1- 

o 

Ol 

n 

ON 

o 
m 

m 

X 

in 

mi 

<- 

es 

1 

no 

1 

00 
00 

1 

1 

i    3! 

1 

X 

^ 

x 

1 

X 

1 

C 

~ 

I 

X 

* 

X 

NO 
O] 

i£> 
r- 

CI 

ON 

lO 

oo 

o 

o 

r- 

ON 

+ 

00 
m 

^f 

•— ' 

vq 

ON 

rl 

r- 

NO 

NO          _ 

x 

—' 

rJ 

-" 

r-i 

<*> 

ri 

r-i    + 

II 

II 

II 

ii 

II 

II 

ii 

II 

>. 

>. 

>CTx 

>, 

>Ax. 

>> 

>CTx 

>, 

>,|X 

^_ 

^ 

as 

5 

_j 

oo 

■s. 

to 

C/} 

en 

'^ 

rz 

p 

S3 

_£< 

" 

Si 

n 

5 

(3 

■a 

0i 

a; 

OS 

a: 

as 

CC 

-7* 

^ 

_J 

_J 

_1 

_j 

_j 

_j 

_l 

j 

O 

o 

> 

c 

i/i 

- 

IM 

0 

n 

E 

£ 

■a  a 

9    E 


.a  £ 

o  jj 

£  -° 

3  ° 


C     «l    2 


^: 

ft 

1) 

c 

>. 
O 

i/) 

a 
■a 

E 

3 

^ 

^ 

o 

r 

0 

-,j 

ft 

-3 

0 
E 

E 

b 
o 

c 

£=5  S 


.2,tS   « 

to  '5 


2,«  a  as  > 


o  o 


OS  I  Q 


ii    n     ii 

C/N      H       — 


o  o 

JO! 


■o 

T3 


g  .a 


a  T3 
1.1 


_    £    t-  C  NO  ft^ 

—  C  3        o  m  x 

2  K  «,?  5  -  5 

Si  «.     C     J  .  -  — '  X 

•-i  e  '7  ~"  n 

■2  a  a  .=  it.  ii  'S  ■ 

~  "   °  —   rt  "  n 
'~t-  r-    •  ■    rz    i-    3^ 

ss-  <. 

II    II    5"  " 

-jit/1  -*— 


84 


Pi 


10.00 


8189 


si-  .  • 


60.00 


110.00 


160.00 


210.00  260.00 


EFFORT:    TAN  DAYS    (X105) 


10.00  60.00  110.00  160.00  210.00 

EFFORT:    TAN  DAYS    (X105) 


260.00 


Figure  8.  —  Soviet  and  Japanese  king  crab  fishery  in  the  Eastern 
Bering  Sea,  1955-67:  Catch,  effort,  and  catch  per  unit  of  effort. 


85 


IE 
v 

>. 


02 

e 


1) 

X 


X) 


T3 

C 


O 


— 


—  c  M 

fi 

on 

ON 

NO 

rN 

rN 

0 

00 

00 

"* 

i/-> 

r-j 

NO 

r- 

■* 

00 

0 

■<t 

CI 

1/1 

vn 

10 

rN 

m 

■•Oa 

■* 

■* 

*fr 

0 

•* 

00 

NO_ 

O 

00 

x  _  n 

CNJ 

«c2 

0 

■* 

•* 

■* 

u-> 

CI 

<♦ 

io 

S^ 

~ 

"" 

^H 

~ 

^ 

- 

™h 

v>  <r 

x>-o 

M  C 

vO 

ON 

00 

rN) 

fS 

■* 

NO 

00<_  3 

0 

On 

00 

tn 

t- 

ON 
ON 

00 

ON 

O 
O 

500 

ON 

00 

ro 

in 

CM 

00 

ON 

NO 

r~ 

ro 

m 

r-r 

lO 

SO 

90 

0*-1 

vD 

00' 

00* 

00' 

•rf 

r— 

00* 

t~ 

r^ 

Z.E 

^^ 

0 

— 

m 

< 

(N 

< 

^Ss 

1; 

Z 

I~- 

Z 

3 

£ 

0 

z 

01 

> 

O 

<^l 

ON 

** 

00 

■* 

0 

ON 

r<> 

OO 

00 

rsi 

ON 

NO 

<-h 

CNI 

"* 

«o 

a 

(N 

"* 

l/N, 

oj 

lO 

r» 

00 

ON 

r~ 

a: 

r» 

00 

00 

00 

00 

00 

00 

00 

00 

~H 

1 

*j 

>T 

X 

_ 

* 

r-l 

1 

CNl 

ro 

NO 

00 

NO 
■* 

00 

w 

rN 

00 

1 

x 

. — 1 

■  , 

rn 

CNl 

+ 

1 

2 

n 

X 
lO 

+ 

*x 

ON 

ed 

x 

£ 

<^l 

00 
CO 

X 

X 

NO 

1" 

ON 

l" 

x" 

r-i 

X 
in 

X 

00 

0 

00 

10 

l" 

*i 

QO 

rs 

Tj- 

^^ 

ON 

rH 

rs 

0 

+ 

00 

NO 

-d- 

' — ' 

m 

_ 

1 

r~ 

06 

m 

10 

r3 

r-1 

O 

N  <- 

1 

V> 

r^ 

r~ 

& 

Ol 

X 

X 

X 
r- 

X 

ci 

vn 

1 

1 

ro 

t1 

r 

.7^ 

PS 

nO 

m 

r-- 

on 

^^ 

X 

1 

X 

II 

f*"l 

("*N 

ON 

00 

10 

10 

rs 

ON 
NO 

On 

O 

1 

p 

X 

NO 

ON 

r 

X 

X* 

ON 

X 

^^ 

. 

1 

1 

WO 

tH 

1 

1 

x 

X 

NO 

X 

# 

CO 

NO 

00 

rs 

CNl 

ro 

1 

ON 
ON 

NO 

10 

00 

ci 

NO 

+ 

O 

0 

00 

t»* 

•* 

00 

ON 

CN 

Tfr 

r-4 

■* 

0 

(N 

w 

c*l 

r~i 

00 

m 

q 

Ti- 

NO 

r~ 

X 

rN) 

II 

— i 

II 

NO 

ll 

r^ 

II 

II 

II 

11 
>> 

+_, 

II 

^*~> 

^ 

'Z.' 

>. 

€~ 

>J>< 

>, 

>>|x 

2 

ai 

C 

0 

C/) 

C/N 

tSl 

W5 

~ 

n 

X) 

n 

X> 

a) 

JO 

a 

X) 

u 

■a 

OS 

tC 

OS 

OS 

OS 

OS 

ai 

Oi 

E 

O 

U 

U 

U 

a 

Q 

a 

Q 

— " 

*-* 

J 

_) 

_1 

_l 

_i 

-J 

_) 

—1 

o 

■S  cn| 

a  e 

o   E 
c    x 

I! 


*T3      o 


3     cfl 
^3    T3 


5  s 


o  t:  *a 
""  c  o 


3  — 

.§  1 

x  « 


e  g 


'J 


0i  = 


11  11  11  11 

c/3,  #— 1  rN 


E 

"3 

> 

a 

u 

13 

-0 

a 

id 

U 

M 

■^ 

U 

S3 

O    «j 

o  c 

■  —  n 

tc  c 

o  o 
-JO 


3  O  no 

2  §  o£ 

"5  3  u 

u  fl  i- 

i  —  <U    II 


J3  E 


11    11    11    11    11 


— J  U  Q    rs  X)  C/5 


86 


2bh-3 


1 1 1 1 

0.00  20.00  1(0.00  60.00  80.00 

EFFORT:   NUMBER  OF  SKATES   (XLO3) 


100.0 


^    \. 


0.00 


20.00 


1(0.00 


60.00 


80.00 


EFFORT:   NUMBER  OF  SKATES    (X10    ) 


100.00 


Figure   9.  — Cape  Flattery  sablefish  fishery,  1917-52:  Catch,  effort, 
and  catch  per  unit  of  effort. 


87 


4) 


O 


M 

—  .* 

S  o 

CJ  ■*" 

o 

m 

O 

ON 

o 

lO 

ro 
ro 

in 

NO 

NO 

ro 
in 

NO 

ro 
oo 

ro 

a  0 

ON 

r— 

NO 

ON 

•* 

ON 

ON 

NO 

ro 

ON 

O 

00 

ON 

**■ 

NO 

ro 

lO 

ro 

ON 

■* 

•"• 

lO 

V. 

-a 

s 

3 

C"~ 

NO 

ON 

r- 

ON 

■* 

>",? 

rh 

r- 

ro 

lO 

l> 

m 

NO 

r- 

&oa- 

■*r 

•* 

NO 

00 

NO 

m 

© 

m 

00 

S    3 

cm 

■* 

O 

r- 

00 

OO 

NO 

O 

o 

■a- 

(S 

NO 

in 

ro 

00 

NO 

o 

O 

rs 

cm 

cm 

CM 

CI 

CM 

CM 

£ 

H 

.* 

Cm 

ON 

lO 

lO 

o 

^<A 

•w 

"" 

o 

■* 

ca 

X 

Q 

z 

< 

<3 

> 

< 

> 

5S 

ro 

ON 

t— 

o 

t- 

co 

(—4 

r-~ 

vo 

vq 

00 

CM 

00 

q 

NO 

■* 

q 

a 

cm 

CM 

CM 

i/i 

t~ 

ON 

ON 

■* 

o 

r~ 

•<t 

no 

Cl 

m 

<* 

r- 

>n 

OO 

io 

00 

i 

>, 

ft 

>. 

X 

CO 

N-^ 

ON 

r-H 

ON 

NO 

CO 

00 

ts 

— ; 

— ; 

^H 

CM 

CM 

+ 

CM 

+ 

CM 

*j 

+-• 

X 

^^ 

•^"N 

X 

X 

O 

X 

o 

CM 

X 

CD 

+ 

„ 

>T  K 

>n 

l" 

X 

I/O 

WO 
ro 

ca 

X 

i 

1 

»-H 

• 

|" 

E 

1 

CM 

>> 

00 

CM 

•* 

CO 

^ 

!/5 

m 

ON 

CM 

lO 

w 

CM 

CM 

NO 

in 

00 

^H 

r~ 

"c3 

O 

ON 

1 

cm 

CM 

+ 

00 
CM 

+ 

ON 

ro 
m 

-1 

CM 

r~ 

CM 

ro 

00 
O 

'ft 

tN 

X 

X 
O 
oo 

NO 

00 

CM 

x" 

NO 

o 

^H 

1 

r^ 

1 

E 

ON 

q 

II 

NO 
NO 

ro 
•ti- 
ro 

X 

ON 

r~ 

NO 

o 

CM 

1 
X 

o 

m 

X 

o 

CM 

r 

X 

1 
X 

■JO 
lO 

X 
in 
ro 

f 

c^l 

1 

1 

1 

|" 

1 

X 

00 
ro 

00 
ON 

1 

X 

OO 

CM 

OO 

m 

ro 

tn 

o 

00 

o 

X 

+ 

00 

ro 

OO 
ON 

NO 
ON 
1^ 

O 
if 

00 

ro 

NO 

m 

1/0 

ON 
ON 

x 

_; 

II 

ON 

II 

CM 

II 

CM 

II 

— 

ii 

^ 

II 

^ *■' 

II 

^* 

II 

^~> 

>> 

>T 

>>|X 

>i 

>>|X 

>T 

>.|X 

>, 

>;|X 

_ 

X 

c 
0 

oi 

-1 

c/3 

on 

w 

V5 

eg 

X 

ca 

XI 

ca 

XI 

ca 

X 

o 

T3 

OS 

a 

at 

OS 

OS 

CA 

OS 

OS 

E 

U 

u 

U 

U 

Q 

Q 

Q 

Q 

— 

,£3 

J 

_1 

-j 

-J 

J 

-J 

J 

_) 

3 

w 

c  ■- 

.2  5 

*-    ca 


,0  (-4 

E* 
3  E 

11 

ca    x 

a    ca 

fc  E 

2    w 

a;    ca 


T3    u 

•3  ~ 


|s 

_-c   0 
cu  *-  ■  — 

^   ^     O     n 

.2"  S  § 

O   S  ^    h, 

Sao 

S  S  i  "- 

^    Cl  u    g 

■~  c  o  « 

«gaE 

flCr-     0 

E  ia-s 

g  XT  >>  <u 

£-£*>  "S 

t«.a  ^  CM 

3  «S  as 

o 

CL*     "^ 


T3  • J3 
ca  ca 


^  ts 


00 


o  o"  3 
tS  c  as  I 


c^  *. 


ca  "O 

C  => 
cu  cu 
u  ca 
B-S5 


-         c  a 

S      o  -a 

3       3  ca 


o  •£ 

o 

S    3 

3  X) 

"•a 


O 


o  g 

w5   ca 


ca 


r3 


W   cu 

ex: 
o  " 

CA     — 

«  ° 
^  s 

c  « 
S  E 

Uh    ■  — 

5  « 
< 


Q 
> 


u  5 

U    o. 

3    i_ 

2^ 

ca 

cu    C 

3-2 

71 

b 

H 

0 

O  ca 

o 

X 

X 
-3 

T3 

u 

X 

X 

.s^ 

mi 

0 

o 

cs 

Gfl     " 

>n 

O      - 

09 

3 

C  r~ 

B.£ 

XI    t 

3 

c 

_0 

ON 

0 

u 

— 1 

o  S 

c 

u 

C 

3 

in 

^e 

ac-w 

3 

u  S 

3 
Lm 

611 

ii 

i-  ■  — 

o  g 

U 

■a 

"5 

0 

73 

u 

> 

"3 

0 

ii  ii 

ii 

II 

II 

E 

n 

a:a: 

-J  o 

U 

« 

X  OO 

-i— 

relationship  (the  LDRaS  model  gave  a  larger 
R2,  but  y,_  i  was  not  statistically  significant). 
However,  the  LDRa  model  was  marginally 
significant  over  the  LCRa  model  (R2  of  0.88 
versus  0.85).  There  is  evidence  of  positive  auto- 
correlation for  the  LDRa  function.  The  Gulland- 
LCRb  does  give  somewhat  different  estimates  of 
y*  and  x*  than  unadjusted  LCRb.  Figure  8  shows 
the  LCRa  and  LDRa  functions. 


So  far,  we  think  that  the  logistic-decreasing- 
returns  function  has  considerable  merit.  It 
should,  of  course,  be  tested  further.  Other  func- 
tions should  also  be  tried,  including  those 
assuming  a  Gompertz  growth  function  and  the 
more  generalized  function  used  by  Tomlinson 
and  Pella.  It  is  also  hoped  that  this  effort  by 
economists  will  be  reviewed  by  people  in  the 
field  of  biology. 


Cape  Flattery  Sablefish 

Table  5  shows  the  results  for  the  Cape  Flattery 
sablefish  fishery.  Again,  the  LDRa  model  is 
superior  in  explaining  the  catch-effort  relation 
with  an  R2  of  0.54.  The  stock  adjustment  co- 
efficient was  not  statistically  significant  at  the 
5%  level.  Positive  autocorrelation  was  found  for 
the  LDRa  function.  There  does  not  seem  to  be  an 
appreciable  difference  between  the  Gulland- 
LCRb  and  the  unadjusted  LCRb. 

On  the  basis  of  the  sample  fisheries  it  would 
seem  that  the  LDRa  function  is  a  more  realistic 
description  of  the  catch-effort  relation  than  the 
LCRa  model  employed  by  Schaefer.  In  addition, 
it  is  apparent  that  for  the  above  species  the  Gul- 
land  method  of  adjusting  this  data  yields  very 
similar  results  to  the  unadjusted.  Catch-effort 
data  have  been  gathered  on  49  stocks  of  fish  by 
the  Economic  Research  Laboratory.  We  plan  to 
carry  out  similar  investigations  for  the  other 
stocks  since  the  basic  computer  programs  have 
been  written.  Figure  9  shows  the  LCRa  and 
LDRa  functions. 

CONCLUSIONS 

We  do  not  claim  to  have  discovered  the  "true" 
relation  between  effort  and  yield  for  the  stocks 
of  fish  discussed  in  this  paper.  We  have  no 
guarantee  either  that  biological  growth  is  exactly 
a  logistic  function,  or  that  yr  =  m,(l  —  z,x ) 
is  exactly  the  relation  of  effort  to  yield  from  a 
fixed  biomass.  But  we  believe  that  (1)  the 
decreasing-returns  functions  yt  =  m,  (1  —zt  t) 
is  theoretically  better  than  the  constant-returns 
function  y,  =  km,  employed  by  Schaefer;  and 
(2)  the  decreasing  returns  function  also  gives 
better  statistical  results  as  shown  graphically 
in  Figures  6  to  9  and  is  confirmed  by  the 
correlation  coefficients. 


LITERATURE  CITED 

BARNETT,  H.  J.,  and  C.  MORSE.  1963.  Scarcity  and 
Growth:  The  Economics  of  Natural  Resource  Avail- 
ability, Baltimore. 

BELL,  FREDERICK,  DARREL  NASH,  ERNEST 
CARLSON,  FREDERICK  WAUGH,  and  RICHARD 
KINOSHITA.  Manuscript.  The  Future  of  the  World's 
Fishery  Resources:  Forecasts  of  Demand,  Supply  and 
Prices  to  the  Year  2000  with  Recommendations  for 
Public  Policy,  U.  S.  Department  of  Commerce,  National 
Marine  Fisheries  Service,  Economic  Research  Laboratory. 

CARLSON,  ERNEST  W.  1969.  Bio-Economic  Model 
of  a  Fishery,  Economic  Research  Laboratory,  U.S. 
Department  of  Commerce.  Working  Paper  12. 

DAVIS,  HAROLD  T.  1941.  The  Theory  of  Econometrics, 
Bloomington,  Indiana:  Principia  Press,  Chapter  11. 

FISHER,  IRVING.  1925.  Our  Unstable  Dollar  and  the 
So-Called  Business  Cycle,  Journal  of  the  American 
Statistical  Association,  pp.  179-202. 

FOX,  WILLIAM  JR.  1970.  An  Exponential  Surplus- 
Yield  Model  for  Optimizing  Exploited  Fish  Populations, 
Transactions  of  the  American  Fisheries  Society,  No.  1. 

GULLAND,  J.  A.  Manual  of  Methods  for  Fish  Stock 
Assessment.  Part  1.  Fish  Population  Analysis.  FAO 
Manuals  in  Fisheries  Science  No.  4.  Rome. 

KOYCK,  L.  M.  1954.  Distributed  Lags  and  Investment 
Analysis,  Amsterdam,  pp.  9-14. 

MARSHALL,  ALFRED.  Principles  of  Economics,  Mac- 
millan,  8th  edition  1920  and  reprints  to  1930,  p.  166. 
London. 

NERLOVE,  MARC.  1958.  The  Dynamics  of  Supply.  The 
Johns  Hopkins  Press,  Baltimore. 

PELLA,  JEROME  J.,  and  PATRICK  K.  TOMLINSON. 
1969.  A  Generalized  Stock  Production  Model,  Inter- 
American  Tropical  Tuna  Commission,  Bulletin  13(3). 


89 


RYTHER.    JOHN    H.    1969.    Photosynthesis    and    Fish 
Production  from  the  Sea,  Science,  Vol.  166,  pp.  72-76. 


SCHAEFER,   MILNER    B.    1954.   Some  Aspects   of  the 
Dynamics  of  Populations  Important  to  the  Management 


of  Commercial  Marine  Fisheries,  Inter-American  Tropi- 
cal Tuna  Commission,  Bulletin  1(2),  27-56. 

SCHAEFER,  MILNER  B.  1956.  Some  Aspects  of  the 
Dynamics  of  Population  Important  to  the  Management 
of  Commercial  Marine  Fisheries,  Inter-American  Tropi- 
cal Tuna  Commission,  Bulletins  1  and  2. 


APPENDIX  I.  METHODS  OF  ADJUSTING  CATCH  AND 
EFFORT  DATA  TO  REPRESENT  EQUILIBRIUM  OBSERVATIONS 


The  Schaefer  (1957)  Method 

The  Schaefer  analysis  (using  his  notation)  is 
based  on  the  assumption  that  the  rate  of 
population  change  can  be  represented  by  the 
equation 

(l)^  =  klP(L-P)-k2FP 

where  k\  is  the  rate  of  population  increase, 
k2  is  the  catchability  coefficient,  L  the  maximum 
population  size,  F  is  fishing  effort,  and  P  is 
the  current  population  size.  Further,  it  is 
assumed  that  at  level  P,  in  year  i,  equilibrium 
yield,  Ye  is  estimated  by  P  +  Catch,  and  that 


(2)  AP 


Pt  +  1-  Pr- 


Ct*i/Ft+i  -Ct-i/Ft- 


where  C  is  catch.  To  use  these  equations  it  is 
necessary  to  relate  P  and  u,  catch  per  unit  effort, 
that  is 

(3)  P=k2u. 

If  P  in  equation  (1)  is  replaced  by  P,  then  all 
three  parameters  k,  k2,  and  L  can  be  estimated 
from  a  series  of  data  on  catch  and  catch  per  unit 
of  effort.  This  1957  procedure  of  Schaefer's  was 
first  tried  as  a  basis  for  a  decision  rule. 

Initially  a  15-year  series  of  data  was  divided 
into  three  equal  parts,  that  is,  1  to  5,  6  to  10,  and 
11  to  15  years.  The  three  parameters  were 
estimated  from  the  three  sets  of  data  by  solving 
the  simultaneous  equations  of  the  form 


nt 


11    ri 


A  Ui  =  k]     X 


r  "       —  ■ 

L  i   u] 
y=l    n. 


-2 


-k2    v  nt    Iil 

7=1    n, 

where  k\,  and  k2,  and  L  are  parameters,  A u,  is 
the  change  in  catch  per  unit  effort,  u,  is  the 
average  catch  per  unit  effort  u,~  is  the  average 
catch  per  unit  effort  squared,  f,  the  number 
of  units  of  effort  and  n,  the  length  of  the  period 
in  years. 

Pella  and  Tomlinson  suggested  that  the  series 
of  data  be  divided  into  periods  with  the  greatest 
differences  in  stock  levels  to  avoid  absurd  results. 
They  also  pointed  out  the  lack  of  a  unique 
solution,  since  different  partitioning  of  the  data 
may  give  different  results.  There  is  also  no 
statistical  basis  on  which  to  infer  properties  of 
the  parameters  such  as  bias,  consistency,  or 
efficiency,  etc. 


Gulland  (1961, 1968b)  Method 

This  method  involves  relating  the  mean 
annual  catch  per  unit  of  effort  in  a  given  year 
to  the  fishing  effort,  averaged  over  that  year 
and  a  certain  number  of  previous  years  cor- 
responding to  the  mean  number  of  years  that  a 
year-class  contributes  to  the  fishery. 

For  example,  the  catch  in  period  t  would  be 
related  to  the  average  effort  over  the  last  3  years 
for  the  yellowfin  tuna  since  a  year-class  con- 
tributes to  the  fishery  for  about  3  years.  We  are 
doubtful  of  the  validity  of  this  since  it  gives 
equal  weight  to  each  year  of  effort  in  computing 
the  average  effort.  We  feel  the  hypothesis  ex- 
pressed in  this  paper  is  more  realistic.  In  addi- 
tion,   the    statistical    properties    of   a    moving 


90 


average  of  effort  as  used  in  regression  are  not  in  the  Eastern  Tropical  Pacific  Ocean,"  Inter- 
well  known.  Finally,  the  technique  is  not  as  Amer.  Trop.  Tuna  Commission.  Bull.  2(6)  1957, 
direct  a  test  for  adjustment  of  the  population  pp.  245-285  and  Gulland,  M.  Manual  of  Methods 
to  effort  as  the  one  used  in  this  paper  (see  for  Fish  Stock  Assessments.  Part  1.  Fish 
below).  See  Schaefer,  M.  B.  "A  Study  of  the  Population  Analysis.  FAO  Fish  Technical 
Dynamics  of  the  Fishery  for  Yellowfin  Tuna  Paper,  1968,  FRs/T40  (Rev.  2),  97  pp. 


91 


Some  Suggestions  for  the  Development  of  a 
Bioeconomic  Theory  of  the  Fishery1 


Russell  G.  Thompson2 
ABSTRACT 

In  this  study,  the  fundamental  characteristics  of  the  Schaefer  model  and  the 
Thompson-George  (TG)  production-investment  model  are  reviewed,  and  extensions  of 
the  TG  model  are  discussed.  It  is  then  indicated  how  a  bioeconomic  model  for  the 
sole  ownership  fishery  may  be  obtained  by  adjoining  the  Schaefer  model  to  the  TG 
model  (or  any  of  the  extensions).  This  leads  into  a  discussion  of  the  fundamental  variables 
in  a  dynamic  analysis  of  the  fishery  problem  and  the  limitations  of  published  bioeconomic 
analyses.  It  is  further  pointed  out  that  further  work  needs  to  be  directed  to  the 
formulation  of  catch  functions  allowing  for  varying  marginal  returns  with  respect  to 
fishing  effort,  in  particular. 


INTRODUCTION 


In  1954  Schaefer  used  the  first-order  terms  of 
the  sigmoid  growth  law  to  describe  the  dynamics 
of  an  unexploited  fish  population  and  assumed 
the  catch  to  be  proportional  to  effort3  to  describe 
the  exploitation  by  man.  The  catch  function  was 
subtracted  from  the  natural  growth  law  to 
obtain  the  following  model  (which  is  commonly 
referred  to  as  Schaefer's  model): 

(1)  x(t)  =  rx(t)  (v—x(t))  -  ay(t)x(t) 

where  x  is  the  fish  biomass,  y  is  fishing  effort, 
t  is  time,  x(t)  =  dx(t)ldt,  and  the  remaining 
symbols  are  parameters. 

In  1968  Thompson  and  George  formulated  a 
production-investment  model  for  the  firm  in- 
volving stocks  and  flows.  Less  than  full  use  of 
the  capacity  was  allowed  for  by  introduction  of 
a  production  scale  variable.  Short-  and  long-run 
distinctions  in  economics  were  thus  possible. 
The  firm  could  increase  the  capital  stock  by  the 


1  Partially  supported  by  the  National  Science  Founda- 
tion as  a  part  of  the  Sea  Grant  Program  for  1970. 

2  Russell  G.  Thompson  is  Professior  of  Quantitative 
Management  Science,  University  of  Houston. 

3  As  indicated  by  Schaefer  and  Beverton  (1963),  this 
assumption  is  common  to  the  Beverton-Holt  approach 
as  well. 


purchase  of  capacity  in  excess  of  attrition.  None 
of  the  capital  stock  could  be  sold  within  the 
decision  interval  of  finite  length;  it  could  only 
be  sold  at  the  end  of  the  interval.  Therefore,  the 
problem  was  irreversible  during  the  finite  period. 
Extensions  to  allow  for  increasing  marginal 
costs  are  straightforward  and  were  left  to  the 
reader.  The  decision  rules  for  the  optimal 
production  and  investment  controls  were  derived 
by  use  of  control  theory  methods.  An  algorithm 
was  developed  by  which  to  compute  solutions  to 
the  controls  so  that  the  model  had  practical  as 
well  as  theoretical  value. 

In  1970  George  showed  that  solutions  to 
the  optimal  controls  for  a  cash  flow  form  of 
analysis  (as  used  by  Thompson  and  George)  were 
identical  to  those  for  a  discounted  form  of 
analysis.  That  is,  in  reference  to  the  TG  model, 
the  optimal  controls  are  the  same  for  the  case 
where b(t)>o and  D(t)  =  o  as  for  the  case  b(t)  =  o 
and  D(t)  is  evaluated  at  the  market  rate  of 
interest  i(t).  George  further  showed  that  one 
model  or  the  other  must  be  used  (in  an  exclusive 
sense). 

In  1971  Thompson,  Hocking,  and  George 
showed  how  the  initial  values  for  the  physical 
and  money  capital  accounts  can  be  derived 
optimally  as  a  part  of  the  solution  to  the 
investment-production  problem  (as  well  as  the 
values  for  the  controls  during  the  decision- 
making period).  In  1970  Proctor  studied  the 
investment  problem  for  the  firm  in  a  reversible 
and  also  in  an  irreversible  setting  (where  the 


92 


firm  may  buy  and  sell  its  capital  stock  during 
the  period  as  well  as  at  the  end).  He  further 
derived  the  demand  functions  for  capital  in  each 
case  and  deduced  their  economic  characteristics. 


CONCEPTUAL  MODIFICATIONS 

By  adjoining  the  Schaefer  model  to  any  one 
of  these  formulations,  a  production-investment 
model  for  the  sole  ownership  fishery  is  obtained.4 
Such  a  formulation  has  a  number  of  distinct 
advantages:  First,  the  inherently  dynamic  prob- 
lem of  the  fishery  is  formulated  accordingly  in  a 
mathematical  sense,  second,  the  model  (since  it 
encompasses  the  economic  and  biological  rela- 
tions) is  bioeconomic  in  form;  third,  given  mean- 
ingful expressions  for  the  functions  involved, 
decision  rules  for  the  production  and  investment 
controls  (and  hence  the  basis  for  a  bioeconomic 
theory)  may  be  derived  by  the  straightforward 
use  of  published  mathematical  methods. 

Lack  of  such  a  methodology  may  be  the  reason 
for  the  historical  development  of  the  bio- 
economic theory  for  the  fishery.  For  example, 
virtually  all  economists  who  have  published 
in  the  professional  journals  (or  by  the  way  of 
Resources  for  the  Future)  have  commonly  as- 
sumed the  inherently  dynamic  problem  of  the 
fishery  to  be  static  at  the  outset  of  their 
analyses  (cf.  Smith  1969),  Christy  and  Scott 
(1965),  Gordon  (1954),  and  Crutchfield  and 
Pontecorvo  (1968). 

Another  example  is  provided  by  the  form  of 
the  catch  function  used.  Until  recently,  econo- 
mists have  not  seriously  questioned  the  form  of 
the  catch  function  introduced  by  Schaefer,  oyx. 
This  formulation  implies  constant  marginal 
returns  with  respect  (w.r.)  to  effort  and  in- 
creasing returns  to  scale. 

Crutchfield  and  Zellner  (1962)  made  static 
and  dynamic  analyses  of  the  fishery  problem 
(with  this  catch  function)  and  found  different 
constant  solutions!  They  failed  to  note  that  a 
capacity  limitation  must  be  imposed  on  fishing 
effort.  The  problem  is  similar  to  maximizing 
the  function  y  =  x  in  which  the  domain  must  be 

4  Any  of  these  forms  of  the  problem  are  consistent 
with  Turvey's  formulation  (1964).  Variations  in  mesh 
size  would  be  associated  with  different  capital  character- 
istics, and  require  the  introduction  of  more  than  one 
capacity  variable  and  possibly  functions  relating  vessel 
types  and  mesh  size. 


bounded  from  above  for  the  problem  to  have 
finite  solution. 

Following  this  analysis,  Crutchfield  and  Zell- 
ner introduced  a  Cobb-Douglas  form  for  the 
catch  function  and  made  a  partial  analysis  of 
this  case.  This  problem  also  requires  a  capacity 
limitation  on  effort  to  be  well  posed.  In  addition, 
increasing  returns  to  scale  in  capacity  for 
sufficiently  small  expenditures  may  be  neces- 
sary as  well  as  decreasing  returns  beyond  some 
point.  This  is  particularly  relevant  when  the 
competitive  model  is  desired  for  a  reference 
framework.  Decreasing  returns  everywhere  are 
inconsistent  with  the  market  requirements  for 
a  competitive  structure  (Proctor,  1970). 

Still  another  example  of  the  unusual  approach 
used  to  date  is  the  specification  of  an  infinite 
horizon  for  the  completely  irreversible  invest- 
ment problem.  The  optimal  length  of  the  horizon 
in  a  common  property  resource  problem  might 
well  be  one  of  the  fundamental  results  being 
sought  in  the  analysis,  and  not  an  input  to  the 
analysis,  as  specified  by  Crutchfield  and  Zellner. 
There  are  no  transferable  rights  to  the  fishery 
resource;  and  hence,  the  entrepreneur  might 
desire  to  take  all  of  the  resource  within  a  finite 
period  of  time.  Thus,  the  optimal  solutions  to 
the  investment  and  production  controls  and  the 
length  of  the  decision  horizon  would  be  expected 
to  be  the  fundamental  variables  for  a  bioeconomic 
theory  of  the  fishery. 

For  the  case  of  the  Schaefer  model,  the  decision 
rules  for  the  production-investment  controls 
follow  immediately  from  the  TG  model.  The 
necessary  condition  for  the  optimal  length  of 
the  decision  interval,  if  one  exists,  follows  as 
an  immediate  extension  of  their  results.  In 
fact,  the  decision  rules  for  investment  and 
production  are  particularly  straightforward  and 
easy  to  state.  Let  v  —  investment,  m  =  the 
investment  upper-bound,  7  =  the  fish  price, 
6  =  production  cost  per  unit  of  effort,  f  =  in- 
vestment cost  per  unit  of  capacity,  <P  =  the 
discount  function,  z  =  fishing  capacity,  p\  — 
the  marginal  value  of  the  fish  per  unit  weight, 
and  p2  =  the  marginal  value  of  capacity.  Then 
the  decision  rules  are: 


(2  =0iip:  -0f  <0, 

m  if  p:  —  0f  >0, 


93 


(3)    v,, 


0  if  0ioxo  —  pi  ox,,  —  <f>0  <0, 
z,t  if6yox0  —pi  axo  —  <pO  >0. 


with  the  subscript  on  v,  y,  x  and  z  denoting 
optimum  values. 


method  may  be  further  enhanced  considerably 
by  the  development  and  estimation  of  more 
robust  forms  of  the  catch  function. 


The  sole  owner  firm  invests  the  maximum  pos- 
sible amount  if  the  marginal  value  of  capacity 
is  greater  than  the  discounted  marginal  cost  of 
capacity  and  does  not  invest  at  all  if  the  opposite 
is  the  case.  The  firm  uses  all  of  its  capacity  if 
the  discounted  net  marginal  revenues  from 
fishing  effort,  <p( ■yox,,—  0),  are  greater  than  the 
marginal  value  of  the  fish  resource,  p/  ox,,,  and 
the  firm  does  not  fish  at  all  if  tht  marginal  value 
of  the  fish  resource  is  greater  than  the  net 
marginal  revenue  from  fishing. 

The  difference  between  a  sole  owner  firm  and 
a  competitive  firm  is  immediate.  In  the  latter 
case,  the  effects  of  fishing  on  the  resource  are 
ignored;  and  hence,  the  marginal  value  of  the 
fish  resource  is  always  zero  (since  px  (t)  =  o).  It 
can  further  be  shown  that  pi^(t)  for  all  £.  Thus, 
the  marginal  value  of  the  fish  resource  reduces 
the  value  of  the  decision  rule  for  fishing  effort. 

If  the  Schaefer  model  is  augmented  to  allow 
for  a  Cobb-Douglas  type  of  catch  function,  for 
example,  then  an  interior  solution  (in  the 
interval  [o,  z0])  for  fishing  effort  is  possible. 
Similarly,  an  interior  solution  (in  the  interval 
[o,  m])  for  investment  costs  is  possible  if 
increasing  marginal  costs  of  capacity  are 
specified. 

The  main  difficulty  in  applying  the  TG  model 
(as  first  developed)  is  specification  of  the  invest- 
ment upper-bound.  It  is  clearly  a  proxy  for 
various  limitations  on  investment.  For  instance, 
there  might  be  borrowing  limitations  imposed 
by  the  financial  community.  If  so,  Rahman's 
extension  (1970)  of  the  TG  model  may  be  ap- 
propriate. On  the  other  hand,  the  investment 
upper-bound  may  be  superfluous  if  the  catch 
function  is  of  a  traditional  production  function 
form.  Few  serious  efforts  have  been  directed  to 
investigations  of  alternative  forms  for  the  catch 
function.  Further  efforts  of  the  type  being 
pursued  by  Carlson  (1969)  surely  need  to  be 
given  top  priority  in  fishery  research. 

In  summary,  an  operational  methodology  for 
the  management  of  a  fishery  is  available  by 
adjoining  the  Schaefer  model  to  the  TG  model, 
or  one  of  its  extensions.  The  potential  for  this 


LITERATURE  CITED 

CARLSON,  ERNEST  W.  1969.  A  Bio-economic  Model 
of  a  Fishery.  Working  Paper  No.  12,  Division  of 
Economic  Research,  National  Marine  Fisheries  Service, 
U.S.  Department  of  Commerce. 

CHRISTY,  F.  T.,  JR.,  and  A.  SCOTT.  1965.  The  Common 
Wealth  in  Ocean  Fisheries.  Published  for  Resources 
for  the  Future,  Inc.,  by  the  Johns  Hopkins  Press, 
Baltimore,  281  pp. 

CRUTCHFIELD,  J.  A.,  and  G.  PONTECORVO.  1969. 
The  Pacific  Salmon  Fisheries,  A  Study  of  Irrational 
Conservation.  Published  for  Resources  for  the  Future, 
Inc.,  by  the  Johns  Hopkins  Press,  Baltimore. 

CRUTCHFIELD,  J.  A.,  and  A.  ZELLNER.  1962. 
Economic  Aspects  of  the  Pacific  Halibut  Industry. 
Fishery  Industrial  Research.  Vol.  1,  No.  1. 

GEORGE,  M.  D.  1970.  Discounting  and  Cash  Flow 
Analysis  in  Investment  Problems.  Unpublished  manu- 
script available  from  author  on  request. 

GORDON,  H.  S.  1954.  The  Economic  Theory  of  a 
Common  Property  Resource:  The  Fishery.  Journal  of 
Political  Economy,  62(2):  124-142. 

PROCTOR,  M.  S.  1970.  Investment  Theory  for  the 
Firm:  Deterministic  and  Stochastic  Models.  Unpub- 
lished Ph.D.  dissertation,  Texas  A&M  University. 

RAHMAN,  QUAZI  MD.  MAFIZUR.  1970.  An  Optimal 
Investment  and  Financial  Control  Model:  Theoretical 
Solutions  and  an  Application.  Unpublished  Ph.D. 
dissertation,  Texas  A&M  University. 

SCHAEFER,  M.  B.  1954.  Some  Aspects  of  the  Dynamics 
of  Populations  Important  to  the  Management  of  the 
Commercial  Marine  Fisheries.  Inter-American  Tropical 
Tuna  Commission,  Bulletin,  1(2):  26-56,  La  Jolla, 
California. 

SCHAEFER,  M.  B.,  and  R.  J.  H.  BEVERTON.  1963. 
Fishing  Dynamics  —  Their  Analysis  and  Interpretation. 
In  M.  N.  Hill  (editor)  The  Sea,  pp.  464-483.  Interscience, 
Vol.  2,  New  York. 

SMITH,  V.  L.  1969.  On  Models  of  Commercial  Fishing. 
Journal  of  Political  Economy.  77(2):  181-198. 

THOMPSON,  R.  G.,  and  M.  D.  GEORGE.  1968.  Optimal 


94 


Operations  and  Investments  of  the  Firm.  Management  767-772. 
Science,  15(1):  49-56. 

THOMPSON,    RUSSELL    G.,    R.    R.    HOCKING,    and  TURVEY,   R.    1964.   Optimization   and   Suboptimization 

MELVIN    D.    GEORGE.    1971.    A    Nonconvex    Control  in    Fishery    Regulation.    American    Economic    Review, 

Problem  for  the  Competitive  Firm.  Econometrica.  39(5):  54(2, 1):  64-70. 


95 


Practical  Problems  of  Constructing  Bioeconomic 
Models  for  Fishery  Management 


Paul  Adam1 

ABSTRACT 

In  many  practical  cases  it  is  impossible  to  construct  a  complete  bioeconomic  model 
of  a  given  fish  stock,  such  as  when  one  or  several  fleets  move  irregularly  from  one  stock 
to  another,  or  when  fishing  effort  increases  so  rapidly  that  it  is  not  possible  to 
accurately  specify  a  reliable  yield/effort  relationship.  A  continuing  bioeconomic  model 
is  proposed  here  which  will  allow  inclusion  of  these  dimensions  while  allowing  both 
for  year-to-year  fluctuations  in  managed  effort  and  also  for  gradual  adjustment  of  labor 
and  capital  to  those  levels  designated  as  optimal  within  the  broad  ranges  of  this 
continuing  model.  Year-to-year  re-evaluation  offish  stocks  and  capital-labor  requirements 
is  stressed. 


INTRODUCTION 

This  paper  is  devoted  to  the  problem  of  mixed 
fisheries.  Few  fish  stocks  are  exploited  by  one 
fishing  fleet  only  and  few  fishing  fleets  are 
dependent  upon  only  one  fish  stock.  In  the  rare 
cases  of  isolated  fisheries  (one  main  species, 
one  fleet,  one  market)  there  are  often  incidental 
catches  which,  although  they  may  be  relatively 
small,  are  important  for  the  overall  profitability 
of  the  fleet.  It  can  be  said  that  in  most  fisheries 
the  rule  is  to  switch  from  one  type  of  fishing  to 
another  or  from  one  stock  to  another,  according 
to  the  seasons  or  to  the  variable  fish  abundance 
in  the  different  stocks.  These  continuous  adjust- 
ments, occurring  irregularly,  make  the  problem 
of  fishery  management  a  most  complex  one. 

Furthermore,  it  must  be  added  that  in  the 
last  10-15  years  the  techniques  used  in  some  of 
the  most  important  world  fisheries  have  been 
considerably  improved.  These  developments 
include:  long  distance  stern  trawling  associated 
with  freezing  at  sea,  purse  seining  for  pelagic 
species  in  the  North  Atlantic,  purse  seining  for 
tuna  species  in  the  Central  Pacific  and  Atlantic, 
double  beam  trawling  in  the  North  Sea,  etc.  As 
a  consequence  of  these  recent  developments,  it 
is  more  difficult  to  study  those  fisheries  which 


1  Head  of  the  Fisheries  Division,  Organization  for 
Economic  Cooperation  and  Development.  The  author  is 
solely  responsible  for  the  ideas  and  information  presented 
in  this  paper. 


are  the  most  advanced  and  consequently  the 
most  interesting. 

The  study  made  in  this  paper  will  obviously 
be  economic,  but  no  serious  or  complete  eco- 
nomic study  of  any  fishery  can  be  undertaken 
without  consideration  of  the  available  resources. 
In  other  words,  the  work  of  the  economist  in  this 
context  cannot  begin  or  would  have  no  solid 
basis  without  starting  with  the  findings  of 
marine  biologists.  It  is  therefore  indispensable 
to  examine  the  nature,  the  scope  and  especially 
the  shortcomings  of  the  biological  findings  inas- 
much as  they  have  to  be  used  by  the  fishery 
economists. 

SHORTCOMINGS  OF  THE 
BIOLOGICAL  MODELS 

The  whole  process  of  the  fishing  operations 
is  expressed  in  Figure  1.  The  arrows  indicate 
the  basic  components  of  an  operating  fishery. 
It  makes  it  apparent  that  any  research  which 
would  isolate  either  biology  or  economics  would 
be  cut  off  from  the  feedback  occurring  in  reality. 
Any  model  used  to  describe  reality  will  be  false 
if  it  is  divided  into  two  isolated  parts. 

The  traditional  catch  curve  derived  from  the 
biological  findings  on  one  fish  stock  cannot  be 
directly  used  by  the  economists.  In  fact,  this 
curve,  which  is  an  average  catch  curve,  should 
be  supplemented  with  two  curves  indicating  the 
maximum  and  minimum  yields  according  to  the 


96 


BIOLOGICAL  RESEARCH 


ECONOMICAL  RESEARCH 


Fishing  Effort 
(vessels  +  gear) 
expressed  in 
physical  terms 


Cost  of  Fishing  Effort 
+ 
Profit 


Fish  stocks 


Catches 
(in  weight  of  fish 

for  each  stock) 


Returns 
from  the  sales 
of  the  landings 


Figure  1.  —  The  basic  components  of  the  fishing  process. 


fluctuations  of  abundance.  As  shown  by  Figure 
2,  these  curves  of  maximum  and  minimum  yields 
accentuate  departure  from  MSY  as  compared 
to  the  average  curve  with  increasing  fishing 
effort  (and,  after  the  point  of  MSY,  increasing 
overfishing).  The  reason  is  that  the  more  in- 
tensive is  the  fishing  effort,  the  faster  the  year 
classes  are  exhausted,  as  there  are  often  rather 
wide  fluctuations  in  the  strength  of  the  succes- 
sive year  classes.  The  fluctuations  of  the  catches 
can  only  be  increased  with  a  faster  exhaustion 
of  the  best  year  classes. 

As  shown  by  Figure  2,  it  is  difficult  to 
evaluate  the  social  cost  of  fishing  effort  unless 
we  have  the  simple  case  of  a  given  fleet  exploit- 
ing a  given  fish  stock.  In  such  a  case,  the  losses 


of  years  of  bad  catches  are  compensated  by  the 
profits  made  in  better  years.  Or,  if  the  market 
for  the  landings  is  also  isolated,  it  might  be  that 
the  returns  are  more  or  less  equalized  by  higher 
prices  when  there  is  a  scarcity  in  landings  and 
lower  prices  when  the  landings  are  more 
abundant. 

For  mixed  fisheries,  Figure  2  should  be 
transformed  into  Figure  3,  thereby  taking  into 
account  the  fact  that  the  fishing  fleet  exploiting 
a  given  stock  at  a  given  average  level  is  maxi- 
mum when  the  abundance  in  the  given  stock  is 
maximum  and  when  the  abundance  in  the  other 
stocks  that  can  be  fished  by  the  same  fleet  is 
minimum,  and  vice  versa.  No  stock  can  be 
subject  to  a  stable  fishing  effort.  It  will  vary 


97 


Catches 


Average  cost  curve 


Average  yield  curve 


Fishing  effort 


Figure  2.  —  Maximum,  average,  and  minimum  catch  curves  for  a  single  fish  stock. 


Average  yield  curve 


Fishing  effort 

Figure  3.  —  Maximum,  average,  and  minimum  catch  curves  for  a  multiple  stock  fishery. 


between  two  extremes  determined  by  the 
abundance  in  the  stock  considered  but  which  also 
depend  upon  the  abundance  in  the  neighboring 
stocks.  It  should  be  noted  that  the  shape  of  the 
resulting  curve  and  the  location  of  the  point  of 
equilibrium  would  have  to  be  determined  for 


each  particular  case.  Each  case  would  not  only 
be  the  result  of  the  structure  of  the  given  fish 
stock  and  of  the  exploitation  borne  by  this 
stock,  it  would  also  be  the  result  of  the  structure 
of  the  other  stocks  which  would  be  more  or  less 
attractive,  i.e.,  profitable.  The  findings  of  the 


98 


biologists  should  therefore  cover  all  the  stocks 
which  are  exploited  by  the  fleets  considered  by 
the  economist,  otherwise  there  would  be  a 
substantial  gap  in  an  essential  part  of  the 
needed  information. 

The  previous  paragraphs  were  based  on  the 
assumption  that  the  pattern  of  the  recruitment 
to  the  fish  stocks  remains  unchanged  whatever 
the  size  of  the  stock  and  the  level  of  the  fishing 
effort.  In  practice,  this  assumption  is  certainly 
not  realistic.  But  the  opposite  assumption  that 
the  level  of  recruitment  is  linked  solely  to  the 
size  of  the  stock  is  certainly  equally  erroneous. 

These  two  remarks  oblige  us  to  enter  some- 
what into  the  intricacies  of  the  computations 
made  by  the  marine  biologists.  When  these 
scientists  are  examining  the  past  catches  they 
proceed  along  analytical  lines  which  are  cor- 
rected every  year  according  to  what  has  hap- 
pened. Their  analyses  are  summarized  and 
systematized  with  the  help  of  mathematical 
functions.  These  functions  can  serve  the  addi- 
tional purpose  of  making  forecasts  about  the 
effect  of  a  diminishing,  sustained,  or  increased 
fishing  effort  in  the  years  to  come,  ceteris 
paribus. 

Among  these  other  factors  the  main  one  is 
the  pattern  of  recruitment.  When  a  constant 
rate  of  recruitment  is  assumed,  the  mathematics 
lead  to  a  curve  tending  asymptotically  to  a 
minimum  yield  equal  to  an  exploitation  level 
associated  with  average  yearly  recruitment. 
When  recruitment  is  assumed  to  be  aligned  with 
the  size  of  the  stock,  mathematics  lead  to  a 
curve  asymptotic  to  the  X  axis  or  to  a  parabola. 
In  fact,  both  assumptions  are  false  and  known  to 
be  false;  the  real  curve  for  each  stock  is  in 
between  these  two  different  mathematical 
formulations,  but  present  scientific  knowledge 
in  marine  biology  does  not  allow  us  to  know 
when  the  pattern  of  recruitment  becomes 
different. 

The  resulting  margin  of  error  is  of  course 
without  practical  importance  when  there  is  a 
stable  fishing  effort.  When  the  increase  of  fishing 
effort  is  slow,  the  impact  can  be  surveyed  step 
by  step  and  the  margin  of  error  remains  small. 
But  when  the  increase  of  fishing  effort  is  fast 
and  furthermore  when  fishing  effort  is,  as  is 
true  in  complex  fisheries,  significantly  varying 
from  one  year  to  the  other,  the  margin  of  error 
is  bound  to  be  as  large  as  the  distance  between 


the  two  curves.  This  precludes  an  accurate 
forecast.  In  any  case,  it  seems  that  most  often 
the  yield  curve  is  relatively  flat  around  the 
maximum.  The  Schaefer  model  tends  to  exag- 
gerate the  sharpness  of  the  turning  point  at 
MSY,  whereas  the  Beverton  and  Holt  model 
may  tend  to  exaggerate  the  flatness  after  MSY. 
Let  us  imagine  a  fish  stock  exploited  as  in 
Figure  4  at  a  variable  level  of  fishing  effort, 
with  fluctuations  stabilized  at  maximum  and 
minimum  levels  unchanged  for  a  number  of 
years.  The  calculations  of  the  biologists  lead 
to  a  derivation  of  a  yield  curve  as  drawn  in 
Figure  4.  The  margins  of  error  in  the  calculations 
are  such  that,  if  there  were  a  change  in  recruit- 
ment function  around  the  point  of  average  yield, 
it  could  not  be  easily  seen;  the  actual  average 
yield  curve  could  well  be  drawn  by  the  dotted 
lines  and  no  one  could  prove  which  is  the  real 
one.  This  is  not  critical  if  the  fishing  effort  is 
not  increased,  but  assuming,  as  it  is  often  the 
case  at  present,  an  increasing  demand  for 
protein  and  improved  productivity  due  to  tech- 
nological change,  the  only  practical  problem 
would  be  the  problem  of  an  increased  fishing 
effort  .  .  .  for  which,  with  such  data,  no  forecast 
at  all  could  be  made  before  a  new  stabilization 
of  fishing  effort  for  a  subsequent  number  of 
years.  Before  such  a  stabilization,  the  most 
detrimental  consequences  could  have  materi- 
alized (cf.  the  California  sardines).  The  faster 
the  increase  in  fishing  effort,  the  more  difficult 
are  the  assessments. 


PARTIAL  BIOECONOMIC  MODELS 

While  initially  I  attempted  to  prove  that 
biological  models  cannot  be  complete,  at  least  in 
the  most  important  cases  of  increasing  fishing 
effort,  it  is  not  necessary  to  stress  that  com- 
plete bioeconomic  models  cannot  exist.  It  is  an 
obvious  fact  that  in  bioeconomic  models  biology 
comes  first;  they  are  fully  dependent  on  the 
reliability  of  the  basic  biological  data.  This  is 
a  very  big  drawback  which  would  well  render 
the  whole  exercise  of  very  little  practical  help 
in  managing  fish  resources.  But  it  should  not  be 
forgotten  that,  in  most  cases,  the  biologists  can, 
with  reasonable  accuracy,  indicate  the  level  of 
maximum  sustainable  yields.  This  limit  gives 
a  very  important  and  solid  basis  for  assessment 


99 


M.S.Y. 


Fishing  effort 


Figure  4.  —  Alternative  yield  curves  for  a  fish  stock  exploited  at  variable  levels 

of  fishing  effort. 


as  such  a  limit  cannot  be  overstepped  without 
economic  losses. 

It  could  also  be  added  that  the  impossibility 
of  constructing  complete  bioeconomic  models  is 
not  as  harmful  as  might  be  thought.  In  many 
cases  of  advanced  overfishing,  complete  bio- 
economic models  would  not  necessarily  supply 
practical  management  policies.  In  a  situation 
of  advanced  fishing  effort,  the  benefits  to  be 
expected  from  fishery  management  are  benefits 
which  could  not  be  reaped  before  the  stock  is 
rebuilt  to  its  MSY  level.  In  the  meantime  the 
reductions  likely  to  be  made  in  fishing  effort 
would  cause  problems  of  de-investments  (e.g., 
scrapping  premiums  .  .  .)  and  of  employment 
(re-employment  of  the  fishermen  concerned). 
Furthermore,  a  reduced  and  less  costly  fishing 
effort  exploiting  a  rebuilt  stock  would  give 
rents;  it  is  possible  to  imagine  regulatory  means 
by  which  such  rents  would  be  at  least  partly 
taken  from  the  remaining  fishermen,  but  this 
could  only  be  made  on  the  basis  of  the  fishing 
techniques  prevalent  at  the  time  of  making  the 
regulation.  It  would  often  be  difficult  to  find 
the  regulations  which  would  result  in  the  desir- 
able aggregate  effort  while  permitting  new 
technological  developments  at  the  same  time. 
Some  success  has  been  achieved  in  the  Canadian 


salmon  program  toward  attaining  both  of  these 
ends.  In  other  words,  even  if  complete  bio- 
economic models  would  exist  they  would  not 
as  such  provide  complete  solutions  to  the 
problems  of  re-establishing  overfished  stocks  to 
the  ideal  situation  of  MSY. 

Before  going  further  it  is  necessary  to  say  a 
few  words  about  the  techniques  of  communica- 
tion between  biologists  and  economists.  In  fact, 
there  is  not  much  difficulty  with  the  basic 
Schaefer  model  which  is  widely  used  in  the 
United  States.  Its  mathematical  expression  is 
as  follows: 


(1)    Y  =  aE  +  bE^ 


where 


Y     = 
E     = 

a,b  = 


yields,  expressed  in  weight  of 

catches 

fishing    effort,    expressed    in 

number  of  given  vessels  during 

given  times 

parameters   characterizing 

each  particular  stock. 


The  economist  has  little  difficulty  in  following 
and  utilizing  biological  results  from  this  model. 


100 


Unfortunately  the  Beverton  and  Holt  model 
is  not  so  easy  to  handle.  In  its  simplest  expres- 
sion, it  reads: 

(2)  Z  =  M  +  F  = 

log(,  number  of  fish  at  beginning  of  year 
number  of  fish  at  end  of  year 

where 

Z  —  total  mortality 

M  =  natural  mortality 

F  =  fishing  mortality. 

If  calculated  on  a  weight  basis  instead  of  a 
number  basis,  account  should  also  be  taken  of 
the  rate  of  growth  of  live  fish. 

With  such  a  model  converting  the  figures  of 
the  biologists  into  units  which  can  be  utilized 
by  the  economists  is  most  often  impossible. 
No  mathematical  barrier  exists  as  long  as  it  is 
understood  that  the  natural  logarithm  of  a 
ratio  between  the  catches  or  the  stocks  of  two 
years  is,  in  fact,  a  percentage.  However,  an 
important  part  of  the  data  utilized  by  the 
biologists,  when  it  is  all  published,  is  scattered 
in  many  different  publications.  It  is  not  suf- 
ficient to  know  the  ratio  of  abundance  derived 
from  fishing  effort  (F)  and  the  ratio  of  natural 
mortality  (M);  the  ratio  of  the  growth  of  the 
fish  and  the  assumed  recruitments  are  also 
indispensable  but  not  easily  available.  Further- 
more, the  relationship  between  ratios  and  actual 
figures  are  too  often  summarized  to  an  extent 
which  forbids  reconstruction  of  the  details  of 
the  computations  and  of  the  results. 

While  the  present  paper  is  mainly  directed 
toward  an  improvement  of  the  cooperation 
between  biologists  and  economists,  it  should 
be  stressed  that  a  prerequisite  is  to  have  access 
to  the  results  of  the  computations  of  the  other 
discipline.  Cooperation  does  not  require  working 
at  the  same  desk,  but  it  would  ask  for  this 
minimum  of  understanding. 

Unfortunately,  the  facility  with  which  the 
Schaefer  model  can  be  used  by  the  economists 
does  not  always  mean  that  there  is  a  perfect 
and  total  understanding  between  fishery  biolo- 
gists and  economists.  More  important  perhaps 
than  the  unit  of  measurement  are  a  few  basic 
concepts  which  are  commonly  used  with  different 
meanings.  The  fishing  effort  concept  is  by  far 
the  most  important  one. 

Fishing  effort  is  in  fact  usually  expressed  in 


many  different  ways:  either  by  its  physical 
characteristics  or  by  its  returns  in  weights  of 
different  fish  species  or  in  money  values  (either 
returns  or  costs,  or  profits).  The  usage  of  these 
different  units  should  be  systematized,  other- 
wise the  concept  of  fishing  effort  would  be 
misleading  as  is  too  often  the  case  when  so 
many  researchers  use  it  with  different  and 
implied  assumptions  on  the  way  it  should  be 
expressed.  In  fact,  there  could  not  be  one  single 
way  of  expressing  fishing  effort;  fishing  effort 
considered  in  its  full  and  general  meaning  is  a 
combination  of  the  different  units  by  which  it 
could  be  expressed. 


Physical  Characteristics  of  Vessel  and  Gear 

This  could  include  any  kind  of  measure 
describing  the  characteristics  of  the  vessel: 
GRT,  power,  length  .  .  .  also  taking  into  account 
items  like  the  number  of  berths  (which  might 
be  significant  for  pole-and-line  techniques),  or 
the  sonar  (for  purse  seining),  or  the  number  of 
pots  (for  crab  or  lobster,  etc.).  Obviously,  for 
each  specific  case  the  most  important  character- 
istic^) to  be  used  as  a  measure  of  the  impact  of 
the  fishing  on  the  stock  or  as  a  measure  of  the 
fishing  power  in  relation  to  a  given  fish  stock 
will  vary.  Therefore,  a  multipurpose  vessel  has 
a  different  fishing  power  according  to  the  gear  it 
is  utilizing;  it  might  even  be  that  the  fishing 
power  has  to  be  different  when  the  same  vessel 
with  the  same  gear  is  exploiting  different  stocks. 
As  a  result  the  fishing  effort  of  the  same  boat 
would  have  to  be  expressed  differently  for  each 
type  of  exploitation,  each  season,  each  year, 
each  stock,  etc. 


Cost  of  Fishing  Effort 

Building  costs  and  operating  costs  which 
could  be  combined  by  using  operating  costs 
including  depreciation  plus  overhead  are  a  more 
permanent  type  of  unit.  First,  the  costs  of  a 
given  boat  are  not  so  much  changed  when  it 
changes  gear.  Secondly,  many  boats  have  been 
built  for  a  definite  type  of  usage.  The  costs  of 
a  boat  will  be  easily  defined  by  so  much  per 
day  at  sea. 


101 


Unit  of  Time 

The  biologist  and  the  economist  will  be 
naturally  inclined  to  use  different  units  of  time 
(time  at  sea  for  the  second  and  time  fishing  for 
the  first).  Anyway,  the  distance  to  the  grounds 
will  have  an  opposite  effect  for  both  researchers, 
the  longer  the  distance,  the  higher  the  costs  or 
the  fishing  effort  for  the  economist;  the  shorter 
the  distance  the  higher  the  impact  on  the  fish 
stocks,  or  the  fishing  effort  for  the  biologist. 

The  conclusion  is  obvious.  There  cannot  be 
such  a  unit  as  a  unit  of  fishing  effort.  Fishing 
effort  is  a  complex  concept;  it  is  a  ratio  or  a 
relationship  between  different  units.  To  assume 
that  it  can  be  defined  once  and  for  all  and  be 
used  indifferently  by  researchers  of  both  dis- 
ciplines, economics  and  biology,  is  a  complete 
mistake.  Each  time  that  the  concept  of  fishing 
effort  is  utilized  it  should  be  made  clear  what 
it  really  means.  Attached  to  a  stock  or  fishing 
technique,  its  value  is  limited  to  this  stock  or 
technique.  Given  in  money  terms  its  compar- 


ability is  attached  to  the  economic  systems  of 
which  it  forms  part. 


CONCLUDING  REMARKS 

A  substantial  complexity  is  the  consequence 
of  the  impossibility  of  building  up  a  complete 
bioeconomic  model,  of  the  difficulty  of  converting 
to  economic  measurement  the  ratios  used  by  a 
number  of  biologists,  of  the  lack  of  a  clear 
understanding  of  what  fishing  effort  is,  of  the 
impossibility  of  forecasting  the  pattern  of 
recruitment  of  the  fish  stocks.  To  overcome  this 
complexity  it  does  not  seem  that  one  can  in- 
definitely rely  upon  equations  which,  whether 
they  are  Schaefer's  or  Beverton  and  Holt's,  are 
mostly  used  analytically  to  give  account  of 
past  developments  but  cannot  make  apparent 
the  mechanisms  through  which  future  develop- 
ments are  taking  place.  Figure  5  shows  that 
these  biological  equations  only  concern  the 
squares  1,  2,  and  3  when  a  complete  simulation 


BIOLOGICAL   RESEARCH 
(by  fish  stock) 


ECONOMIC  RESEARCH 
(by  fishing  fleet) 


Actual  vessels 
catch  rates 


Direct  assessments 

e.g.  by  accoustic 

methods  . 


Corrected  for 
standard  vessels 


Assessments  of 
mortalities  and  of  the 
size  of  the  fish  stock 


Growth  and  recruit- 
ment observations 


Forecasts  of  the 
future  stock  and 
mortalities 


Market 
assessments 


Existing 
fishing  fleets 


Economic  assessment 
of  the  past  and 
present  situation 


Expected 
changes  in  the 
markets 


Expected 
changes  in  the 
fleet 


Necessary 
adaptation 
leading  to 


New  market 
conditions 

Redevelopment 
of  the  fleet 

J 

Figure  5.  —  A  simulated  flow  chart  of  a  fishery. 
102 


model  should  incorporate  the  14  squares  includ- 
ing independent  measures  of  the  size  of  the 
stocks  and  of  recruitment  and  the  feedback 
from  the  economic  side. 

It  is  often  said  in  international  fishery  dis- 
cussions that  no  regulation  should  be  adopted 
or  even  proposed  before  it  can  be  justified  by 
sufficient  "scientific  evidence."  Nobody  is  fooled 
any  more  by  this  sophisticated  expression  which 
means  that  national  economic  short  term 
interests  should  prevail  as  long  as  there  is  no 
definite  proof  that  such  national  interests  are 
leading  to  detrimental  international  economic 
consequences.  It  is  obvious  that  such  scientific 
evidence     has    often     been     supplied     by    the 


biologists,  if  only  when  they  stated  that  numer- 
ous stocks  are  exploited  beyond  the  point  of 
MSY.  But  the  precise  economic  consequences 
of  these  statements  are  very  rarely  available; 
and  there  is  practically  no  case  where  the 
economic  consequences  of  the  cuts  to  be  made  in 
the  fishing  effort  have  been  evaluated  (short 
term  costs  or  losses  and  long  term  benefits 
according  to  the  possible  regulations  to  be 
adopted).  It  is  obvious  that  such  "practical" 
evidence  will  never  be  supplied  without  a  close 
cooperation  between  biologists  and  economists. 
The  possibility  of  successful  fishery  management 
is  entirely  dependent  on  such  bioeconomic 
research  work. 


103 


ISSUES  RELATED  TO  FISHERY  MANAGEMENT 
RESEARCH  RESULTS 


In  the  final  section  concerning  other  issues 
related  to  fishery  management,  the  first  paper 
by  Holmsen  summarizes  the  results  of  his  study 
of  the  Peruvian  anchoveta  fishery.  His  is  very 
much  an  applied  study,  for  he  is  interested  in 
indicating  the  critical  components  of  what  they 
have  done  in  the  past,  the  faults  that  may  exist, 
and  an  evaluation  of  alternative  management 
programs. 

By  his  measure  the  current  excess  capacity 
in  the  fleet  should  be  reduced  by  14-38%  depend- 
ing upon  the  biological  or  social  constraints 
imposed  (length  of  closed  season).  Alternative 
plans  which  might  correct  this  situation  are 
reviewed,  including: 

(1)  restrictions  on  fleet  size. 

(2)  government  purchase  of  scrap  fleet,  the 
cost  to  be  covered  by  an  assessment  on  the 
remainder  of  the  fleet;  new  entry  would  be 
restricted  simultaneously. 

(3)  require  private  scrapping  to  permit  new 
private  construction  —  a  scrapping  ratio. 

(4)  tie  fleet  size  to  licensed  capacity  of  fac- 
tories. 

(5)  a  quota  system  with  variable,  long-lived 
shares  allocated  via  an  auction  system. 

As  there  is  excess  capacity  at  the  processing 
level  also  this  becomes  part  of  the  consideration. 
Possible  controls  here  would  be  (1)  reducing 
licensing  capacity  leading  to  forced  insolvency, 
(2)  government  purchase  of  plants,  or  (3)  trans- 
ferable factory  quotas. 

Holmsen  recommends  a  combination  program 
including  both  levels.  Emphasized  would  be  a 
high  scrap/rebuild  ratio  and  lifting  the  debt 
moratorium  on  plants. 

In  the  paper  by  Thompson,  Callen,  and  Wolken 
the  Thompson  and  George  model,  as  previously 
referred  to,  is  expanded  to  account  for  income 
taxes  and  depreciation.  Emphasizing  the  desire 
for  survival  as  a  key  decision  element  the 
authors  apply  this  model  to  sample  firms  in  the 
Gulf  shrimp  fishery,  using  alternative  sets  of 
price  and  landings  data.  The  critical  nature  of 
each  decision  variable  is  noted  for  each  set  of 
inputs. 

Anderson,  Connolly,  Halter,  and  Longhurst 
present  another  version  of  a  simulation  approach 


to  evaluation  of  management  alternatives, 
relating  experience  in  the  management  of  deer 
population  subject  to  different  hunting  strategies 
defined  by  alternative  sets  of  regulations. 

Some  interesting  general  methodological 
points  are  made  in  this  paper.  Among  these  is 
the  stress  on  the  iterative-feedback  elements  of 
the  simulator.  By  stressing  this  mechanism  in 
fisheries  we  could  obtain  a  continuing  evaluation 
of  the  quality  of  the  input  in  addition  to  the 
quantitative  dimensions  of  alternative  programs. 
Thus,  a  type  of  continuing  sensitivity  analysis 
can  be  performed  on  such  items  as  estimates  of 
MSY,  alternative  measures  of  fishing  power, 
the  existence  of  diminishing  returns,  social 
transfer  costs,  and  alternative  discount  rates. 

As  does  Adam,  the  authors  consider  biological 
issues  to  be  the  essence  of  first  generation  models. 
Second  generation  models  would  include  eco- 
nomics and  other  considerations.  This  differs 
somewhat  from  Pontecorvo,  who  would  have 
biology  and  economics  as  first  and  second 
generation  models,  respectively,  and  other  con- 
siderations as  part  of  third  generation  models. 

A  final  element  of  general  interest  is  the  use 
of  a  random  number  generator  to  create  an 
array  of  "forage  factors."  This  would  be  a 
method  of  considering  the  many  combinations  of 
environmental  factors  that  affect  recruitment  in 
fish  stocks.  In  particular,  as  Pontecorvo  suggests, 
there  may  be  tradeoffs  between  levels  of  accuracy 
and  the  costs  of  these  levels.  This  analysis 
could  be  performed  within  a  complete  simulated 
fishery  system  with  the  aid  of  this  generator. 

The  paper  by  Stevens  and  Mattox  is  actually 
a  report  on  two  separate,  but  related  studies, 
one  on  the  economics  of  salmon  hatchery  opera- 
tions and  the  other  on  the  supply  response  of 
fishing  vessels  (boats)  to  changes  in  catch/effort 
ratios  and  market  conditions.  The  hatcheries 
issue  is  one  which  has  achieved  little  attention 
in  the  economics  literature  and  is  timely  con- 
sidering the  growth  in  salmon  hatcheries  and 
the  increasing  research  and  development  work 
being  conducted  for  other  species. 

That  these  hatcheries  programs  are  critical 
to  the  overall  management  plans  is  a  patently 


104 


obvious,  but  seldom  mentioned,  fact.  As  pointed 
out  by  the  authors,  with  hatchery  fish  ranging 
from  30-80%  of  all  fish  caught  from  hatchery 
streams  and  20%  of  all  Pacific  salmon,  no 
management  program  could  be  successful  with- 
out explicit  Consideration  of  the  hatcheries.  In 
this  examination  of  15  Oregon  hatcheries  pro- 
duction functions  were  estimated  which  indicated 
fixed  input  proportionality,  constant  returns 
to  size  and  substitution  between  the  fixed 
proportional  input  and  water  temperature. 

In  the  study  of  entry  and  exit  an  irreversible 
function  was  found  to  exist.  Entry  followed 
good  years,  but  exit  did  not  follow  bad  years 
to  the  same  degree.  Thus,  successful  "hatchery 
years"  would  lead  to  entry  and  expanded  fleet 
size  which  could  not  be  justified  by  lesser,  even 
average  years.  This  is  a  further  enforcement  of 
the  argument  for  limited  entry  as  the  effective- 
ness of  hatcheries  programs  in  raising  fisher- 
men's incomes  will  be  mitigated  unless  the 
countervailing  tendency  to  overcapitalize  is 
restricted.  Part  of  this  restrictive  element  may 
include  a  deliberate  effort  to  increase  opportunity 
costs,  as  discussed  previously. 

Keen  is  the  only  author  here  reflecting  on  a 
historical  system  used  to  limit  entry,  the 
Japanese  experience.  When  reviewing  this  work 
it  is  necessary  to  recall  that  the  principal 
objective  of  the  Japanese  program  has  always 
been  "to  maintain  the  viability  of  the  individual 
enterprise."  As  this  objective  is  somewhat  akin 
to  "maintaining  the  family  farm"  it  differs  from 
the  objective  held  by  most  economists  to  be 
desirable.  If  the  Japanese  program  can  be  judged 
successful  in  meeting  its  own  objective,  it  may 
still  not  be  suitable  to  our  purposes.  Neverthe- 
less, we  can  proceed  to  evaluate  the  components 
of  the  program  to  determine  its  failure  and 
successes  and  to  gain  an  appreciation  of  the 
critical  decisions  which  need  to  be  made  in  a 
management  program  as  it  evolves  over  time. 

The  Japanese  system  began  in  1946  when  all 
craft  greater  than  10  tons  had  to  be  licensed.  It 
evolved  to  include  area  restrictions  and  to  be 
divided  into  tonnage  groupings,  with  different 
restrictions  for  distant-water  fisheries  as  these 
developed.  Its  principal  overall  characteristic 
was  its  pliability.  When  pressures  for  additional 
development  of  certain  fisheries  mounted,  ad- 


justments were  made  to  allow  for  some  of  this 
investment.  In  some  instances,  when  certain 
fishing  operations  were  no  longer  viable,  attrac- 
tions to  divert  excess  effort  to  other  fisheries 
were  established.  The  principal  thrust  of  these 
regulations  was  to  modify  the  tendency  to  over- 
invest  and  dilute  capital  values.  In  some  in- 
stances, the  growing  value  of  fishing  licenses 
attest  to  the  success  of  this  program. 

Critical  is  the  effect  of  these  programs  on  the 
development  of  technology.  It  can  be  shown  that 
in  some  cases  technology  took  some  strange 
courses  because  of  the  regulations,  somewhat 
akin  to  our  own  Alaskan  limit  seiners.  This  and 
other  elements  of  an  existing  scheme  could 
prove  a  fruitful  area  of  examination  in  the 
future,  now  that  substantial  progress  has  been 
made  in  theoretical  studies. 

The  final  paper  by  Huq  is  so  timely  as  to 
appear  to  be  at  the  unanimous  request  of  the 
other  authors  and  participants  in  the  workshop. 
This  is  because  the  subject  is  labor  mobility  and 
social  transfer  costs,  with  the  study  reported  on 
being  confined  to  three  representative  com- 
munities in  the  Maine  pot-lobster  fishery. 

In  this  study  the  goal  is  to  evaluate  such 
measures  of  labor  mobility  as  age,  level  of 
education,  income  levels,  technical  skills,  other 
employment,  time  in  present  occupation,  invest- 
ments in  the  fishery,  attitudes  toward  fishing 
as  an  occupation,  and  attitudes  toward  certain 
elements  of  the  harvesting  process  so  that  alter- 
native forms  of  limited  entry  would  be  evaluated. 
Results  indicate  that  immobility  is  substantial, 
but  that  this  may  not  be  a  problem  as  the 
limitation  may  successfully  be  applied  to  capital 
inputs  with  little  reduction  in  the  labor  input 
for  much  of  the  sample  examined  in  the  three 
communities.  For  the  remainder,  some  form  of 
an  adjustment  assistance  program  may  be 
necessary,  particularly  since  a  portion  of  the 
labor  force  in  the  fishery  is  currently  supple- 
menting public  assistance  or  social  security 
incomes  with  its  lobstering  activity.  These 
members  of  the  labor  force  truly  have  limited 
opportunities.  Restricting  their  participation 
would  place  a  greater  burden  on  other  family 
members,  who  may  also  be  in  the  lobster  fishery. 

A.A.S. 


105 


Management  of  the  Peruvian  Anchoveta  Resource 


Andreas  A.  Holmsen1 


ABSTRACT 

The  best  available  estimate  of  the  maximum  sustainable  yield  of  the  Peruvian 
anchoveta  resource  is  9.5  million  metric  tons  (  ±  1  million).  The  productive  capacity  of 
the  purse-seine  fleet  and  the  fishmeal  factories  far  exceed  this  tonnage  with  the  result 
that  the  open  season  is  becoming  shorter  year  by  year.  This  paper  describes  the  current 
fishery  management  program  in  Peru  and  the  degree  of  overinvestment  in  the  industry. 
It  further  outlines  the  alternative  methods  which  can  be  used  to  reduce  excess  capacity 
in  the  catching  and  processing  phase  and  the  advantages  and  disadvantages  of  the 
various  alternatives. 


INTRODUCTION 

It  is  well  known  among  fisheries  people  that 
Peru  is  the  leading  fishing  country  of  the  world 
in  terms  of  tonnage  landed.  About  97%  of  the 
catch  is  anchoveta,  which  is  used  strictly  for 
production  of  fishmeal  and  oil.  Besides  the 
employment  and  earnings  derived  from  the 
harvesting  and  processing  of  this  resource,  fish- 
meal makes  another  valuable  contribution  to  the 
economy  of  Peru.  Like  many  other  less  developed 
countries,  Peru  has  balance  of  payments  prob- 
lems and  exports  of  fishmeal  and  oil  account  for 
approximately  one-third  of  foreign  earnings. 
With  the  exception  of  Iceland,  I  doubt  that  any 
other  country  is  as  dependent  on  its  fishery 
resource  as  Peru,  and  few  are  so  concerned 
about  it. 

To  protect  the  resource  Peru  claims  a  200- 
mile  fisheries  limit  which  may  be  twice  as  much 
as  is  necessary.  Seventy  miles  is  the  maximum 
distance  from  shore  that  anchoveta  fishing 
takes  place.  The  stock  is  concentrated  in  the 
waters  off  the  southern  two-thirds  of  the  country, 
so  except  for  some  mixing  on  the  Chilean  border 
it  is  entirely  a  national  resource. 

Peru's  emergence  as  a  fishing  nation  began 
in  the  1950's,  but  most  of  the  growth  of  the 
industry  has  taken  place  during  the  last  decade. 
During  the  1960-61  fishing  season  (September- 
August)  Peru's  landings  of  anchoveta  were 
about  4  million  metric  tons.  During  the  1969-70 

1  Department  of  Resource  Economics,  University  of 
Rhode  Island. 


season,  landings  reached  about  11  million  metric 
tons,  and  every  season  during  the  decade  land- 
ings were  higher  than  the  previous  year. 

During  the  early  years  of  the  decade,  the 
rapid  development  of  the  industry  took  place 
with  little  planning,  basic  knowledge,  and 
experience.  As  a  result,  overexpansion,  particu- 
larly in  processing  capacity,  has  plagued  the 
industry  ever  since. 

The  number  of  vessels  in  the  fleet  reached  a 
high  of  1,778  vessels  during  the  1963-64  season, 
but  later  gradually  declined  to  the  current 
size  of  about  1,400.  The  vessels  have  become 
bigger  every  year,  however.  While  5-6  years 
ago,  a  vessel  with  180-ton  hold  capacity  was  a 
large  vessel,  the  smallest  built  today  has  a 
capacity  of  275  tons  and  most  vessels  built 
during  the  last  2  years  have  a  350-ton  capacity. 
Thus,  the  fleet  capacity  has  increased  from 
about  180,000  tons  capacity  in  the  mid-60's  to 
somewhat  above  200,000  tons  during  the  1969-70 
season. 

A  large  part  of  the  fleet  is  considered  obsolete, 
consisting  of  wooden  vessels  built  from  1962 
to  1964  (in  Peru,  7  years  are  considered  the 
economic  life  of  such  vessels).  In  recent  years, 
most  vessels  have  been  built  of  steel  and  con- 
struction of  fiberglass  vessels  has  started.  Echo 
sounder,  powerblock,  and  fish  pump  are  standard 
equipment  in  the  fleet,  and  the  most  modern 
vessels  also  have  sonar.  A  fishing  trip  normally 
is  a  day  trip,  the  vessel  leaving  early  in  the 
morning  and  returning  with  or  without  catch 
in  the  afternoon. 


106 


Most  of  the  Peruvian  fishing  fleet  is  owned  by 
firms  who  also  own  factories  and  only  about  20% 
of  the  fleet  is  owned  by  independent  vessel 
owners.  A  fair  number  of  these  are  tied  to  a 
particular  factory,  however,  and  have  to  deliver 
their  catch  there,  owing  to  financial  help 
rendered  when  buying  the  vessel  or  for  similar 
reasons. 

As  the  number  of  vessels  has  declined  so  has 
the  number  of  processing  plants.  A  consolidation 
has  taken  place  into  fewer  and  larger  units. 
Currently,  Peru  has  127  fishmeal  factories  with 
a  total  capacity  of  close  to  8,000  tons  of  fish 
per  hour.  About  10  of  these  plants  did  not 
operate  last  season.  While  most  firms  own  only 
one  factory,  a  number  of  larger  firms  own 
several  each.  These  are  generally  located  in 
different  ports  or  geographic  regions  as  a  hedge 
against  poor  fishing  in  one  particular  area. 

CURRENT  MANAGEMENT  PROGRAMS 

Both  the  Peruvian  authorities  and  the  Peru- 
vian fishing  industry  have  for  several  years 
been  aware  of  the  danger  of  overexploiting  the 
anchoveta  stock,  and  have  taken  steps  to  reduce 
the  pressure  on  the  resource.  Fishing  effort 
expanded  quickly  until  the  1963-64  season  when 
the  total  catch  reached  a  level  of  8  million  tons. 
Thereafter,  first  closed  seasons  and  then  overall 
catch  quotas  were  established.  At  the  present 
time,  the  following  programs  or  restrictions 
are  in  force: 

1.  The  fishery  is  closed  on  Saturdays  and 
Sundays. 

2.  The  fishery  is  closed  about  1  month  in 
summertime  during  the  "peladilla"-season. 
That  closure  ("veda")  takes  place  when 
there  are  large  amounts  of  small  fish 
(peladilla)  in  the  catch.  The  time  of  the  veda 
varies  from  year  to  year.  In  1970  the 
closure  was  from  mid-February  to  mid- 
March,  which  was  too  late. 

3.  During  the  fishing  season,  after  the  pela- 
dilla have  entered  the  fishery  and  explora- 
tory cruises  to  assess  the  recruitment  have 
taken  place,  an  overall  quota  is  established 
for  the  season.  When  this  quota  is  reached, 
the  fishery  is  closed.2 


2  Except   from  the   port   of  Ilo   close   to  the   Chilean 
border. 


4.  Each  factory  has  been  given  a  license  for 
a  certain  daily  input  of  raw  material.  The 
license  capacity  is  stated  in  terms  of  tons 
per  hour.  This  quantity  multiplied  by  24 
is  the  maximum  quantity  a  factoiy  is 
permitted  to  accept  in  one  day.  Due  to  the 
fact  that  both  the  licensed  and  the  technical 
capacities  of  the  fishmeal  factories  have  far 
exceeded  landings,  factory  licenses  have 
not  been  effective  in  reducing  fishing 
pressure. 

THE  CURRENT  SITUATION 

The  Anchoveta  Resource 

Anchoveta  generally  spawns  in  late  winter 
(August)  and  reaches  a  harvestable  stage  in 
midsummer  (December-February).  It  has  a  life 
span  of  2  to  3  years.  In  the  early  and  middle 
60's,  fish  1-year  old  or  more  contributed  to  most 
of  the  catch,  while  later  the  zero  year  class  has 
become  dominant  in  the  annual  catch  and  its 
percentage  of  the  total  catch  is  increasing.  This 
is  considered  a  warning  signal.  Actually  at  the 
beginning  of  last  season,  September-November 
1970,  the  catch  was  lower  per  month  than  in  any 
month  in  1965,  five  years  ago.  The  rich  1969-70 
fishery  did  not  perform  well  before  the  zero 
year  class  came  of  size.  An  FAO  panel  on  stock 
assessment  which  met  in  Peru  in  January  1970 
came  to  the  conclusion  that  the  maximum  sus- 
tainable yield  of  the  Peruvian  anchoveta  resource 
most  probably  was  9.5  million  tons  (  ±  1  million 
tons).  The  experts  recommend  that  the  authori- 
ties permit  a  10-million  ton  catch  coupled  with 
close  observation  of  the  fishery  to  see  what  effect 
this  fishing  pressure  would  have.  The  authori- 
ties, however,  permitted  11  million  tons  to  be 
caught,  which  biologists  think  will  significantly 
hurt  the  fishery  in  1970-71,  both  because  too 
much  of  the  1-year  class  already  might  have 
been  harvested  and  possibly  also  due  to  reduced 
reproductive  stock. 


Fishing  Pressure 

While  the  summer  veda  is  of  biological  sig- 
nificance since  it  prevents  the  catching  of  large 
quantities  of  very  small  fish,  and  while  the 
prohibition  of  weekend  fishing  might  have  some 


107 


social  advantage,  the  long  winter  veda,  which 
has  been  increasing  over  time  to  reach  SV2 
months  in  1970,  is  only  due  to  an  excessive 
catch  capacity  of  the  fleet  relative  to  the  resource 
available.  Given  a  maximum  sustainable  yield 
of  9X2  million  tons,  increases  in  capacity  or 
technological  improvements  of  the  fleet  will 
mean  a  longer  winter  veda. 

During  the  two  years  from  1966-67  to  1968-69, 
hold  capacity  increased  by  16,000  tons  per 
year,  or  about  equal  to  old  tonnage  leaving  the 
fleet.  During  1969-70,  however,  about  32,000 
tons  of  new  construction  entered  the  fleet  and 
according  to  interviews  with  the  various  ship- 
yards that  rate  of  construction  has  continued 
for  the  remainder  of  1970  (Holmsen,  1970b). 
Thus,  the  fishing  season  (the  number  of  fishing 
days  permitted)  has  gradually  declined  from 
289  in  1963-64  to  166  days  in  1966-67  and  155 
days  in  1969-70.  To  catch  a  quota  of  9V2  million 
tons,  a  145-day  fishing  season  would  have  been 
sufficient  in  1969-70.  Due  to  the  amount  of  new 
construction,  with  the  same  abundance  and 
availability  of  fish  as  last  season,  the  fleet 
would  be  able  to  catch  9V2  million  tons  in  less 
than  140  days  in  1970-71.  As  long  as  fishmeal 
prices  are  high  and  factories  have  to  pay  con- 
siderably more  to  independent  owners  per  ton 
of  fish  than  the  cost  per  ton  for  operating  their 
own  vessels,  construction  will  continue,  resulting 
in  a  shorter  and  shorter  season,  to  the  detriment 
of  the  industry  as  a  whole.  There  are  similar 
examples  from  other  fisheries  where  overall 
catch  quotas  have  been  established  with  no 
limit  to  entry,  such  as  Pacific  halibut  and 
yellowfin  tuna. 

Peru  is  short  of  investment  capital  and  par- 
ticularly short  of  foreign  exchange.  In  addition 
to  being  a  misallocation  of  capital,  however,  the 
pressure  of  an  excessive  fleet  poses  the  danger 
of  pressure  on  government  to  keep  the  season 
open  longer  than  the  period  recommended  by 
stock  assessment  experts. 


Processing  Capacity 

The  same  problem  of  overcapacity  is  found  in 
the  processing  phase.  Some  years  ago,  the  gov- 
ernment prohibited  the  building  of  more  fac- 
tories and  issued  licenses  restricting  the  input 
to  a  specific  tonnage  per  hour  for  the  existing 


fishmeal  plants.  The  technical  capacities  of 
various  plants  were  increased,  however,  without 
regard  to  the  license.  Last  year,  the  government 
started  to  enforce  the  law  and  several  firms  had 
to  buy  plants  to  bring  their  own  licensed  capac- 
ity up  to  their  technical  capacity,  even  when 
they  had  no  use  for  the  purchased  plant's  build- 
ings or  equipment.  Thus,  some  consolidation 
took  place  and  the  total  licensed  capacity  now 
reasonably  reflects  the  total  technical  capacity. 
The  licensed  capacity  is  about  50%  more  than  is 
needed,  however,  even  with  the  short  season 
now  in  effect,  and  the  excess  capacity  would  of 
course  be  even  greater  if  the  fleet  size  were  re- 
duced so  the  season  became  longer. 

The  fishmeal  industry  as  a  whole  is  deep  in 
debt,  liabilities  about  equal  to  assets.  Since 
some  firms  are  in  a  good  financial  position,  this 
means  that  many  firms  are  thoroughly  insolvent, 
and  would  have  been  bankrupt  but  for  a  mora- 
torium on  debt  collection. 


DESIRABLE  OBJECTIVES 

The  problem  facing  the  Peruvian  anchoveta 
industry  is  how  to  reduce  the  excess  capacity 
both  in  the  catching  and  the  processing  phase, 
so  that  excessive  closed  seasons  can  be  prevented 
and  the  productivity  of  the  remaining  production 
units  improved. 

A  reduction  in  capacity  and  lengthening  of 
the  fishing  season  has  a  fourfold  advantage: 

1.  Less  pressure  will  be  placed  on  the  govern- 
ment to  exceed  recommended  levels  of  catch. 

2.  Fewer  investment  funds  will  be  needed  for 
the  industry. 

3.  The  remaining  units  will  be  more  produc- 
tive and  thereby,  the  economic  situation 
for  the  industry  will  improve. 

4.  The  sustainable  yield  in  the  fishery  will 
increase,  as  more  fish  will  be  caught  at  a 
higher  age  or  larger  sizes. 

The  cost  savings  which  will  accrue  depend  on 
the  percentage  of  fixed  and  variable  costs  in  the 
catching  and  processing  phase.  For  the  catching 
phase,  it  will  also  depend  on  what  percentage 
of  the  variable  costs  are  associated  with  volume 
and  what  percent  with  time. 

Based  upon  budgetary  data  for  1970-71  from 
a  handful  of  companies,  the  following  break- 


108 


down  might  be  a  reasonable  approximation.3 
Forty-seven  percent  of  the  cost  of  harvesting 
was  found  to  be  fixed  and  not  related  to  the 
number  of  fishing  days,  nor  the  size  of  catch. 
Thirty-five  percent  of  the  cost  was  apportioned 
to  the  size  of  the  catch,  of  which  34%  was  the 
crew  share  and  social  benefits.  The  remaining 
18%  was  related  to  the  number  of  days  the 
vessels  were  out  fishing,  catch  or  no  catch. 

For  a  fishmeal  factory  the  cost  of  fish  is  a 
variable  expense  and  this  item  alone  amounted 
to  59%  of  total  cost.  The  variable  cost  of  pro- 
ducing meal  and  oil  amounted  to  75%  of  total 
costs  and  the  fixed  cost  25% .  Excluding  the  cost 
of  the  fish,  the  variable  costs  were  39%  and  the 
fixed  costs  61%  (Holmsen,  1970a). 

What  the  current  overcapacity  in  industry  is 
depends  on  what  kind  of  management  program 
one  has  in  mind  —  whether  one  recommends  a 
1-  or  2-month  peladilla  veda,  whether  one  sticks 
to  the  5-day  week  rather  than  a  7-day  week,  etc. 
Based  on  various  alternatives  from  a  7-day 
week  and  no  veda  to  a  5-day  week  and  a  2-month 
peladilla  veda,  the  fleet  reduction  necessary 
was  found  to  range  from  38%  to  14%  (Boerema 
and  Holmsen,  1970).  By  using  the  coefficients 
above,  this  would  lead  to  savings  ranging  from 
about  $20  million  annually  in  the  first  case  to 
about  $6  million  in  the  latter  case.  The  savings 
in  the  processing  phase  would  also  be  significant. 
An  FAO  management  panel,  which  met  in  Peru 
in  June  1970,  concluded  that  the  technical 
capacity  of  the  factories  could  be  reduced  nearly 
50%  under  year-round  fishing,  and  that  total 
savings  to  industry  from  reduction  of  fleet  size 
and  number  of  plants  could  perhaps  run  as  high 
as  $50  million.  No  value  can  presently  be  put 
on  the  lessened  risk  of  overfishing  and  depletion 
of  the  stock. 

ALTERNATIVE  CONTROLS 

A  fisheries  management  program  should  have 
a  double  goal:  1)  to  protect  the  resource  from 
overexploitation,  and  2)  to  prevent  overinvest- 
ment and  economic  wastes  in  harvesting  and 
processing.    To    achieve    these    goals    in    the 


3  The    percentages  are    median    observations    based 

upon     representative  vessel     size     (140-     to     220-ton 

capacity)    and   plants  with   technical   capacity   of  60-90 
tons  per  hour. 


anchoveta  fishery,  restrictions  can  be  put  on  the 
fleet  or  on  the  factories  or  on  both.  Some 
programs  might  achieve  the  desired  result 
rather  fast,  while  others  might  take  more  time. 
Alternative  programs  related  to  the  catching 
phase  will  first  be  discussed. 


Restrictions  on  Fleet  Size 

(1)  A  reduction  in  the  size  of  the  fleet  to  the 
desired  level  can  be  achieved  by  an  embargo  on 
new  construction.  Despite  the  fact  that  a  number 
of  vessels,  which  otherwise  would  have  been 
scrapped,  would  be  repaired  and  remain  in  the 
fishery,  a  fair  number  of  vessels  would  disappear 
from  the  fishery  each  year  and  the  season  for 
those  remaining  would  become  longer.  Argu- 
ments against  such  a  proposal  would  be  that 
older,  smaller  vessels  in  the  fishery,  which  are 
the  highest  cost  producing  units,  would  get  an 
additional  "lease  on  life"  and  the  fleet  would 
stagnate  technically. 

(2)  Another  possibility  with  immediate  effect 
would  be  for  the  government  to  buy  up  the  scrap 
part  of  the  fleet  (the  high  cost  producer),  and 
assess  the  cost  on  the  remainder  of  the  industry, 
preferably  through  a  fee  per  ton  of  meal  produced. 
A  large  number  of  such  vessels  would  have  to 
be  bought  since  each  contributes  very  little  to 
the  total  catch.  The  industry  would  be  better 
off,  however,  since  the  marginal  cost  of  the 
remaining  vessels  would  be  far  below  the  average 
cost  of  the  vessels  removed  from  the  fishery. 
Such  a  program  would  have  no  long  run  effect, 
however,  if  restrictions  on  new  construction 
were  not  implemented  at  the  same  time. 

A  scrapping  ratio  would  have  to  be  intro- 
duced limiting  the  annual  output  of  productive 
capacity  to  the  amount  of  productive  capacity 
leaving  the  fleet  during  the  year.  Such  a  pro- 
gram, which  has  some  support  in  Peru,  still 
leaves  a  difficult  question  unanswered.  Which 
vessels  should  the  government  buy  and  scrap 
and  what  would  the  prices  be?  Two  6-year  old 
150-ton  vessels  are  not  necessarily  worth  the 
same  price.  Appraisal  and  judgment  are  called 
for,  which  could  easily  result  in  kickbacks  in  a 
country  where  civil  service  salaries  are  low  and 
where  bribery  has  not  been  unfamiliar  in 
doing  business. 


109 


(3)  A  third  alternative  would  be  to  rely 
entirely  on  a  scrapping  ratio.  If  a  firm  or  in- 
dividual wants  to  build  a  new  vessel,  he  would 
have  to  scrap  a  larger  tonnage  of  old  vessels.  If 
a  firm  has  no  vessels  to  scrap,  it  will  have  to 
buy  tonnage  for  scrapping.  The  time  necessary 
for  an  adjustment  of  the  fleet  size  to  the  desired 
level  will  be  longer  than  under  the  previous 
alternative.  A  scrapping  ratio  (based  either  on 
gross  tonnage  or  tonnage  capacity)  has  to  be 
relatively  high  in  the  beginning,  possibly  three 
to  one,  but  will,  over  time,  come  close  to  one  to 
one,  just  sufficiently  high  to  offset  the  effect  of 
technological  improvements  in  vessels  and  gear. 
If  nobody  is  willing  to  scrap  vessels  at  the 
initial  ratio  (except  for  credits  obtained  when 
vessels  sink  or  burn)  the  effect  will  be  the  same 
as  an  embargo  on  vessel  construction.  The 
price  of  obsolete  vessels  will  then  be  close  to 
zero,  however,  so  some  new  construction  will 
surely  take  place.  Some  vessels,  which  ordinarily 
would  not  have  been  removed  from  the  fishery, 
might  be  removed  if  the  owner  can  sell  them  to 
someone  needing  tonnage  to  scrap. 

This  program  falls  somewhere  between  the 
two  previously  mentioned,  but  neither  does  it 
involve  government  outlays  nor  does  it  prevent 
technological  improvements  in  the  fleet  during 
the  transition  period.  All  these  three  programs 
would  necessitate  a  scrapping  ratio  when  the 
fleet  is  reduced  to  the  desired  level. 

(4)  Recommendations  have  been  made  to  the 
government  of  Peru  to  reduce  fishing  effort  by 
tying  the  size  of  the  fleet  to  the  licensed  capacity 
of  the  factories.  The  recommendations  called 
for  a  maximum  of  1.4  tons  of  hold  capacity  per 
ton  of  daily  processing  capacity.4  Even  if  a 
ratio  were  imposed  on  a  firm  (some  firms  have 
several  factories)  rather  than  on  a  factory  so 
that  vessels  can  be  used  where  fish  are  abundant, 
there  seems  to  be  certain  disadvantages  with 
such  a  program.  While  previous  programs 
mentioned  have  not  differentiated  between 
factory  owned  and  independently  owned  vessels, 
the  question  now  arises  as  to  how  to  deal  with 
the  20%  of  the  fleet  which  is  independently 
owned.  Secondly,  such  a  program  would  lessen 
competition  and  freeze  the  industry  in  a  given 
pattern. 


4  This  ratio  is  too  high,  as  few  firms  currently  have 
a  higher  ratio. 


(5)  To  reduce  the  size  of  the  fleet  and  expand 
the  fishing  season,  a  quota  system  can  also  be 
implemented.  Catch  quotas  can  be  established 
for  individual  vessels,  factories,  or  firms.  To 
reduce  uncertainties  about  investment,  quotas 
should  be  given  for  a  number  of  years  and  not 
for  one  season  at  a  time.  Further,  due  to  changes 
in  recruitment  and  the  amount  of  effort  the 
resource  can  bear,  quotas  should  be  allotted 
as  a  percentage  of  the  overall  annual  quota. 

A  quota  system  for  the  purpose  of  reducing 
the  number  of  producing  units  would  most  likely 
have  to  be  based  on  an  auction  system.  Such  a 
system,  whether  introduced  on  the  vessel, 
factory,  or  company  level,  would  tend  to  elimin- 
ate not  only  the  less  efficient  producers  but 
also  those  which  are  financially  weak.  Such  a 
program  would  transfer  significant  funds  from 
the  fishing  industry  to  the  public  treasury.  Due 
to  the  structure  of  the  Peruvian  anchoveta 
industry,  a  company  quota  would  seem  prefer- 
able as  this  would  reduce  the  size  of  the  fleet 
(overhead  costs)  more  than  a  quota  on  factories 
or  vessels.  Under  the  two  latter  arrangements, 
many  vessels  may  be  tied  up  because  they  have 
reached  this  quota,  while  others  still  are  fishing 
because  a  factory  may  be  located  in  an  area 
where  availability  of  fish  is  low  in  a  particular 
season  resulting  in  excessive  steaming  time  by 
the  factory  fleet.  Even  company  quotas  would 
result  in  an  excessive  fleet,  however,  as  each 
company  would  keep  a  fleet  big  enough  to  be 
sure  it  will  catch  its  quota. 

The  various  management  alternatives  so  far 
mentioned  have  been  directed  towards  reducing 
the  capacity  of  the  fleet  and  extension  of  the 
fishing  season  and  thus,  reducing  the  size  of 
investment  in  the  catching  phase.  Some  of  the 
alternatives  will  have  little  or  no  impact  on  the 
excess  investment  and  low  capacity  utilization 
of  the  fishmeal  factories,  while  others  will  have 
a  significant  impact. 


Reduction  in  Processing  Capacity 

(1)  Reduction  of  the  total  licensed  capacity  of 
fishmeal  plants  will  indirectly  affect  the  fleet. 
As  indicated  earlier,  the  industry  as  a  whole  is 
in  a  poor  financial  position.  By  lifting  the 
moratorium    on    debt    collection,    many    firms 


110 


would  go  bankrupt  and  this  would  improve  the 
situation  for  those  remaining.  Since  most  of  the 
debt  is  to  the  public  sector,  it  would  mean  the 
government  would  have  to  write  off  some  bad  or 
uncollectible  claims. 

(2)  Spokesmen  for  Sociedad  National  de 
Pesqueria  (a  trade  organization  for  the  fishmeal 
producers)  are  extremely  concerned  about  excess 
capacity  and  have  indicated  a  willingness  to 
bail  out  the  government  through  a  program 
where  the  government  buys  up  the  high-cost 
plants  and  assesses  the  cost  on  the  remainder 
of  the  firms  over  2-3  years  by  a  fee  per  ton  of 
meal  produced.  Whatever  methods  are  used 
for  eliminating  the  excess  capacity,  they  will  be 
beneficial  for  the  industry  as  a  whole  and  reduce 
the  pressure  on  the  government  to  increase  the 
overall  catch  quota. 

(3)  In  addition  to  eliminating  the  insolvent, 
high  cost,  or  marginal  producers,  a  further 
reduction  in  the  licensed  processing  capacity 
will  be  needed.  Capacity  should  be  reduced  to 
a  level  just  sufficient  to  process  the  catch  over 
an  extended  fishing  season.  Currently  the 
licensed  capacity  of  a  plant  is  for  tons  of  fish  per 
hour,  and  only  rarely  does  a  factory  produce  at 
full  capacity.  To  encourage  fleet  reduction,  the 
license  should  be  issued  to  companies  rather 
than  on  a  factory  basis  and  as  previously 
mentioned,  should  be  a  percentage  of  the  overall 
catch  quota.  A  quota  might  be  either  on  input 
of  fish  or  output  of  meal.  The  latter  is  easier 
to  control  since  the  meal  is  exported  through  a 
government  monopoly.  A  quota  on  input,  how- 
ever, would  give  a  strong  incentive  to  increase 
the  yield  (output  per  ton  of  fish)  and  improve- 
ment in  this  respect  is  badly  needed.  Quotas 
could  be  based  on  the  company's  current  share 
of  the  market,  or  be  put  up  for  auction. 


Quotas  or  licenses  to  operate  might  be  trans- 
ferable or  nontransferable.  A  transferable  quota 
could  put  large  and  small  companies  (one-plant 
operators  and  multiplant  operators)  on  a  more 
equal  competitive  basis.  The  author  can  see 
little  advantage  in  a  nontransferable  quota 
except  for  the  fact  that  it  might  prevent  con- 
solidation of  the  industry  into  too  few  hands. 

CONCLUSIONS 

To  manage  the  anchoveta  industry  solely 
through  regulation  of  the  processing  phase 
would  very  likely  put  the  independent  vessel 
owners  at  a  serious  disadvantage.  To  prevent 
this,  a  management  program  for  the  Peruvian 
anchoveta  industry  should  include  both  regula- 
tions at  the  catching  and  the  processing  level. 
Of  the  various  alternatives  available  for  manage- 
ment of  the  Peruvian  anchoveta  industry,  the 
author  would  be  in  favor  of  relying  on  a  fairly 
high  scrap  and  rebuild  ratio  to  reduce  the  fleet. 
Lifting  of  the  moratorium  on  debt  collection, 
combined  with  transferable  licenses  for  factories, 
so  market  forces  could  be  effective,  might  be 
sufficient  to  reduce  processing  capacity  to  the 
desired  level. 

LITERATURE  CITED 

BOEREMA,  L.  K.,  and  A.  HOLMSEN.  1970.  Some 
Economic  Aspects  of  Management  of  Fleet  Size  in  the 
Peruvian  Anchoveta  Industry.  Unpublished  manuscript. 

HOLMSEN,  A.  1970a.  Cost  Structure  in  the  Peruvian 
Anchoveta  Industry.  Unpublished  manuscript. 

HOLMSEN,  A.  1970b.  Factors  Affecting  the  Potential 
Productivity  of  the  Peruvian  Anchoveta  Fleet.  Un- 
published manuscript. 


Ill 


A  Stochastic  Investment  Model 
for  a  Survival  Conscious  Fishing  Firm 

Russell  G.  Thompson,  Richard  W.  Callen, 
and  Lawrence  C.  Wolken1 


ABSTRACT 

In  this  study,  the  stochastic  investment  model  for  a  survival  conscious  firm  developed 
by  Thompson  and  George  (1970)  is  extended  to  take  into  account  income  taxes  and 
depreciation  of  the  capacity.  This  model  is  applied  to  shrimp  fishing  on  the  Texas  Gulf 
coast.  Values  of  the  parameters,  as  in  the  deterministic  application  by  Thompson  et  al. 
(1970),  were  based  on  proprietory  information,  current  market  conditions,  and  present 
institutional  restrictions.  The  effect  of  growth  in  real  per  capita  income  on  shrimp 
prices  is  estimated,  and  two  different  rates  of  income  growth  are  analyzed.  Solutions 
to  six  problems  based  on  two  different  sets  of  random  sequences  are  computed  and 
discussed.  The  results  indicate  the  effect  of  the  survival  constraint  on  investment 
decisions,  and  the  importance  of  revealed  information  in  decisionmaking. 


INTRODUCTION 


In  1970,  Thompson  and  George  formulated  a 
stochastic  dynamic  investment  model  for  the 
survival  conscious  firm,  derived  the  optimal 
decision  rules  for  investment,  and  computed 
solutions  to  several  problems.  This  model  takes 
into  account  the  probability  distribution  of  the 
yield  (output  per  unit  of  capacity)  and  output 
price,  as  well  as  all  of  the  information  known  to 
the  decisionmaker  at  the  time  of  each  investment 
decision.  The  entrepreneur  is  initially  assumed 
to  be  in  a  financial  position  where  a  feasible 
investment  solution  always  exists  if  the  lowest 
output  price  and  yield  occur  in  every  period  of 
the  planning  horizon.  In  the  model,  the  objective 
of  the  firm  is  to  maximize  expected  net  worth  at 
the  end  of  the  planning  horizon.  All  production 
expenses,    investment    outlays,    interest   costs, 


1  Russell  G.  Thompson  is  Professor  of  Quantitative 
Management  Science,  University  of  Houston;  Richard 
W.  Callen  and  Lawrence  C.  Wolken  are  Lecturers  in 
Quantitative  Management  Science,  University  of  Hous- 
ton. This  work  was  partially  supported  by  the  National 
Science  Foundation  GH  59  as  a  part  of  the  Sea  Grant 
Program  for  1970. 


and  planned  cash  withdrawals  must  be  paid 
for  as  incurred  (or  scheduled). 

In  this  study,  the  Thompson-George  model  is 
extended  to  take  into  account  income  taxes  and 
depreciation.  This  requires  the  introduction  of 
another  state  variable  to  account  for  the  value 
of  the  firm's  capital  —  the  investment  in 
capacity.  Straightforward  extensions  of  the 
fundamental  constructs  (developed  by  Thompson 
and  George)  were  required,  and  are  available 
from  the  authors  upon  request. 

Because  of  the  vagaries  in  fish  prices  and 
catches,  this  model  would  be  expected  to  be  a 
particularly  appropriate  decision  aid  for  invest- 
ments in  fishing  capacity.  There  are  generally 
few,  if  any,  alternative  uses  for  specialized 
fishing  equipment.  Also,  fishermen  typically 
have  poor  alternative  opportunities  by  which  to 
earn  a  living.  Low  prices  and  small  catches 
would  be  expected,  as  a  result,  to  be  dreaded 
much  more  than  high  prices  and  large  catches 
are  desired.  A  sequence  of  worse  than  expected 
net  revenues  (even  in  the  case  of  a  very  favorable 
expectation)  could  terminate  the  existence  of 
the  fishing  firm.  This  could  well  be  an  unaccept- 
able risk  of  failure.  Hence,  survival  of  the 
fishing  firm  would  be  expected  to  be  a  funda- 
mental factor  influencing  the  firm's  investment 
decisions. 


112 


DEVELOPMENT  OF  THE 
SURVIVAL  MODEL 

In  the  survival  model,  the  decisionmaker 
evaluates  the  worst  sequence  of  net  revenues 
that  could  occur  in  every  year  of  the  decision- 
making period.  This  sequence,  in  conjunction 
with  the  value  of  the  initial  investment  in 
capacity  and  the  value  of  the  money  account, 
determine  the  survivable  set  of  fishing  capacity 
purchases  at  the  beginning  of  the  first  year. 
The  decisionmaker  selects  from  this  set  the 
investment  that  contributes  the  most  to  his 
terminal  net  worth.  After  the  first  year  and 
before  the  second  operating  year  begins,  the 
output  price  received  and  the  yield  obtained  in 
the  first  year  have  been  observed.  This  is  now 
a  part  of  the  information  known  to  the  decision- 
maker for  planning  in  the  second  year.  The 
decisionmaker  again  evaluates  the  worst  se- 
quence of  yields  and  prices  that  could  occur  in 
every  remaining  year  of  the  decisionmaking 
period.  This  abbreviated  sequence  is  now 
evaluated  in  conjunction  with  the  capacity  and 
money  position  at  the  end  of  the  first  year.  It 
determines  the  survivable  set  of  capacity  pur- 
chases for  the  second  year.  Again,  as  in  the 
first  year,  the  decisionmaker  selects  from  this 
second  set  the  investment  that  contributes  the 
most  to  his  terminal  net  worth.  This  procedure 
is  repeated  in  every  year  throughout  the 
decisionmaking  period.  Investment  decisions 
are  conditioned  by  experience,  and  are  not  based 
solely  on  expected  values. 

By  definition,  the  firm  survives  in  a  given  year 
if  the  value  of  the  capacity  exceeds  the  value  of 
the  indebtedness.  A  survivable  investment  is 
defined  in  the  following  way:  the  decisionmaker 
has  completed  operations  in  year  k-1  and  is  now 
planning  for  year  k.  He  wants  to  survive  above 
all  else  during  the  remaining  N  —  (k-1)  years 
of  the  decision  period,  even  if  all  future  yields 
and  prices  are  the  lowest  possible.  An  invest- 
ment decision  in  the  fcth  year,  s^,  is  said  to  be 
survivable  if  the  value  of  the  capacity  in  every 
remaining  year  is  never  less  than  the  indebted- 
ness owed  (with  capacity  not  being  purchased 
in  any  of  the  years  after  the  fcth  0ne  and  the 
lowest  net  revenues  being  visualized  in  every 
year  of  the  yet  undisclosed  future). 

Under  these  conditions,  a  survivable  ca- 
pacity   purchase    in    year    k    is    found    to    be 


equivalent  to  the  following  one:  the  product  of 
the  capacity  units  purchased  in  year  k  and  the 
marginal  value  of  capacity  calculated  under  the 
assumption  of  the  lowest  net  revenue  occurring 
in  every  forthcoming  year  —  the  marginal  cost 
of  capacity  visualizing  the  worst  —  is  never 
greater  than  the  value  of  the  money  account  in 
year  k  —  1  plus  the  terminal  value  of  the  ca- 
pacity in  all  of  the  remaining  years  (with  the 
lowest  prices  and  smallest  catches  occurring) 
minus  any  fixed  cash  withdrawals  in  the  rest 
of  the  planning  period.  (All  money  flows  are 
adjusted  for  the  values  of  alternative  oppor- 
tunities, income  taxes,  and  depreciation.)  This 
upperbound  would  be  the  value  of  the  firm's 
assets  if  the  worst  possible  sequence  of  net 
revenues  occurred  —  the  decisionmaker's  final 
asset  position  visualizing  the  worst. 

To  reflect  the  fear  of  low  net  revenues,  revenue 
per  unit  of  capacity  when  the  lowest  price  and 
yield  occurs  is  assumed  to  be  less  than  the 
operating  cost  per  unit  of  capacity.  It  is  also 
assumed  that  per  unit  prices  of  capacity  are 
not  increasing  so  rapidly  that  operating  losses 
per  unit  may  be  covered  by  value  appreciation 
in  capacity.  (Speculation  is  never  a  sure  bet.) 
This  implies  that  the  marginal  cost  of  capacity 
visualizing  the  worst  is  positive.  Hence,  dividing 
the  lower  bound  for  the  firm's  final  asset  position 
by  this  positive  marginal  cost,  the  upper  bound 
for  a  survivable  purchase  of  capacity  in  a  given 
year  is  obtained.  This  represents  the  maximum 
amount  of  capacity  that  the  decisionmaker  can 
purchase  and  still  insure  survival  of  the  firm 
throughout  the  rest  of  the  decision  period.  It 
depends  upon  the  value  of  the  firm's  money 
account,  the  amount  of  capacity  owned,  and 
the  value  of  that  capacity  in  the  previous  year. 
This  upper  bound  function  in  year  k  is  denoted 
by  Hk(zk  - 1 ,  Vk-it  %k-i),  where  at  the  end 
of  the  k  —  lst  year  zk  _  /  is  the  cash  balance, 
yk  _  i  is  the  units  of  capacity  owned,  and  xk  _  y 
is  the  purchase  value  of  the  firm's  capacity.  The 
firm  is  in  debt  if  zk  /  is  negative  and  has 
savings  if  zk_1  is  positive. 

We  will  also  introduce  the  following  notation 
now;  St  is  the  units  of  capacity  purchased  at 
the  beginning  of  the  ith  year  (and  used  for  the 
first  time  in  year  i);  r,  is  the  operating  costs  per 
unit  of  capacity  in  year  i\  a,  is  the  per  unit  pur- 
chase price  of  capacity  before  the  beginning  of 
the  operating  season  in  year  i;  A,  is  the  cash  with- 


113 


drawal  in  year  i  for  sundry  expenses;  7  is  the 
interest  rate  paid  (or  received)  on  the  cash  ac- 
count z;  co,  is  the  unknown  revenue  per  unit  of 
capacity  in  the  ith  year;  N  is  the  number  of  years 
in  the  planning  period;  |3  is  the  fraction  of  the 
value  of  the  capacity  recoverable  at  the  end  of 
the  planning  period;  5  is  the  income  tax  rate; 
and  e  is  the  straightline  depreciation  fraction. 
Also  E  will  be  used  to  denote  the  mean  of  the 
random  variable  go,-;  and  L  will  be  used  to  denote 
the  smallest  possible  annual  net  revenue  having 
a  positive  probability  of  occuring.  The  symbol 
a,  is  used  to  denote  the  output  price  where  only 
the  yield  is  a  random  variable  in  the  application 
below. 

Using  the  above  development,  the  survival 
model  may  be  stated  as  follows: 

Maximize  E(zN  +  j3a/v  +  iy/v)  over  all  n-tuples 
of  functions  s,(coi ,  C02 ,  .  .  .,  to,_i  ),  i  =  1,  2,  .  .  ., 
N,  satisfying  the  difference  equations 


(1)  Xf—Xi-i  =OiSj,x0  =o0yo, 
where 

y(  —  y,_,  =  S{,  yo  given  and  non-negative, 

(2)  zt  —  zt-\  =  yzt-\  +  y,-  (co  —  77)  —  07  s/ 

A-  -  5  |y,  (co~ t{)  +  yzi-i  -  a,. 
—  exi],  e  =  0.091, 


where  z0  given,  and  i  =  1,2,. 
ing  the  inequalities 


.,  N,  and  satisfy - 


O^si^Hi  (z/_i ,  y,_, ,  x/_i ),  ;  =  1,2,..  .,  N. 

In  words,  the  decisionmaker  desires  to  maxi- 
mize expected  net  worth  at  the  end  of  the 
decision  period  where  the  purchases  of  capacity 
are  selected  from  the  survivable  set  in  each 
year  (delineated  by  the  inequality  restrictions). 
Thus,  in  the  maximization  process,  the  decision- 
maker, who  takes  into  account  all  of  the  informa- 
tion known  at  the  time  of  the  decision,  selects 
the  investment  from  the  survivable  set  of  capa- 
city purchases  that  maximizes  expected  net 
worth  at  the  end  of  the  planning  horizon. 


THE  DECISION  RULE  FOR  INVESTMENT 

By  the  use  of  dynamic  programming 
methods,  the  method  developed  by  Thompson 
and  George  was  extended,  as  mentioned  above, 
to  allow  for  depreciation  and  income  taxes.  The 
extended  rule  for  optimal  investments  is  sum- 
marized in  the  following  theorem. 

Theorem:  Suppose  H|(z0,  yo,  x0  )>0,  i.e.  the 
upper  bound  for  investments  in  the  first  year 
is  non-negative.  Let  Rk  be  the  expected  mar- 
ginal value  of  capacity  for  survival  investment 
decisions— the  marginal  value  of  capacity  vis- 
ualizing the  worst.  Then  the  decision  rule  for 
optimal  survivable  investment  is  as  follows: 

( 3 )    sk°  =  Hk  (zk°.  1 ,  yfc°_, ,  xfc°_, )  if  R , >0, 
and  sic  ~  o  if  Rk<0 

with  the  feasible  value  of  Sk  being  immaterial 
if  Rk  =0. 

In  other  words,  the  decisionmaker  buys  the 
survivable  limit  of  capacity  in  year  k  if  the 
marginal  value  of  capacity  visualizing  the  worst 
is  positive  in  that  year,  and  he  makes  no  capacity 
purchases  if  this  marginal  value  is  negative.  It 
also  follows  that  the  optimal  purchase  is  im- 
material in  any  year  (because  of  the  linearity 
of  the  problem)  whenever  the  decision  rule  is 
zero.  The  upper  bound  for  investments  in  the 
first  year  insures  the  existence  of  a  feasible 
investment  solution  in  each  year  of  the  planning 
horizon. 


An  Application  to  Shrimp  Fishing 

To  indicate  how  the  model  may  be  applied  to 
a  shrimp  fishing  firm,  parameters  were  specified 
for  a  relatively  small  fishing  firm  operating 
73-foot  steel  hull  trawlers  (see  Table  1).  In  the 
specifications,  the  values  of  the  parameters  were 
specified  to  reflect  prices,  costs,  and  landings  per 
vessel  as  reported  by  the  firms  cooperating  in 
the  study.  There  is  an  exception  with  regard  to 
Problem  3.  Average  landings  per  vessel  which 
were  found  to  be  57,560  pounds  of  heads-off 
shrimp  per  year  in  the  years  1958  through  1969 
were  specified  to  be  one  standard  deviation 
above  the  mean  to  evaluate  the  effect  of  better 
than  average  management.  That  is,  in  Problem 


114 


Table  1,  —  Values  of  the  parameters  for  four  survival  problems:  the  Gulf  shrimp  fishery. 
Parameters  Problems 


N   --  number  of  years  in  planning  period 

Z    ■-  initial  cash  balance  in  dollars 
o 

y    -  initial  number  of  boats  in  fleet 

x     -  initial  investment  in  dollars 
o 

J     -  annual  interest  rate  per  dollar 

T     -  annual  production  cost  per  vessel  in  dollars 

a     -  per  vessel  purchase  price  in  dollars 

e      -  annual  depreciation  fraction  per  dollar  invested 

t      -  annual  income  tax  rate  per  dollar  of  taxable 
income 

p      -  recoverable  fraction  of  investment  in  fishing 
capacity 


1 


5 

5 

5 

5 

0 

96,145 

0 

0 

1 

(t 

1 

1 

100,000 

(l 

100,000 

100,000 

0.085 

0.085 

0.085 

0.085 

30,000  x 

30,000  x 

30,000  x 

30,000  x 

(1.03)r 

(1.03)* 

(1.03)' 

(1.03)f 

100,000  x 

100,000  x 

100,000  x 

100,000  x 

(1.03)* 

(1.03)' 

(1.03/ 

(1.03)' 

.091 


.25 


.65 


.091 


.25 


.65 


.091 


.25 


.65 


.091 


.25 


.65 


annual  cash  withdrawal  for  sundry  expenses  in 

3,600  x 

3,600  x 

3,600  x 

3,600  x 

dollars 

(1.03)' 

(1.03/ 

(1.03)r 

(1.03/ 

owner's  expected  annual  revenue  per  vessel  in 

49,790  x 

49,790  x 

54,400  x 

49,790  x 

dollars 

p/1.03)' 

p/1.03/ 

p/1.03)* 

p/1.03)' 

owner's  lowest  annual  revenue  per  vessel  in 

dollars 

22,500  x 

22,500  x 

22,500  x 

22,500  x 

(1.03)f 

(1.03/ 

(1.03/ 

(1.03)' 

3  landings  per  vessel  were  63,291  pounds  of 
heads-off  shrimp  per  year. 

Since  the  real  price  of  shrimp  —  the  price 
adjusted  for  the  purchasing  power  of  money  — 
is  highly  influenced  by  growth  in  real  per 
capita  income,  and  since  it  appears  that  the 
economy  may  be  entering  a  period  of  modest 
growth  (possibly  much  like  the  late  1950's), 
the  real  price  of  shrimp  was  specified  to  reflect 
a  1.5%  rate  of  growth  in  real  per  capita  income 
in  Problems  1,  2,  and  3,  and  to  reflect  a  3.3% 
rate  of  growth  (as  observed  in  the  mid  1960's) 
in  Problem  4. 

To  evaluate  the  economic  attractiveness  of 
shrimp  fishing  versus  the  best  alternative  to 


fishing  (as  reflected  by  the  interest  rate  on 
money),  the  decisionmaker  in  Problem  2  initially 
has  the  approximate  money  equivalent  of  an 
investment  in  one  vessel.  Recall  that  the  en- 
trepreneur is  a  profit  maximizer,  given  that  he 
can  survive.  Thus,  the  decisionmaker  would 
opt  for  the  savings  alternative  whenever  the  net 
rate  of  return  from  a  dollar  invested  in  fishing 
capacity  is  less  than  the  interest  rate  on  money. 
That  is,  the  second  problem  indicates  the  eco- 
nomic advantage  (or  disadvantage)  of  investing 
in  fishing  relative  to  loaning  the  money  to 
someone  else. 

Since  the  model  takes  into  account  the  infor- 
mation obtained  through  time  as  the  values  of 


115 


Table  2.  —  Solutions  to  four  survival  problems  in  table  1;  landings  per  vessel  are  random. 


Marginal  value 

Investment 

Debt  to 

of  another  vessel 

in  boats 

Boats  owned 

Cash  balance 

gross  asset 

Problem 

Year 

(dollars) 

(number) 

(number) 

(dollars) 

ratio 

0 

_ 

_ 

1.00 

0 

1 

5,843 

1.44 

2.44 

-146,356 

.57 

2 

-      784 

0 

2.44 

-127,678 

.48 

1 

3 

-  7,896 

0 

2.44 

-116,862 

.43 

4 

-15,474 

0 

2.44 

-108,022 

.38 

5 

-23,490 

0 

2.44 

-  74,436 

.26 

0 

- 

- 

2.44 

96,145 

_ 

1 

5,843 

2.44 

2.44 

-145,083 

.57 

2 

-      784 

0 

2.44 

-126,507 

.48 

2 

3 

-  7,896 

0 

2.44 

-115,728 

.43 

4 

-15,474 

0 

2.44 

-106,908 

.38 

5 

-23,490 

0 

2.44 

-  73,534 

.26 

0 

— 

_ 

1.00 

0 

_ 

1 

21,419 

1.44 

2.44 

-136,487 

.53 

2 

16,198 

1.13 

3.57 

-216,534 

.56 

3 

3 

7,080 

4.03 

7.59 

-581,958 

.68 

4 

-  5,562 

0 

7.59 

-511,662 

.58 

5 

-18,570 

0 

7.59 

-358,977 

.40 

u 

_ 

_ 

1.00 

0 

_ 

1 

10,655 

1.44 

2.44 

-145,128 

.56 

2 

9,943 

.80 

3.23 

-240,502 

.58 

4 

3 

2,624 

3.15 

6.38 

-503,596 

.70 

4 

-  7,595 

0 

6.38 

-462,898 

.63 

5 

-19,119 

0 

6.38 

-341,999 

.45 

the  random  variables  are  revealed,  solutions  to 
two  sets  of  problems  were  computed.  In  the  first 
set,  the  landing  per  vessel  is  random;  whereas 
in  the  second  set,  the  price  received  is  random 
as  well.  The  first  set  of  results  is  presented  in 
Table  2,  and  the  second  set  in  Table  3. 

It  is  important  to  note  that  this  application 
of  the  survival  model  is  not  exhaustive  of  the 
many  that  could  be  made,  or  to  imply  that  the 
normative  results  presented  are  likely  to  occur. 
This  work  is  only  meant  to  indicate  how  an 
investor  interested  in  shrimp  fishing,  who  has 
a  limited  amount  of  money  capital,  might 
obtain  bench  marks  (from  the  model)  for  in- 
vestment planning. 


Values  of  the  Parameters 

In  this  application,  the  firm's  initial  fishing 
capacity  was  specified  to  be  one  vessel  in  Prob- 


lems 1,  3,  and  4.  The  values  of  the  data  (excluding 
the  basis  for  the  expected  shrimp  price  in  the 
first  set  of  problems)  are  given  in  Table  1.  The 
initial  purchase  price  of  one  vessel  was  taken 
to  be  $100,000.  In  Problems  1,  3,  and  4,  the 
firm  is  visualized  as  having  an  initial  debt-free 
investment  of  $100,000  with  no  savings.  This 
relatively  large  amount  of  initial  equity  was 
necessary  for  the  survival  problem  to  have  a 
feasible  solution.  The  minimum  value  for  the 
firm's  initial  equity  in  Problem  1  was  found 
to  be  $97,000. 

In  Problem  2,  where  the  entrepreneur  has 
his  equity  in  savings  rather  than  invested  in 
fishing  capacity,  the  initial  value  for  savings  is 
$96,145.  This  is  approximately  equivalent  to 
owning  one  vessel  initially  because  of  the  pro- 
cedure used  to  calculate  interest  earnings  and 
tax  allowances  in  the  model. 

There  is  only  one  money  account  in  the  model, 
and  accordingly  one  interest  rate.  This  rate  was 
specified  to  be  8V2%  per  year. 


116 


Table  3.  —  Solutions  to  two  survival  problems  in  table  1 ;  shrimp  prices  and  landings  per  vessel  are  random. 


Marginal  value 

Investment 

Debt  to 

of  another  vessel 

in  boats 

Boats  owned 

Cash  balance 

gross  asset 

Problem 

Year 

(dollars) 

(number) 

(number) 

(dollars) 

ratio 

0 

1.00 

0 

0 

1 

5,843 

1.44 

2.44 

-156,026 

.60 

2 

-      784 

0 

2.44 

-126,538 

.48 

1 

3 

-  7,896 

(l 

2.44 

-118,206 

.43 

4 

-15,474 

0 

2.44 

-118,517 

.42 

5 

-23,490 

0 

2.44 

-100,311 

U 

0 

_ 

_ 

1.00 

0 

— 

1 

21,419 

1.44 

2.44 

-147,856 

.57 

2 

16,198 

.69 

3.13 

-176,191 

.52 

3 

3 

7,080 

4.22 

7.34 

-569,170 

69 

4 

-  5,562 

0 

7.34 

-533,307 

.63 

5 

-18,570 

0 

7.34 

-438,263 

.50 

To  reflect  inflation,  the  purchase  price  of  new 
vessels  was  specified  to  increase  at  3%  per  year. 
This  rate  is  2%  below  reported  price  trends, 
which  include  costs  of  technological  improve- 
ments. Newer  vessels  have  been  powered  by 
larger  engines.  This  has  allowed  for  larger 
trawls  to  be  towed  at  faster  rates.  This  rate  of 
improvement  in  technology  is  believed  to  have 
increased  investment  costs  by  2%  per  year. 

From  the  cost  records  of  the  cooperating 
firms,  the  annual  cost  of  operating  a  73-foot 
trawler  was  found  to  be  $30,000  in  1969.  This 
cost  figure  includes  an  allowance  for  overhead 
and  insurance.  Representatives  of  the  firms 
interviewed  indicated  these  costs  have  increased 
3%  per  year  in  recent  years.  Thus,  the  annual 
production  cost  per  vessel,  rf,  was  specified  to 
be  30,000  (1.03)f. 

Straight  line  depreciation  methods  were  used 
for  tax  purposes  with  an  11-year  depreciation 
period  being  used  for  a  fully  outfitted  vessel. 
This  average  was  estimated  on  a  value  weighted 
basis  from  the  records  of  a  number  of  firms. 
The  reciprocal  of  this  figure,  0.091,  was  the 
value  used  in  the  depreciation  function. 

Income  is  the  sum  of  the  revenues  received 
(by  the  owner  after  the  "lay")  less  operating 
costs,  interest  costs  (or  plus  interest  earnings), 
and  taxes.  The  income  tax  rate,  which  is  denoted 
by  £,  was  taken  to  be  25%  of  the  taxable  income. 
This  rate  was  paid  in  the  late  60's  by  a  number 
of  the  firms  studied. 

In  shrimp  fishing,  the  captain  and  first  mate 
on  a  vessel  are  commonly  paid  on  a  "lay"  basis 
wherein   they   receive  for  services   rendered   a 


percentage  of  the  revenue  earned  by  the  vessel. 
The  header,  who  is  the  third  crew  member,  is 
typically  paid  on  a  per  box  basis;  his  wages  are 
included  in  the  production  cost  per  vessel.  For 
73-foot  vessels,  the  "lay"  for  the  captain  and 
first  mate  is  commonly  35%  (with  the  owner 
getting  in  effect  65%  of  the  ex-vessel  price) ;  they 
typically  pay  for  all  of  the  groceries. 

In  interviewing  the  cooperating  firms,  the 
relative  resale  value  of  the  vessels  sold  was 
found  to  be  fairly  well  approximated  (for  vessels 
5  to  6  years  old)  by  summing  the  accumulated 
depreciation  fractions  with  an  appropriate  ad- 
justment for  technological  improvement.  This 
procedure,  which  implies  that  the  resale  value 
of  a  vessel  5  years  old  would  be  65%  of  the 
purchase  price,  was  used  as  the  basis  for  specify- 
ing (3  to  be  equal  to  0.652. 

To  project  per  vessel  expected  revenue  re- 
ceived by  the  owner,  the  log  of  the  real  shrimp 
price  received  by  the  cooperating  firms,  Pt,  was 
regressed  on  the  log  of  the  index  of  real  per 
capita  income  (in  the  United  States  after  taxes), 
yt,  and  the  log  of  per  unit  effort  landings,  h, 
caught  in  depths  beyond  10  fathoms  off  the 
Texas  coast.  (See  the  earlier  study  by  Thompson 
et  al.  (1970,  p.  12)  for  data.)  The  estimated 
regression  equation  was  as  follows: 


(4)    In  pt  =  -4.571  +  1.175  In  yt  - 

(t  =  3.6) 

R2  =  .748,  oe  =  .0888. 


-.379  In  /,, 
(*=3.5) 


2  This  approximating  procedure  was  necessary,  since 
the  vintage  was  not  accounted  for  in  the  model. 


117 


Variations  in  landings  per  unit  effort,  which 
were  found  to  be  highly  correlated  for  the  Texas 
and  Gulf-South  Atlantic  fisheries,  are  still 
regarded  by  biologists  as  being  largely  random. 
Thus,  to  remove  the  effect  of  landings  on  price, 
landings  were  specified  to  be  equal  to  the  mean 
value  observed  for  the  Texas  fishery  in  the 
period  1958  through  1967.  Hence,  the  price 
estimating  equation  with  an  adjustment  to  a 

1969  base  year  was  as  shown  below. 

(5)  lnpf  =-1.332  +  1.175  In  yt 

To  use  this  equation,  the  index  of  real  per 
capita  income  had  to  be  projected  for  the  years 

1970  through  1974.  This  was  done  by  regressing 
In  yt  on  time,  t,  for  the  years  1953  through 
1960,  and  also  for  the  years  1961  through  1968. 
The  following  two  income  projection  equations 
were  developed  for  the  period  t  =  1970,  1971, 
.  .  .,  1974. 

Specification  I:  1.5%  rate  of  income  growth 

(6)  lnyr  =  4.94  +  .015* 

Specification  II:  3.3%  rate  of  income  growth 

(7)  lnyr  =4.94  +  .033? 

By  substituting  the  desired  specification 
from  (6)  into  (5),  the  price  projection  equation 
was  obtained.  The  effective  expected  real  shrimp 
price,  ar,  was  0.65  of  the  antilog  of  pt.  To  convert 
to  money  terms,  the  projected  prices  were 
multiplied  by  the  consumer  price  index  value 
for  1969,  1.277,  and  by  a  price  inflating  factor 
of  3%  per  year  thereafter.  In  Table  1,  fit  denotes 
the  price  reflecting  the  high  rate  of  income 
growth  and  pt  the  low  rate. 

For  the  first  set  of  four  problems,  the  estimate 
of  the  owner's  lowest  annual  revenue  per 
vessel,  Lr,  was  found  by  taking  the  lay  residual 
of  the  product  of  the  1969  shrimp  price,  a69, 
and  the  projected  lower  bound  for  landings  per 
vessel.  This  lower  bound  was  taken  to  be  3.4 
standard  deviations  (in  t  units  for  11  degrees 
of  freedom)  below  the  mean  landing  per  vessel 
of  57,560  pounds  with  the  sample  standard 
deviation  being  5,731  pounds.  Thus,  the  prob- 
ability of  the  landings  per  vessel  being  greater 
than  this  lower  bound  (assuming  this  to  be  a 
valid  probability  basis)  is  greater  than  0.99. 
Moreover,  since  the  growth  rate  in  real  per 
capita  income  is  not  taken  into  account  in  Lt, 


the  probability  of  revenue  per  vessel  falling 
below  the  implied  estimate  of  the  owner's  lowest 
annual  revenue  per  vessel  (where  the  price  is 
projected  under  either  specification)  decreases 
steadily  as  the  planning  period  unfolds.  In 
other  words,  the  estimate  of  L,  is  very  conserva- 
tive for  the  year  1970  and  becomes  increasingly 
conservative  thereafter  in  the  planning  period.3 

For  the  second  set  of  two  problems  in  which 
the  shrimp  price  is  random  as  well  as  the 
landing  per  vessel,  the  same  value  was  used  for 
the  owner's  lowest  annual  revenue  per  vessel. 
This  resulted  in  a  slightly  smaller  probability 
of  survival  than  in  the  first  four  problems 
(because  of  the  additional  randomness  in  the 
price),  but  one  still  greater  than  0.99.  Thus,  in 
the  interest  of  simplicity,  the  same  value  of 
Lt  was  used  in  both  sets  of  problems. 

Knowledgeable  industry  representatives  (who 
were  consulted  with  regard  to  the  above  specifi- 
cations) indicated  a  5-year  survival  period 
would  be  especially  meaningful  for  firms  operat- 
ing the  73-foot  trawlers.  Accordingly,  two  5- 
year  sequences  of  random  revenues  per  vessel 
were  developed  with  only  the  landing  per  vessel 
being  random  in  the  first  sequence.  Landings 
per  vessel  were  regarded  as  independent  of 
price,  since  the  fishery  is  relatively  competitive; 
moreover,  for  the  period  studied,  per  vessel 
landings  for  the  cooperating  firms  were  not 
highly  correlated  with  landings  per  unit  of 
effort  in  the  Texas  fishery4  (r2  =  0.16).  Using 
the  regression  estimate  for  price  in  each  year 
1970  through  1974  and  the  estimated  standard 
error  of  the  regression,  and  also  using  the  sample 
mean  and  standard  deviation  for  landings  per 
vessel  of  the  cooperating  firms,  the  random 
prices  and  landings  per  vessels  were  calculated 
as  follows:  (1)  By  use  of  the  Box-Muller  (1958) 
method,  normal  random  deviates  for  prices  and 
landings  per  vessel  were  independently  gener- 
ated ;  and  (2)  the  products  of  these  two  random 
variables  were  adjusted  for  the  lay  and  expected 
changes  in  the  purchasing  power  of  money.  The 
following  random  sequences  were  accordingly 
obtained  and  used  in  the  analysis. 


3  To  have  a  probability  support  at  Lf,  this  small 
probability  of  non-survival  is  implicitly  assumed  to  be 
insurable. 

4  Landings  per  unit  effort  in  the  Texas  Fishery  are 
highly  correlated  with  those  for  the  Gulf  and  South 
Atlantic. 


118 


Random  Sequences  of  Revenues  per  Vessel 


Seq 

xence  No.  1 

Problems  1  &  2 

Problem  3 

Problem  U 

$30,741 

$36,141 

$31,413 

42,572 

48,233 

44,457 

39,859 

45,795 

42,531 

39,797 

46,020 

43,393 

50,784 

57,308 

56,583 

Sequence  No.  2 
Problem  1  Problem  3 


$25,450 
47,261 
38,810 
36,077 

44,747 


$29,920 
53,546 
44,589 
41,719 
50,495 


It  may  be  helpful  to  recall  that  the  decision- 
maker is  regarded  as  being  a  better  than  average 
manager  in  Problem  3.  The  1.5%  rate  of  real 
economic  growth  per  capita  is  used  in  Problems 
1,  2,  and  3;  and  the  3.3%  rate  of  economic 
growth  is  used  in  Problem  4. 

In  evaluating  the  solutions  to  the  first  set  of 
four  problems  in  Table  2,  the  results  indicate 
the  profitability  of  investing  in  shrimp  fishing 
capacity  during  the  5-year  planning  period. 
The  model  fisherman  opted  for  investing  in 
fishing  capacity  in  Problem  2,  even  though  he 
could  have  left  his  money  in  savings  at  8.5% 
interest.  Thus,  the  rate  of  return  over  cost  from 
shrimp  fishing  was  greater  than  8.5% .  In  further 
analysis,  it  was  found  to  continue  to  be  so  until 
the  rate  of  interest  reached  9.5%  ;  then  the  rate 
of  return  over  cost  switched  in  favor  of  savings. 

The  value  of  better  than  average  management 
is  indicated  by  the  results  in  Problem  3.  There, 
the  average  landing  per  vessel  was  taken  to  be 
one  standard  deviation  (5,731  pounds)  greater 
than  in  Problem  1.  The  same  amount  was 
invested  in  the  first  year;  but  in  the  second  and 
third  years  there  were  striking  differences.  The 
model  fisherman  bought  5.2  vessels  in  Problem 
3,  while  he  did  not  buy  any  in  Problem  1.  He 
chose  to  pay  off  debt  in  the  first  problem  after 
the  initial  investment,  since  that  represented 
a  more  profitable  use  of  the  money.  It  may  be 
noticed  that  the  investment  upper  bound  limited 
the  size  of  the  purchases  in  the  first  3  years  of 


Problem  3  (and  the  first  year  of  Problem  1). 
The  marginal  value  of  another  vessel  was 
positive;  however,  the  money  was  not  available 
for  investment  given  the  desire  to  survive. 

Success  in  shrimp  fishing  is  clearly  influenced 
by  the  rate  of  income  growth  in  the  economy  — 
compare  Problems  1  and  4.  In  Problem  4,  the 
marginal  value  of  another  vessel  is  almost 
twice  as  large  in  the  first  year  as  in  Problem  1, 
and  remains  large  in  the  second  year  when  the 
value  in  the  first  problem  goes  negative.  This 
increased  growth  in  per  capita  income  results 
in  an  increased  ability  to  invest  in  the  second 
year  in  Problem  4  and  still  further  increased 
ability,  at  a  lower  marginal  incentive,  in  the 
third  year.  The  model  fisherman  carries  a  con- 
siderably larger  debt  load,  as  a  result  of  the 
increased  profitability,  in  Problem  4  than  in 
Problem  1. 

In  evaluating  the  second  set  of  results  given 
in  Table  3  and  comparing  these  solutions  to 
the  ones  in  Table  2,  only  slight  differences 
between  the  results  may  be  noticed.  Somewhat 
less  is  invested  over  the  planning  period  in 
Problem  3  in  the  second  case  than  in  the  first. 
Also,  a  slightly  larger  debt  load  was  generally 
carried  in  most  of  the  planning  period.  Of 
course,  the  marginal  investment  incentives 
were  the  same  in  both  sets  of  problems;  they 
are  based  on  expected  values.  Vagaries  in  land- 
ings seem  to  be  much  more  important  than 
unexpected  variations  in  price. 


LITERATURE  CITED 


THOMPSON,  RUSSELL  G.,  and  MELVIN  D.  GEORGE. 
1970.  A  Stochastic  Investment  Model  for  Survival 
Conscious  Firm.  Presented  at  Winter  Meetings  of  the 
Econometric  Society,  Detroit,  December,  1970. 

THOMPSON,  RUSSELL  G.,  RICHARD  W.  CALLEN, 
and  LAWRENCE  C.  WOLKEN.  1970.  Optimal  Invest- 
ment and  Borrowing  Decisions  for  a  Model  Shrimp 
Fishing  Firm,  Texas  A&M  University,  Sea  Grant 
Bulletin  No.  205,  April. 

BOX,  G.  E.  P.,  and  MERVIN  E.  MULLER.  1958.  A 
Note  on  the  Generation  of  Random  Normal  Deviates, 
The  Annals  of  Mathematical  Statistics,  Vol.  29,  June, 
pp.  610-611. 


119 


APPENDIX 


Appendix  Table   1.  —  Values  of  projected  index  of  real  Appendix   Table    2.   —  Values  of  projected  real  shrimp 

per  capita  income.  prices. 


Year                          /Specification  1                            Specification  II  Specification  I,  p  Specification  II. p 
Year  (cents  per  pound)  (cents  per  pound) 

1  136.98  139.52 

2  139.06  144.27  1  85.68  87.56 

3  141.17  149.19  2  87.22  91.07 

4  14332  154^7  3  88-78  94-73 


5  145.50  159.53 


4  90.37  98.53 

5  91.99  102.49 


Appendix  Table  3.  —  Values  of  landings  per  vessel  for 
random  sequences  1  and  2. 

Problems  1 ,  2  &  4  Problem  3 

Year  (pounds)  (pounds) 

1  41,965  49,336 

2  55,435  62,806 

3  49,501  56,872 

4  47,140  54,511 

5  57,375  64,746 


120 


Simulation  Experiments  to  Evaluate  Alternative 
Hunting  Strategies  for  a  Deer  Population1 

F.  M.  Anderson,2  G.  E.  Connolly3 
A.  N.  Halter,2  and  W.  M.  Longhurst3 

ABSTRACT 

A  population  dynamics  model  of  the  deer  herd  in  Mendocino  County,  California, 
is  presented.  Environmental  influences  are  modeled  as  density  dependent  birth  and 
death  rate  functions.  The  computer  program  for  this  biomanagement  model  is  outlined 
and  validity  checks  devised  to  improve  the  model  are  discussed.  The  output  shows 
the  impact  of  selected  hunting  strategies  on  productivity,  natural  mortality,  and 
other  population  characteristics.  Tests  of  hunting  strategies  related  to  alternative 
management  goals  are  summarized.  Implications  of  computer  simulation  methodology 
for  the  management  of  wildlife  and  fish  populations  are  discussed. 


INTRODUCTION 

Management  of  a  natural  resource,  such  as 
a  deer  herd  or  fishery,  is  the  manipulation  of 
that  resource  and/or  its  environment  in  an 
attempt  to  satisfy  a  set  of  objectives.  The 
Objectives  can  be  economic  or  noneconomic. 
They  may  or  may  not  be  quantifiable,  and 
hence,  the  management  problem  may  or  may 
not  be  solvable  in  the  framework  of  "extremum" 
problems. 

The  management  of  a  deer  herd,  like  that 
of  a  fishery,  can  be  directed  toward  multiple 
objectives.  The  deer  herd  may  be  maintained 
at  a  particular  level  and  age  composition  to 
achieve  a  hunting  kill  having  the  greatest 
value;  alternatively,  the  herd  may  be  main- 
tained for  purely  aesthetic  reasons.  A  multiple 
objective  of  management  may  be  to  sustain 
a  certain  deer  density  (deer  per  square  mile) 
at  one  time  of  the  year  to  provide  hunting, 
or  at  another  time  of  the  year  to  provide 
scenery  for  sightseers. 

Under  certain  environmental  conditions, 
managers  may  be  prevented  from  knowing 
whether  or  not  the  objective(s)  has  (have) 
been  attained.  In  areas  of  dense  ground  cover, 


1  Technical    Paper    Number    2998,    Oregon    Agricul- 
tural Experiment  Station,  Corvallis,  Oregon. 

2  Department     of    Agricultural     Economics,     Oregon 
State  University,  Corvallis,  Oregon. 

3  Department    of    Agricultural     Zoology,    University 
of  California,  Davis,  California. 


managers  must  often  resort  to  crude  sampling 
techniques  to  derive  population  estimates. 
Other  parameters  can  be  readily  measured. 
For  example,  in  a  deer  herd  where  hunting 
is  done  only  by  license,  the  kill  figures  are 
available  soon  after  the  hunting  season,  and 
can  be  used  in  the  formulation  of  subsequent 
management  strategies.  It  may  be  that  certain 
objectives  will  be  satisfied  if  crucial  parameter 
values  are  between  certain  upper  and  lower 
bounds.  Alternatively,  the  objective  of  man- 
agement may  be  to  maximize  the  value  of  a 
parameter.  Examples  of  these  two  cases  are 
(1)  to  keep  the  average  size  of  the  herd  be- 
tween two  values,  and  (2)  to  maximize  the 
annual  hunter  kill,  respectively.  Other  paral- 
lels to  the  objectives  of  management  for  a 
deer  herd  can  be  found  in  the  management 
of  a  fishery  resource. 

Both  deer  and  fish  populations  are  members 
of  complex,  dynamic  ecosystems.  For  each, 
the  age  composition  changes  over  time  due  to 
the  changes  of  such  parameters  as  birth  rates 
and  death  rates.  In  addition  to  relatively 
simple  variability  about  these  parameters, 
changes  in  the  population  are  compounded  by 
environmental  changes. 

To  illustrate,  assume  there  is  a  functional 
relationship  between  deer  density  and  the 
mortality  rate  of  each  age  category.  Further- 
more, assume  a  fixed  habitat  structure  and 
that  variability  in  the  biosystem  is  introduced 
only  by  changes  in  the  weather.  The  effect  of 
these   changes   will   usually   be   lagged.   Other 


121 


relationships  can  be  hypothesized  to  complete 
the  abstract  model.  For  each  time  period,  the 
mortality  rate  in  each  age  category  depends 
upon  the  density.  Over  time  this  density  will 
change,  as  will  the  inventory  of  deer  in  each 
category.  Hence,  mortality  rates  will  differ 
over  time,  even  if,  the  same  functional  re- 
lationships are  hypothesized. 

Now,  add  in  the  complicating  factor  of 
changes  in  some  or  all  of  these  mortality  func- 
tions consistent  with  an  improved  habitat  and 
a  higher  plane  of  nutrition.  In  the  real  world, 
changes  in  the  deer  habitat  —  and  its  counter- 
parts in  other  fish  and  wildlife  species  —  are 
occurring  continuously. 

In  making  management  decisions,  some 
knowledge  is  assumed  of  the  structure  of  the 
relevant  biosystem.  However,  knowledge  is, 
at  best,  uncertain,  and  heroic  assumptions 
are  aften  made  about  the  effect  of  a  structural 
change.  Thus,  decisions  may  be  made  which 
move  the  biosystem  toward  the  objectives 
desired  in  an  unpredictable  manner.  Manage- 
ment is  usually  carried  out  within  the 
boundaries  described  by  legally  authorized 
regulations,  which  are,  hopefully,  both  con- 
sistent with  a  set  of  objectives  and  flexible 
enough  to  afford  the  on-the-spot  manager  some 
discretionary  action.  When  regulations  are 
for  more  than  one  distinct  resource  unit  this 
flexibility  is  desirable  because  each  unit  is 
unique. 

For  example,  regulations  for  deer  hunting 
in  a  particular  state  usually  embrace  more 
than  one  herd.  No  two  herds  will  be  identical 
at  any  point  in  time,  and  the  regulations  must 
be  sufficiently  flexible  to  allow  for  these  dif- 
ferences. Regulations  are  ideally  formulated 
with  regard  to  the  structure  of  the  relevant 
biosystems,  but  knowledge  of  these  biosystems 
is  not  complete.  The  response  of  the  biosystems 
to  particular  management  actions  cannot  be 
predicted  with  certainty.  Therefore,  there  is 
a  limit  to  the  rigidity  of  the  regulations.  Be- 
yond this  limit,  management  will  be  ineffec- 
tive in  attempts  to  satisfy  the  set  of  objectives. 

Thus  far,  we  have  briefly  described  three  com- 
ponents of  the  management  system  of  a  public 
resource.  These  are  the  complex  biosystem, 
the  set  of  objectives,  and  the  set  of  regulations 
relating  to  the  particular  resource.  One  more 
component  is   necessary   to  complete  a  work- 


able management  system;  that  is,  a  means  of 
monitoring  the  system  is  required.  For  any 
biosystem,  the  selection  of  the  parameters  to 
be  monitored  is  the  result  of  experience  and 
expertise.  However,  to  be  useful  to  manage- 
ment, the  selected  parameters  must  be  indica- 
tive of  the  performance  of  the  biosystem  so  that 
it  can  be  determined  whether,  or  to  what 
extent,  objectives  are  being  accomplished. 

Typically,  only  relatively  few  parameters 
can  be  monitored  accurately  and  rapidly  enough 
to  be  useful.  Information  on  the  state  of  the 
system  is  of  most  value  when  it  is  current. 
The  role  of  time  in  monitoring  systems  cannot 
be  overemphasized.  Information  on  the  state 
of  a  biosystem  at  any  time  is  usually  incomplete. 
For  example,  the  total  number  of  deer  in  a 
herd  is  a  useful  parameter  in  developing  man- 
agement strategies.  In  most  herds  it  is  im- 
possible to  take  an  accurate  annual  census, 
and  estimates  of  the  total  population  must 
be  based  on  samples,  which  often  may  be  col- 
lected only  at  certain  times  of  the  year. 

Historically,  researchers  and  managers  have 
been  restricted  to  experimentation  on  the  real 
biosystem.  However,  with  the  advent  of  com- 
puters and  programming  languages,  it  is  now 
feasible  to  perform  simulated  experiments  on 
biosystems  that  can  be  described  by  mathe- 
matical equations.  This  paper  is  concerned 
with  the  computer  simulation  of  the  deer 
population  in  Mendocino  County,  California. 
The  model  shows  the  population  dynamics 
and  some  of  the  economic  and  recreational 
consequences  associated  with  various  hunting 
strategies. 

COMPUTER  SIMULATION  METHODOLOGY 

Simulation  involves  building  and  operating 
a  model  designed  to  represent  those  features 
of  the  real  system  under  study  and  to  provide 
information  about  the  performance  of  the 
system  under  assumed  controlled  conditions. 

Three  classes  of  simulation  models  can  be 
distinguished:  (1)  physical  models,  such  as 
scale  models  of  river  systems  and  planetar- 
iums,  (2)  mathematical  models  where  a  set  of 
equations  describing  the  system  under  study 
is  written  and  these  equations  are  solved,  per- 
haps analytically,  and  (3)  computer  simula- 
tion  where   the   system   is   described   and   the 


122 


logic  is  programmed  for  computer  calculation. 
In  the  latter  case,  the  intent  is  to  simulate 
complex  systems  which  usually  involve  non- 
linear relationships,  random  components,  and 
time  varying  events. 

Computer  simulations  are  applicable  to  prob- 
lems of  the  type  where  management  can  influ- 
ence the  system's  behavior.  The  purpose  of 
simulating  a  management  system  is  to  test 
the  impact  on  variables  of  interest  within  par- 
ticular management  policies,  before  such 
policies  are  implemented,  and  influence  the 
real  system.  Here,  the  simulation  performs  the 
important  function  of  providing  information 
about  the  possible  consequences  over  time  of 
various  alternative  management  policies.  Thus, 
it  provides  answers  to  the  managers'  questions 
which  are  of  an  if-then  type.  The  computer 
program  is  an  if-then  calculator.  Systems 
could  be  simulated  using  paper  and  pencil,  but 
computers  can  carry  out  these  calculations 
more  efficiently. 

Simulation  should  be  viewed  as  an  iterative 
problem-solving  technique  which  involves  four 
stages:  (1)  problem  definition,  (2)  mathe- 
matical modeling,  (3)  refinement  and  testing 
of  the  resulting  model,  and  (4)  creative  design 
and  execution  of  simulation  experiments  to 
provide  information  relevant  to  the  manage- 
ment problem.  In  Figure  1,  arrows  indicate 
that  the  general  sequence  is  from  problem 
definition  to  application,  but  the  reverse  arrows 
indicate  that  the  process  is  iterative,  or  learn- 
ing in  nature.  A  prior  stage  might  have  to 
be  repeated  on  the  basis  of  information  acquir- 
ed during  a  subsequent  stage  of  the  modeling 
process. 

Problem  definition  is  fundamental  to  build- 
ing a  simulation  model.  This  study's  inter- 
disciplinary team,  composed  of  biologists  and 
agricultural  systems  analysts,  initially  met 
to  determine  the  types  of  questions  the  model 
was  to  answer.  The  questions  fell  into  three 
categories: 

1.  Biological  questions  involving  the  dynam- 
ics of  the  deer  population. 

2.  Economic  questions  involving  the  value 
or  worth  of  certain  events  and  occurrences 
within  the  system. 

3.  Management  questions  which  affect  the 
biological  system  and  have  economic  and 
social  consequences. 


Problem 

Definition 

> 

Mathematical 

Modeling 

&  Simulation 

IP 

Model 
Refinement 
&  Testing 

1 

" 

Model 

Application 

1 

¥ 

Output 


Figure     1.    —    Computer    simulation    as    an    iterative 
problem-solving  process. 


In  its  present  form,  the  model  construction 
cuts  across  all  three  types  of  questions,  and 
should  be  viewed  as  the  first  generation  model 
of  a  sequence  of  models  which,  hopefully,  will 
be  able  to  answer  these  questions  at  more 
sophisticated  levels.  This  first  generation  model 
is  essentially  a  population  simulator  capable 
of  answering  questions  mainly  of  a  biological 
nature,  but  provides  output  for  management 
questions  —  in  particular,  hunting  strategies. 
Other  sections  of  the  output  could  easily  be 
given  economic  interpretation.  The  second 
generation  model  will  include  economic  vari- 
ables such  as  losses  due  to  deer  damage  to 
agricultural  and  forest  lands,  and  gains,  such 
as  hunter  expenditure  and  the  value  of  venison. 
The  proposed  third  generation  model  will  in- 
clude a  management  component  which  would 
be  capable  of  evaluating  management  strate- 


123 


gies  in  the  broader  context  of  their  biological, 
economic,  and  social  consequences. 


DEER  HERD  SIMULATION  MODEL 

A  comprehensive  flow  chart  of  the  compo- 
nents and  interrelationships  of  a  deer  herd  was 
developed.  The  available  data  did  not  permit 
all  relationships  to  be  quantified  and  proxy 
variables  were  devised  to  overcome  this  diffi- 
culty. For  other  relationships  a  complete  speci- 
fication of  the  biological  interactions  would 
have  been  possible,  but  this  would  have  result- 
ed in  a  model  of  substantial  complexity.  Model 
building  is  a  continual  compromise  between 
abstraction  and  complexity.  Models  which 
are  too  abstract  are  devoid  of  interest,  and  the 
results  will  not  be  easily  related  to  the  oper- 
ations of  the  real  system.  When  the  models 
are    large,    and    incorporate    complex    mathe- 


matical formulations,  it  can  be  difficult  to 
extract  meaningful  guidelines  for  management. 
Such  models  may  be  expensive  to  run,  and 
thereby  not  achieve  one  objective  of  the  model- 
ing process,  namely,  to  simulate  the  systems 
and  generate  information  and  knowledge  about 
the  systems  at  a  cost  less  than  alternative 
analytical  techniques. 

The  flows  and  relationships  identified  for 
the  Mendocino  County  deer  herd  are  sum- 
marized in  Figure  2.  In  this  figure  the  time 
series  of  events  is  not  obvious.  These  are  dis- 
cussed in  more  detail  later  in  the  paper.  The 
model  as  depicted  in  Figure  2  is  best  viewed 
as  a  summary  of  the  most  pertinent  inter- 
actions which  occur  each  year  in  the  deer 
biomanagement  system.  The  basic  components 
of  the  system  are  the  birth  and  death  process. 
Each  year  fawns  are  born  into  the  herd,  and 
the  number  of  fawns  born  is  a  function  of  the 
exponential  average  of  the  density  in  particular 


Legend 


—    Weather 


Predators 

and 

Disease 


«r 


==dfc 


Accidents  and 
unclaimed 
hunter  kill 


Competition 
from  wildlife 
and  domestic 
animals 


Functional 
Relationships 
Causal  Relationships 
Information  Flows 
Real  Flows  (Deer) 


Other  factors, 

geophysical, 

destruct 


-     -1 

Regulation 

and 

Management 

-» 

Feed 
Conditions 

Figure  2.  —  Biomanagement  system  of  a  deer  population. 
124 


months  prior  to  the  time  of  birth.4  Thus,  the 
exponential  average  density  is  a  proxy  vari- 
able which  summarizes  all  relevant  causal 
influences  of  the  real  system.  The  casual  in- 
fluences are  indicated  in  Figure  2,  but  are  not 
explicitly  programmed  into  the  computer. 

In  the  model,  losses  are  defined  as  either 
natural  or  due  to  hunting.  Natural  losses  are 
the  residual  of  losses  after  accounting  for  the 
recorded  hunter  kill.  The  natural  losses  will 
include  those  due  to  age,  the  plane  of  nutrition, 
the  action  of  predators,  disease,  and  accidents 
on  the  highways.  Both  natural  and  hunting 
losses  are  computed  each  time  period.  Natural 
losses  are  computed  for  each  category  of  deer 
by  reference  to  functions  relating  the  density 
of  deer  at  the  beginning  of  the  period  to  the 
rate  of  mortality.  Here,  density  is  the  proxy 
variable  for  an  array  of  causal  relationships, 
as  indicated  by  Figure  2.  These  natural  mor- 
tality functions  were  based  upon  biological 
theory  and  the  available  empirical  evidence. 
The  paucity  of  data,  however,  precluded  sta- 
tistical estimation;  hence,  use  was  made  of 
interpolation  techniques  between  data  points 
to  derive  the  mortality  rates  for  particular 
densities.  Natural  losses  are  therefore  endog- 
enous to  the  model. 

Hunting  losses  are  treated  differently.  The 
hunting  loss  rates  are  defined  by  age  category 
and  the  time  period  in  which  hunting  is  allow- 
ed, as  specified  prior  to  the  execution  of  a  com- 
puter run.  The  hunting  losses  could  be  made 
endogenous,  but  in  the  first  generation  model, 
where  accent  is  on  formulating  a  reasonable 
biological  model,  it  is  advantageous  to  man- 
ipulate these  losses  to  test  the  model.  In  the 
real  world,  hunting  strategies  are  fomulated 
cognizant  of  political  considerations,  regula- 
tions, management  capability,  and  the  demand 
for    hunting.    They    are    the    consequences    of 


4  The    exponential    average    density    each    month    is 
computed  as  follows: 

EADt  =  EADt_  ,  +  1/T  (D,  -  EAD,    ,  ) 


interactions  which  are  not  fully  indicated  by 
Figure  2. 

Thus  far,  the  model  has  been  presented  as 
deterministic.  The  real  world  is  characterized 
by  random  variability.  The  response  of  the 
deer  biosystem  to  a  particular  set  of  conditions 
is  variable,  due  to  random,  uncontrollable 
elements  such  as  the  weather  conditions.  Ran- 
domness must  be  accounted  for  in  any  simu- 
lation which  purports  to  model  reality. 

In  the  deer  model,  a  random  number  gener- 
ator is  used  to  generate  variability.5  Vari- 
ability is  due  to  weather  conditions  which  are 
assumed  to  result  in  particular  forage  quality- 
quantity  relationships  or  forage  conditions. 
The  notion  of  a  forage  factor  is  used  as  an 
index  of  forage  conditions.  Each  year,  a  random 
number  is  computed  which,  in  turn,  implies 
a  particular  forage  factor.  Only  five  forage 
conditions  are  identified.  A  forage  factor  of 
five  corresponds  to  average  conditions;  and 
a  forage  factor  of  one  corresponds  to  poor  con- 
ditions. Forage  factors  of  two  and  four  cor- 
respond to  below  and  average  conditions,  res- 
pectively. 

The  probability  distribution  of  forage  factors 
can  be  easily  modified,  consistent  with  the 
investigation  of  the  impact  of  changes  in  the 
pattern  of  forage  conditions  over  time.  Once 
the  forage  factor  is  selected  for  the  year,  it  is 
used  to  modify  the  components  of  the  system 
which  are  considered  to  be  subject  to  vari- 
ability due  to  changes  in  the  forage  conditions 
—  namely,  natural  mortality  rates  and  birth 
rates.  The  notion  of  the  forage  factor  has 
proved  most  useful  in  the  development  of  the 
computer  model,  in  addition  to  its  primary  role 
in  carrying  out  experiments  with  the  model 
after  development. 

Thus,  the  biomanagement  system  is  present- 
ed as  a  network  of  flows,  rates,  and  levels.  The 
system  being  modeled  is  complex,  but  by  suit- 
able abstraction,  a  workable  dynamic  model 
which  permits  examination  of  the  system  in 
a  manner  not  permitted  by  the  usual  compara- 
tive statics  formulation,  can  be  developed. 


where:     t  =    Time  period  (month) 

EAD     =    Exponential    average    density    (deer/ 

square  mile) 
D  =    Density  (deer/square  mile) 

T  =    Exponential     smoothing     time     con- 

stant (number  of  months) 


5  The  computer  program  generates  a  sequence  of 
pseudo-random  numbers  which  provides  the  facilities 
for  comparing  results  of  different  runs  under  identical 
simulated  conditions. 


125 


Time  Sequence  of  Events 

A  flow  chart  of  the  computer  program  of 
the  deer  herd  is  shown  in  Figure  3.  For  any 
simulation  model  concerned  with  the  flow  of 
variables  over  time,  a  unit  of  time  must  be 
defined  for  purposes  of  calculation.  The  com- 
puter moves  in  discrete  steps  through  time, 
and  calculates  the  variables  at  each  step.  In 
the  deer  herd  model,  the  unit  of  time  is  one 
month.  For  each  month  of  a  computer  run, 
the  relevant  calculations  are  made,  and  the 
status  of  the  system  at  the  end  of  that  month 
is  generated.  The  status  of  the  system  is  an 
array  of  rates  and  levels  for  all  variables  in 
the  system.  The  time  counter  is  advanced  one 
unit  (one  month)  and  the  appropriate  calcu- 
lations for  that  month  are  made.  Calculations 
can  be  made  conditional  upon  any  event  or 
series  of  events  in  the  past,  but  not  upon 
future  events,  because  they  have  not  occurred. 

Starting  with  the  opening  inventory  shown 
at  the  top  of  Figure  3,  the  computer  program 
selects  a  forage  factor  for  the  year  as  of  Novem- 
ber 1,  and  computes  natural  losses  as  a  con- 
sequence of  the  forage  factor  and  deer  density. 
Figure  2  shows  the  array  of  interactions  which 
are  summarized  in  the  mortality  rate-density 
functions.  The  mortality  rate  in  each  age  and 
sex  class  is  described  as  an  exponentially  in- 
creasing function  of  density.  Hunting  losses 
are  then  computed  in  accordance  with  the 
hunting  strategy  specified  for  the  simulation 
run,  and  the  closing  inventory  by  age  and  sex 
is  calculated.  Loss  totals  are  then  accumulated, 
and  can  be  included  in  the  output  as  desired 
by  the  analyst.  Each  month,  the  above  sequence 
of  events  is  carried  out. 

After  accumulating  losses  in  May,  the  num- 
ber of  new  fawns  to  be  introduced  into  the 
herd  is  computed.  The  birth  rate  in  each  age 
class  of  does  is  described  as  a  decreasing  func- 
tion of  the  exponential  average  density.  The 
age  categories  are  then  advanced  one  year. 
Bucks  and  does  in  their  sixteenth  year  are 
removed  from  the  system  —  represented  in 
Figure  2  by  the  sink.  Fawns  born  12  months 
previously  are  separated  into  bucks  and  does, 
and  redefined  as  deer  in  their  second  year. 

Two  accounting  years  are  defined  in  the 
computer  program.  The  first  is  from  November 
1  to  October  31.  November  1  is  the  time  when 


managers  are  best  able  to  make  population 
counts  indicative  of  the  age  and  sex  composi- 
tion of  the  herd.  The  second  accounting  year 
used  in  the  model,  July  1  to  June  30,  facilitates 
the  summarization  of  the  hunting  results  for 
each  year.  Selected  parameters  are  printed 
at  the  end  of  each  accounting  year.  After  all 
the  October  operations  are  performed,  the 
year  counter  is  advanced  and  the  simulation 
proceeds  until  the  specified  number  of  years 
has  been  executed.  At  the  end  of  each  run,  sum- 
mary statistics  are  printed. 

Input  Data 

The  model  is  intended  to  simulate  the  Men- 
docino County  deer  herd,  but  the  primary 
data  source  was  the  University  of  California 
Field  Station  at  Hopland,  where  the  deer 
population  has  been  under  continuous  and  in- 
tensive study  since  1951.  The  investigators 
at  Hopland  compiled  these  data  and  integrated 
them  with  the  California  Fish  and  Game  De- 
partment data  for  the  remainder  of  the  county. 

Data  input  for  each  run  is  separate  from 
the  computer  program.  This  permits  changes 
in  the  data  assumptions  to  be  made  without 
altering  the  computer  program.  The  program 
is  designed  to  be  applicable,  with  minor  modi- 
fication, to  other  big  game  populations. 

The  data  block  for  each  run  includes  constants 
to  initialize  the  run,  such  as  the  opening  in- 
ventory, the  area  of  land  available  to  the  herd, 
and  the  length  of  the  run.  Other  data  used  in 
each  year  include  birth  rate  and  natural  mor- 
tality functions,  hunting  loss  percentages,  and 
the  distribution  of  forage  factors. 

HUNTING  STRATEGY  RESULTS 

While  an  infinite  variety  of  hunting  strate- 
gies can  be  tested  in  this  moi..el,  the  options 
of  the  wildlife  manager  are  limited  because 
certain  hunting  strategies  that  are  biologically 
feasible  may  be  socially  or  politically  unaccept- 
able. In  addition,  hunters  can  usually  dis- 
tinguish only  a  few  age  and  sex  classes  in  the 
field.  Limited  hunter  access  to  extensive 
areas  of  private  forest  and  range  lands  pre- 
cludes the  achievement  of  uniform  hunting 
pressure  over  the  entire  county. 


126 


Figure  3.  —  Flow  chart  of  the  computer  program  of  the  deer 

herd. 


The  hunting  strategies  summarized  in  Table 
1  include  the  range  of  options  which  could  be 
practically  implemented  in  Mendocino  County. 
Two  kinds  of  population  parameters  are  shown: 
those  which  can  be  maximized  or  minimized 
as  management  goals,  and  those  comparable 
with  field  data  to  determine  whether  manage- 
ment goals  are  being  achieved.  Some 
parameters,  such  as  the  hunting  kill,  serve 
both  purposes.  The  current  program  prints 
out  many  other  parameters  in  addition  to  those 
presented  in  Table  1. 

Although   it   is   physically   possible  to  hunt 


deer  at  any  time  of  the  year,  in  California  it 
is  customary  to  set  the  deer  seasons  in  late 
summer  and  fall,  for  numerous  biological  and 
sociological  reasons.  In  the  simulation  runs 
presented  in  Table  1,  all  buck  hunting  was  con- 
ducted during  August  and  September,  in  ac- 
cordance with  existing  custom,  and  potential 
doe  and  fawn  hunts  were  set  for  November 
and  December,  the  months  when  antlerless 
deer  are  in  the  best  condition.  All  parameters 
other  than  hunting  specifications  were  held 
constant  throughout  these  runs,  and  the  values 
shown    were    selected    from    the    output    after 


127 


Table  1.  —  Selected  parameters  of  the  Mendocino  County  deer 
population  as  affected  by  alternative  hunting  strategies. 


Strategy 

45%  bucks 

50%  bucks 

No 

25%  adult 

30%  does 

15%  does 

hunting 

bucks 

15%  fawns 

60%  fawns 

(1) 

(2) 

(3) 

(4) 

Total  deer 

June  1 

236,000 

251,000 

141.000 

168,000 

November  1 

191,000 

191,000 

117,000 

141,000 

May  30 

150,000 

148,000 

90,000 

93,000 

Annual  Losses 

Natural 

85,000 

95,000 

15,000 

22,000 

Hunting 

- 

7,900 

36,000 

53,000 

Natural  hunt-loss  ratio 

- 

12:1 

0.4:1 

0.4:1 

Kill  as  percent  of  June  1 

population 

- 

3 

26 

32 

Percent  composition  of  kill 

Bucks 

_ 

100 

42 

22 

Does 

- 

0 

41 

18 

Fawns 

- 

0 

17 

60 

Herd  composition  data 

Fawns/ 100  does 

Spring 

41 

41 

90 

50 

Fall 

64 

64 

83 

96 

Bucks/ 100  does 

Fall 

86 

43 

41 

22 

stability  had  been  attained.  Year-to-year  vari- 
ability was  suppressed  to  highlight  the  dif- 
ferences among  the  hunting  strategies.  The 
principal  features  of  each  strategy  are  sum- 
marized below: 

1.  No  Hunting:  This  strategy  is  presented 
mainly  for  comparison  with  the  other 
runs.  It  is  characterized  by  a  high  buck: doe 
ratio,  low  productivity,  and  high  natural 
mortality. 

2.  Twenty-Five  Percent  Adult  Bucks:  This 
is  an  estimate  of  the  hunting  effected  in 
Mendocino  County  during  the  past  10  + 
years.  Hunting  is  limited  to  males  with 
two  or  more  points  per  antler.  Natural 
mortality  is  higher  than  in  Strategy  1 
because  the  population  includes  relatively 
more  does,  as  indicated  by  the  buck: doe 


ratio,  and  the  number  of  fawns  born  is, 
therefore,  higher.  Fawns  are  most  sus- 
ceptible of  all  age  classes  to  natural  mor- 
tality. Overall  deer  numbers  do  not  differ 
markedly  between  Strategies  1  and  2. 
For  every  deer  taken  by  hunters,  about 
12  die  of  starvation  and  other  natural 
causes.  Although  the  management  goals 
are  not  explicitly  denned,  current  regu- 
lations result  in  the  maintenance  of  maxi- 
mum deer  numbers  and  maximum  natural 
losses.  This  strategy  provides  no  con- 
straint upon  overall  deer  numbers. 
Forty-Five  Percent  Bucks,  Thirty  Per- 
cent Does,  and  Fifteen  Percent  Fawns: 
Where  the  hunter  is  allowed  to  select 
either  bucks  or  does,  this  strategy  repre- 
sents the  results  of  the  heaviest  hunting 


128 


pressure  likely  of  achievement.  Although 
hunters  generally  avoid  killing  fawns  if 
possible,  data  from  other  areas  indicate 
that  fawns  comprise  15%  to  20%  of  the  kill 
in  antlerless  hunts.  The  annual  kill  of 
36,000  Would  probably  require  private 
lands  to  be  hunted  as  heavily  as  public 
lands.  Comparison  with  Strategy  2  indi- 
cates that  the  hunting  kill  would  increase 
about  45% ,  even  though  the  overall  popu- 
lation decreases  about  40%  .  Natural  losses 
are  also  much  reduced. 
4.  Fifty  Percent  Bucks,  Fifteen  Percent  Does, 
and  Sixty  Percent  Fawns:  While  the  previ- 
ous strategy  would  tend  to  maximize  the 
hunting  kill  if  hunters  were  allowed  their 
free  choice  of  animals,  the  kill  could  be  fur- 
ther increased  by  selectively  hunting  fawns. 
This  strategy  is  comparable  with  the  usual 
sheep  management  regime  in  Mendocino 
County,  where  a  high  proportion  of  lambs 
is  marketed  annually.  Although  the  kill 
would  be  considerably  higher  than  in  the 
previous  strategy,  the  total  biomass  yield 
would  be  slightly  lower  because  of  the 
relatively  small  size  of  fawns.  It  may  be 
unrealistic  to  propose  that  50%  of  the 
bucks  can  be  killed  annually.  However, 
if  the  goal  of  management  is  to  maximize 
the  number  of  animals  taken  by  hunting, 
it  is  necessary  to  maintain  the  highest 
possible  proportion  of  breeding  does  in  the 
herd,  and  this  can  be  achieved  only  by 
heavy  hunting  of  adult  males. 

A  convenient  way  of  showing  hunting  yield 
and  population  numbers  at  equilibrium  for 
different  strategies  is  by  plotting  the  results 
from  many  computer  runs  on  graphs  like  these 
shown  in  Figures  4  through  6.  These  graphs 
permit  a  comparison  of  the  relative  effects  of 
selective  hunting  pressure  directed  against 
does,  fawns,  and  buck,  respectively. 

Figure  W-  This  graph  depicts  population 
trends  and  yields  of  deer  when  various  per- 
centages of  does  are  taken  by  hunting  when 
(A)  no  bucks  or  fawns  are  taken,  and  (B)  50% 
of  all  bucks  and  15%  of  the  fawns  are  taken  an- 
nually. Several  pertinent  aspects  of  population 
performance  are  apparent  from  this  graph: 

(1)  With   no  hunting  of  bucks  and  fawns,  a 
slightly  higher  total  population  of  deer 


tends  to  be  maintained  when  any  given 
removal  of  does  is  carried  out. 

(2)  Maximum  productivity  or  yield  of  the 
population  is  achieved  when  approximate- 
ly 25%  of  the  does  are  removed  annually. 
However,  the  total  yield  is  approximately 
five  times  higher  if  bucks  and  fawns  are 
taken  as  specified  in  Strategy  (B). 

(3)  As  hunting  pressure  on  does  increases, 
overall  deer  numbers  decrease  at  an  in- 
creasing rate. 

Figure  5:  Figure  5  indicates  the  effect  of 
increasing  fawn  removals  accompanied  by  (C) 
no  buck  or  doe  hunting,  or  (D)  annual  hunting 
removals  of  50%  of  the  bucks  and  30%  of  the  does. 
It  shows  that: 

(1)  The  total  population  will  decline  only 
slightly  with  the  increasing  removal  of 
fawns  only,  as  depicted  by  (C). 

(2)  Under  the  buck-doe  strategy  in  (D), 
maximum  yield  and  population  size  will 
diminish  rapidly  if  annual  fawn  removal 
exceeds  approximately  30% . 

Figure  6:  The  hunting  conditions  set  forth 
on  this  graph  are,  (E)  no  does  or  fawns  are 
taken  as  related  to  the  increasing  take  of  bucks, 
and  (F)  a  removal  of  30%  of  the  does  and  15%  of 
the  fawns  in  relation  to  an  increasing  take  of 
bucks.  The  graphs  show  that: 

(1)  Buck  removal  alone  has  only  a  slight 
effect  on  yield,  and  even  less  effect  on 
total  population. 

(2)  When  does  and  fawns  are  taken  as  speci- 
fied in  Strategy  (F),  the  total  yield  of 
the  population  is  roughly  doubled,  as  com- 
pared to  taking  bucks  only. 


GENERAL  RELATIONSHIPS 

Consideration  of  the  three  graphs  shows  that: 

(1)  Maximum  yield  of  the  Mendocino  County 
deer  population  is  only  achieved  through  ex- 
ploitation. 

(2)  Reduction  of  the  large,  unexploited  popu- 
lation through  hunting  produces  a  more  dy- 
namic population,  with  greater  turnover.  The 
basic  relationship  is  to  lower  stocking  rate 
on  the  range,  which  reduces  competition  for 
available  feed,  and  thereby  raises  the  plane 
of  nutrition.  This,  in  turn,  improves  fecundity 
and  survival. 


129 


180   _ 


140 


120 


100 


(A)  Stable  November  1  Population 
Buck  Hunting  07. 
Fawn  Hunting  07. 


20        30       40        50 
Percent  Does  Taken  Annually 


Figure  4.  —  Yield  and  population  numbers  at  equilibrium  for  two  buck-fawn 
hunting  strategies  and  variable  doe  hunting  percentages. 


(3)  It  appears  that  maximum  population 
and  yields  will  probably  be  achieved  with  a 
hunting  removal  of  about  20-25%  of  the  does, 
15-30%  of  the  fawns,  and  over  50%  of  the  bucks 
annually.  At  this  rate  of  buck  removal,  there 
is  no  possibility  of  reducing  the  breeding  suc- 
cess of  the  population,  but  it  is  highly  unlikely 
that  such  a  high  rate  of  buck  take  can  ever  be 
achieved  over  the  county  as  a  whole.  The  density 
of  cover  on  much  of  the  deer  range  precludes 
it.  Under  present  hunting  practices,  a  buck 
removal  of  possibly  20-25%  is  being  achieved. 
At  best,  this  might  possibly  be  doubled. 


Likewise,  it  is  highly  unlikely  that  hunters 
can  be  forced  to  take  large  numbers  of  fawns 
selectively.  Most  either-sex  hunting  efforts 
can  be  expected  to  produce  a  take  of  fawns 
of  about  10-20% ,  and  it  is  difficult  to  increase 
this,  as  hunters  try  to  avoid  taking  fawns 
because  of  their  small  size. 

(4)  Removal  of  does  above  the  25%  level  is 
the  most  powerful  means  available  for  total 
population  control,  since  it  reduces  total  re- 
productive potential.  This  finding  is  readily 
applicable  to  the  special  management  problems 
in  National   Parks,  where  big  game  numbers 


130 


160 


140 


Percent  Fawns  Taken  Annually 


Figure    5.   —   Yield    and   population    numbers    at   equilibrium   for   two    buck-doe 
hunting  strategies  and  variable  fawn  hunting  percentages. 


must  be  controlled,  but  public  hunting  is  con- 
sidered incompatible  with  other  management 
goals.  In  such  situations,  it  is  customary  for 
surplus  animals  to  be  shot  by  park  officials. 
Our  calculations  indicate  that  these  removal 
programs  should  be  directed  solely  against 
adult  females  to  provide  the  most  effective 
population  control.  This  would  minimize  the 
number  of  animals  to  be  killed,  as  well  as  the 
manpower  requirements,  and  would  additional- 
ly maintain  a  high  proportion  of  the  aesthetical- 
ly desirable  adult  males  in  the  population. 


CONCLUSIONS 


Computer  simulation  of  dynamic  biomanage- 
ment  systems  appears  to  provide  a  means  of 
generating  information  useful  to  resource 
managers  and  to  research  administrators.  In 
building  computer  simulation  models,  research- 
ers and  managers  put  together  their  theoretical 
and  practical  knowledge  of  a  system.  This 
process  frequently  results  in  finding  existing 
gaps   in   empirical   data,   and   helps   to   revise 


131 


160 


140 


120 


100 


60 


'! 


(E)   Stable  November  1  Population 
Doe  Hunting  0% 
Fawn  Hunting  07. 


Stable  November  1  Population 
Doe  Hunting  307. 
Fawn  Hunting  157. 


|(F)  Yield  at  Stable  Population 
)     Doe  Hunting  30% 


30       40        50       60 
Percent  Bucks  Taken  Annually 


Figure  6.  —  Yield  and  population  numbers  at  equilibrium  for  two  doe- 
fawn  hunting  strategies  and  variable  buck  hunting  percentages. 


research  plans  and  data  collection  procedures 
for  monitoring  the  real  system.  Outside  of  this 
important  research  administration  outcome, 
information  about  consequences  of  manage- 
ment policies  which  might  otherwise  not  be 
obvious  can  be  provided.  For  example,  our 
results  to  date  indicate  that  annual  revisions 
of  the  hunting  regulations  will  not,  in  general, 
cause  management  objectives  to  be  attained 
more  rapidly  than  following  a  fixed  hunting 
strategy.  This  is  due  to  the  compounding  ef- 
fects of  random  variability  and  the  difficulties 
in  monitoring  the  system. 


The  systems  analysis  approach,  and  its  con- 
comitant technique  of  computer  simulation, 
can  and  has  been  used  to  study  other  wildlife 
resources  such  as  fish  populations.  Models  de- 
veloped for  fish  populations  would  necessarily 
incorporate  the  unique  features  of  each  system, 
and  the  output  would  be  designed  according 
to  the  special  needs  of  the  resource  manager. 
However,  further  exploration  of  the  usefulness 
of  computer  simulation  in  studying  fish  popu- 
lations is  needed  before  the  optimism  shown 
for  big  game  management  can  be  expressed 
for  management  of  fisheries. 


132 


Augmentation  of  Salmon  Stocks  through  Artificial 
Propagation:  Methods  and  Implications1 

Joe  B.  Stevens  and  Bruce  W.  Mattox2 
ABSTRACT 

Eighty-one  hatcheries  on  the  Pacific  Coast  now  rear  significant  numbers  of  salmon 
and  steelhead  for  sport  and  commercial  fisheries.  Annual  operation  and  maintenance 
costs  amount  to  $6.6  million.  A  production  function  analysis  of  15  Oregon  Fish  Com- 
mission hatcheries  produced  tentative  conclusions  that  (a)  controlled  inputs  were  com- 
bined in  fixed  proportions,  (b)  constant  returns  to  size  were  realized,  and  (c)  some 
degree  of  factor  substitution  existed  between  the  controlled  "fixed  proportion  input" 
and  water  temperature.  The  latter  relationship  may  allow  hatchery  managers  to  im- 
prove efficiency  at  the  hatchery  level.  Uncertainty  with  respect  to  downstream  en- 
vironmental conditions,  however,  must  be  considered  along  with  returns  to  size  for 
the  hatchery  production  function  when  new  investments  are  undertaken. 

Fixed  asset  theory  was  used  to  conceptualize  exit  and  entry  of  salmon  harvesting 
resources  between  1947  and  1966.  Net  entry  followed  years  of  good  catches,  but  net 
exit  did  not  occur  following  the  bad  years.  If  a  major  objective  of  hatchery  programs 
is  to  augment  fishermen's  incomes,  consideration  must  be  given  to  increasing  the 
opportunity  costs  of  extant  resources  as  well  as  to  limiting  entry  of  new  resources. 


INTRODUCTION 

It  is  a  moot  question  to  ask  whether  or  not 
the  public  sector  should  involve  itself  exten- 
sively in  hatchery  rearing  of  salmon  and  steel- 
head  on  the  Pacific  Coast.  Eighty-one  hatcher- 
ies, valued  at  over  $56  million  with  annual  oper- 
ation and  maintenance  costs  of  $6.6  million, 
now  rear  significant  numbers  of  chinook  and 
coho  salmon  and  steelhead  trout  for  sport  and 
commercial  fisheries.  It  is  a  relevant  question, 
however,  to  ask  under  what  conditions  con- 
tinuing investment  of  this  type  should  be  under- 
taken. Although  this  is  a  question  which  can 
and  should  be  posed,  it  is  not  easily  answered; 
thus  we  do  not  attempt  to  do  so,  aside  from 
exploring    some    obvious    and    not-so-obvious 


implications.  Our  major  attention  herein  is 
devoted  to  asking  and  partially  answering 
the  question:  "Given  the  decision  to  augment 
resource  flows  by  artificial  propagation,  what 
can  be  gleaned  from  existing  data  which  will 
allow  the  public  sector  to  increase  efficiency 
at  the  hatchery  level?"  In  exploring  this  ques- 
tion, we  recognize  the  dangers  of  a  partial 
analysis,  i.e.,  divorcing  hatchery  objectives 
from  higher  order  objectives.  Our  defense  is 
pragmatic,  i.e.,  that  it  is  better  to  start  fitting 
the  pieces  of  the  puzzle  together,  one  by  one, 
than  to  not  start  at  all  or  to  theorize  how  they 
might  all  be  fitted  simultaneously. 

THE  CURRENT  SIGNIFICANCE  OF 
SALMON  AND  STEELHEAD  HATCHERIES5 


1  Technical  Paper  No.  3010,  Oregon  Agricultural 
Experiment  Station. 

2  Associate  Professor  of  Agricultural  Economics, 
Oregon  State  University,  and  Assistant  Professor  of 
Resource  Economics,  University  of  Rhode  Island,  res- 
pectively. This  publication  is  supported  in  part  by  the 
National  Oceanic  and  Atmospheric  Administration 
(maintained  by  the  U.S.  Department  of  Commerce)  In- 
stitutional Sea  Grant  2-35187.  Nothing  stated  herein  is 
to  be  taken  as  representing  the  views  or  policies  of  the 
Oregon  Fish  Commission. 


The  first  Pacific  Coast  salmon  hatchery  was 
constructed  in  Northern  California  by  the 
U.S.  Fish  Commission  almost  a  century  ago. 
Since  that  time,  artificial  propagation  of 
salmon  has  alternately  been  viewed  as  a 
panacea  and  as  no  solution  at  all.  Improve- 
ments in   propagation  methods  have  allowed, 


3  Data  on  the  nature  and  contributions  of  hatchery 
programs  were  taken  freely  and  gratefully  from  Wahle, 
(1970). 


133 


and  environmental  deterioration  has  forced, 
increased  reliance  on  hatchery  operations, 
especially  in  the  past  decade.  Eighty-one 
hatcheries  are  now  operated  by  fishery  agencies 
of  Alaska,  Canada,  California,  Oregon,  and 
Washington,  and  by  the  Bureau  of  Sport  Fish- 
eries and  Wildlife.  Extensive  evaluation  pro- 
grams are  carried  on  by  the  Columbia  Fisheries 
Program  Office  of  the  National  Marine  Fish- 
eries Service  and  by  some  of  the  other  agencies. 
The  evaluative  work  of  the  NMFS  program 
has  included  extensive  fin-clipping,  sampling 
for  marked  salmon,  and  benefit-cost  analyses 
for  brood  years  by  species. 

The  current  status  of  these  resource  augmen- 
tation programs  has  recently  been  summar- 
ized by  Wahle  (1970)  of  the  NMFS,  and  is 
portrayed  in  Table  1.  Survival  rates  of  4  to  5% 
indicate  that  a  multitude  of  fingerlings  must 
be  released  in  order  to  affect  resource  stocks. 
The  cost  of  production  for  one  fingerling,  on 
the  other  hand,  is  relatively  low.  Our  study 
revealed  that  the  15  hatcheries  of  the  Oregon 
Fish  Commission  produced  the  equivalent  of 
about  70  million  salmon  and  steelhead  finger- 
lings  between  October  1,  1968  and  April  30, 
1970,  at  a  cost  of  slightly  over  two  cents  per 
fingerling.4  Assuming  that  the  survival  rates 
in  Table  1  are  appropriate,  the  cost  per  fish 
caught  at  some  time  in  the  future  rises  to  about 
$1.35,  disregarding  any  discounting  for  time. 

The  contributions  of  hatchery-reared  fish  to 
the  ocean  troll  fishery  is  impressive,  ranging 
from  30  to  80%  of  total  catch  in  1968.  Wahle 
points  out,  however,  that  the  proportion  of 
hatchery  fish  to  wild  fish  was  higher  than 
usual  in  that  year.  The  true  contribution  to 
the  sport  catch  of  coho,  for  example,  may  be 
closer  to  50% . 

It  may  be  useful  to  this  group  to  have  the 
hatchery  programs  put  into  perspective  with 
the  total  salmon  catch  for  the  West  Coast 
States  of  Washington,  Oregon,  and  California. 


Table  1.  —  Survival  rates  and  contributions  to  ocean 
troll  fisheries  of  hatchery-reared  salmon  and  steelhead 
in  1968. 


Hatchery-reared  fish  as 

a  percentage  of  total 

Species 

Survival  rate1 

ocean  troll  catch  (1968)2 

Coho 

0.04     (.037) 

Commercial:  30% 
Sport:  80% 

Fall  Chinook 

.004  (.003) 

Commercial:  70% 

Spring  Chinook 

.05 

Sport:  65% 

Steelhead 

.04 

— 

'Survival  rates  for  coastal  streams  are  shown  in  parentheses. 
2The  commercial  fishery  data  for  chinook  salmon  include 
landings  from  the  west  coast  of  Vancouver  Island,  in  addition 
to  landings  in  Oregon,  Washington,  and  California.  The  sport 
landings  include  only  the  latter  three  States. 
SOURCE:   Wahle,  1970. 


To  do  so,  we  have  done  some  quick  (and  dirty) 
calculations  for  which  we  assume  sole  responsi- 
bility. The  total  yearly  landings  of  all  salmon 
in  this  region  fiucuate  widely  because  of  the 
odd-year  cyclical  nature  of  pink  salmon,  an 
important  species  for  which  hatchery  propa- 
gation work  is  now  in  advanced  experimental 
stages  (McNeil,  1969).  Averaging  one  recent 
cycle  year  for  pink  salmon  (1967)  with  one 
non-cycle  year  (1964),  about  two-thirds  of  the 
total  salmon  catch  is  comprised  of  coho  and 
chinook  (U.S.  Department  of  the  Interior, 
1947-1967).  Assuming  that  Wahle's  data  from 
Table  1  are  appropriate  for  coho  and  chinook, 
regardless  of  method  of  capture5  (troll,  gill  net, 
purse  seine),  and  using  a  conservative  hatchery- 
contribution  share  of  30% ,  it  would  appear 
that  perhaps  20%  of  the  total  West  Coast  (U.S.) 
salmon  fishery  is  supported  by  hatchery  pro- 
grams. This  share  is  increasing  over  time, 
and  success  in  rearing  pink  salmon  will  pro- 
vide further  augmentation. 


4  This  assumes  that  coho  and  spring  chinook  were 
released  at  15  fingerlings  per  pound  of  fish,  fall  chinook 
at  100  per  pound,  and  steelhead  at  10  per  pound.  Costs 
include  variable  operating  expenditures  plus  and  im- 
puted 5%  charge  on  the  $7.5  million  replacement  value 
of  fixed  facilities  (Mattox,  1970  and  Wahle,  1970).  The 
latter  sum  is  no  doubt  an  overestimate  of  real  capital 
values. 


5  The  troll  fishery  accounted  for  about  63%  of  total 
coho  and  chinook  capture,  averaging  1964  and  1967 
data.  Ocean  troll  alone  would  constitute  at  least  50% 
of  total  catch. 


134 


PRODUCTION  FUNCTION  ANALYSIS  OF 

HATCHERY  PROPAGATION  OF 

SALMON  AND  STEELHEAD 

The  Incentive  Framework  of 
Hatchery  "Firms" 

As  is  usually  the  case,  our  initial  research 
objectives  were  more  elegant  than  could  be 
accomplished  with  existing  time  and  data. 
Initial  plans  were  to  estimate  marginal  pro- 
ductivities for  each  of  several  factors  of  pro- 
duction relevant  to  the  15  major  hatcheries  of 
the  Oregon  Fish  Commission,  and  to  estimate 
the  total  elasticity  of  production  or  returns 
to  size  for  these  hatcheries.  If  possible,  we 
wanted  to  incorporate  into  the  function  post- 
hatchery  phenomena,  especially  the  physical 
returns  to  the  fishery  of  hatchery-reared  fish. 
In  that  the  NMFS  data  on  the  latter  were 
not  yet  precise  enough  to  identify  differential 
returns  by  hatchery,  it  was  necessary  to 
restrict  the  analysis  to  the  hatchery  produc- 
tion function. 

One  of  the  most  interesting  aspects  of  the 
analysis  was  the  influence  on  model  specifica- 
tion of  the  incentive  framework  of  the  hatch- 
eries. Federal  and  State  hatcheries  receive  no 
price  for  their  product,  have  no  responsibility 
for  realizing  profit,  and  are  managed  by  pro- 
fessionals trained  primarily  in  terms  of  bio- 
logical relationships.  Budget  constraints  are 
imposed  by  the  political  rationing  process. 
Furthermore,  the  nature  of  the  incentive  frame- 
work is  such  that  it  is  only  partially  con- 
ducive to  providing  data  in  a  form  which  is 
useful  for  economic  analysis. 

On  the  other  hand,  hatchery  managers  are 
not  unaffected  by  economic  forces,  since  they 
face  constraints  on  operating  capital  and  tech- 
nology as  well  as  constraints  with  respect  to 
factor  prices,  fixed  facilities,  and  natural  phe- 
nomena. Among  the  latter  are  yearly  and 
seasonal  variations  in  water  quantity,  which 
often  result  in  the  non-use  of  rearing  ponds, 
and  seasonal  variations  in  water  temperatures 
which  affect  metabolic  processes  of  fry  and 
fingerlings. 

The  absence  of  a  product  price,  of  course, 
does  not  mean  that  the  conventional  econo- 
mizing model  is  not  relevant.  The  influence  of 
technological,  budgetary,  and  factor  price  con- 


straints seemed  sufficiently  strong  to  postulate 
that  hatcheries  attempt  to  maximize  output 
subject  to  these  constraints.  In  one  major 
respect,  however,  it  was  anticipated  that  the 
decision  framework  of  the  hatchery  managers 
would  give  rise  to  a  type  of  empirical  result 
not  usually  obtained  in  analyses  of  private 
firms.  That  is,  it  was  hypothesized  that  the 
particular  set  of  hatcheries  we  observed  were 
(a)  combining  controlled  inputs  in  fixed  pro- 
portions, and  (b)  realizing  constant  returns 
to  size. 

The  reasoning  behind  this  hypothesis  largely 
reflects  the  institutional  nature  of  the  hatch- 
eries, although  physical  attributes  of  the  pro- 
ductive factors  serve  as  necessary  conditions. 
The  primary  institutional  factor  is  the  influ- 
ence of  centralized  supervision  on  the  Fish 
Commission  hatcheries.  Resident  managers 
appear  to  operate  within  guidelines  set  by  the 
central  office  with  respect  to  input  combin- 
ations, a  system  which  is  reinforced  by  dis- 
ciplinary training  of  both  groups.  The  physical 
attributes  of  factors  which  would  allow  them 
to  be  combined  in  fixed  proportion  is  a  rela- 
tively high  degree  of  divisibility.  The  latter 
is  elaborated  below. 


The  Biological  Production  Function 

The  underlying  production  function  for 
fingerlings  can  be  viewed  as  consisting  of  three 
controlled  factors  —  food,  labor,  and  rearing 
space  —  and  one  non-controlled  factor  —  water 
temperature.  The  food  variable  is  nutritionally 
complex,  but  a  convenient  one  for  analytical 
purposes  since  the  Oregon  Moist  Pellet  is  a 
"complete"  ration.  This  food,  fed  in  a  variety 
of  pelletized  and  mash  forms,  was  specially 
formulated  to  satisfy  the  nutritional  demands 
of  fingerlings  at  different  ages  as  well  as  for 
prevention  and  treatment  of  disease.  Further, 
the  food  is  centrally  purchased,  thus  elimina- 
ting any  price  differentials  between  hatcheries. 

Although  mechanical  feeders  have  been  tried 
in  some  areas,  the  Fish  Commission  feeds 
entirely  by  hand  application  of  the  pellets. 
In  that  a  pool  of  temporary  labor  is  usually 
available  to  resident  managers,  both  labor 
and  food  variables  are  quite  divisible. 

The  third  major  controlled  variable,  rearing 
space,   might   be   described,    tongue    in   cheek, 


135 


as  water  surrounded  by  concrete.  Water  flows, 
as  noted  earlier,  vary  in  quantity  and  tem- 
perature. Both  of  these  physical  dimensions 
are  largely  outside  the  control  of  management. 
Although  some  low  flow  augmentation  is  ac- 
complished, the  usual  result  of  low  flows  during 
summer  months  has  been  an  inability  to  fully 
utilize  rearing  space.  Since  the  rearing  ponds 
are  fairly  small  and  numerous,  low  flows  are 
adjusted  for  by  maintaining  water  volume 
in  some  ponds  and  temporarily  retiring  others. 
Thus,  the  rearing  space  actually  used  is  also 
fairly  divisible,  although  some  seasonal  excess 
capacity  may  exist. 

Although  our  initial  inclination  was  that 
separate  marginal  factor  productivities  might 
be  estimated,  discussions  with  hatchery  man- 
agers soon  revealed  the  similarity  of  practices 
in  combining  controlled  inputs.  Levels  of 
inputs  and  outputs  at  larger  hatcheries  seem- 
ed to  be  constant  multiples  of  those  found  at 
smaller  hatcheries,  although  opportunities  for 
variable  input  proportions  seemed  to  be  present 
in  a  physical  sense.  One  could,  for  example, 
stock  rearing  ponds  with  fingerlings  at  dif- 
ferent rates,  or  spread  existing  water  flows 
over  all  rearing  ponds.  Centralized  manage- 
ment, of  course,  may  not  be  conducive  to  such 
experiments.  On  the  other  hand,  it  may  well 
be  that  past  "experiments",  intended  or  other- 
wise, have  revealed  that  other  factor  combin- 
ations involve  a  greater  degree  of  risk.  For 
example,  disease  spreads  rapidly  in  rearing 
ponds;  overcrowding  of  fingerlings  might  be 
disastrous.  Similarly,  lower  water  levels  in 
all  ponds  would  increase  water  temperature 
and  accelerate  the  spread  of  disease. 

Our  hypotheses  of  fixed  factor  proportions 
and  constant  returns  to  size  were  equivalent 
to  expecting  that  the  Fish  Commission  acts 
as  if  the  isoquants  for  hatchery  production  of 
fingerlings  are  right-angled,  whether  they 
actually  are  or  not.  The  hypothesis  was  strong- 
ly dependent,  of  course,  on  our  prior  decision 
to  analyze  Fish  Commission  hatcheries.  A 
cross-section  analysis  over  various  agencies, 
in  retrospect,  would  possibly  have  yielded  more 
empirical  information. 

The  non-controlled  variable,  water  tempera- 
ture, can  be  quite  important  during  periods 
of  either  cold  or  warm  weather.  Extremes  of 
either  type  seem  to  effect  primarily  the  volun- 


tary rate  of  metabolic  activity,  rather  than 
the  efficiency  of  food  conversion  (Paloheimo 
and  Dickie,  1966).  It  was  expected,  then,  that 
growth  would  be  retarded  in  the  upper  and 
lower  limits  of  observed  water  temperature. 
This  noncontrolled  variable,  then,  was  viewed 
as  the  principal  shifter  of  a  constant  returns 
production  function. 


Exploratory  Estimation 

The  time  period  selected  for  analysis  was 
October  1,  1968  through  April  30,  1970.  This 
19-month  period  allowed  the  propagation  pro- 
cess to  be  observed  for  at  least  one  brood 
year  for  each  species  of  interest  (Figure  1). 
These  included  coho,  spring  chinook,  fall 
chinook,  chum,  and  steelhead.  In  the  absence 
of  cost  data  which  were  separable  by  species, 
it  was  necessary  to  estimate  an  aggregate 
function  over  all  species.6 

In  view  of  the  fixed  proportions  hypothesis, 
the  initial  attempt  at  estimation  involved 
several  of  the  factors  which  were  thought  to 
be  jointly  combined.  We  were  limited  in  this 
analysis  by  the  absence  of  data  on  either  actual 
water  flows  or  rearing  space  used.  As  a  fairly 
unsatisfactory  proxy,  these  variables  were  re- 
placed by  a  measure  of  the  replacement  value 
of  all  fixed  facilities.  This  variable,  along  with 
food,  operating  expenses  (largely  labor),  and 
cumulative  water  temperature  units7  for  the 
warm  weather  period  and  the  cold  weather 
period,  constituted  the  five  independent  vari- 
ables in  the  initial  run. 

As  anticipated,  a  high  degree  of  intercorre- 
lation  resulted  between  food,  operating  ex- 
penses, and  the  value  of  fixed  facilities  in  both 
Cobb-Douglas  and  linear  estimations.  Correla- 
tion coefficients  between  these  three  variables 
approached  or  exceeded  0.80,  and  resulted  in 
a  considerable  inflation  of  standard  errors. 
Since  it  appeared  that  some  degree  of  factor 
substitution   could   be   estimated  between   any 


6  An  interagency  effort  is  now  underway  to  explore 
cost  accounting  systems  by  species. 

7  A  cumulative  temperature  unit  (CTU)  is  defined 
for  each  day  in  which  the  average  water  temperature 
exceeds  32°F  by  one  degree.  One  month  of  40°  water 
temperature,  for  example,  would  constitute  240  CTU's. 


136 


1968 


1969 


SPECIES   AND   BROOD        o|n|d|j|f|m|      A    |    M     |     J  J        A  S  ON  D        J  F     |    M     |     A 


COHO 


FALL  CHINOOK 


^ 


1968 
1969 


SPRING  CHINOOK 


1967 
1968 
1969 


CHUM 


^^^^^ 


1968 
1969 


^^ 


Figure  1.  -  Brood  year  classification  of  species  propagated  from  the  beginning'  of  October 

1968  through  April  1970. 


one  of  these  variables  and  water  temperature, 
the  food  variable  was  retained  in  further 
analyses. 

Marginal  Factor  Productivities 
and  Returns  to  Size 

Since  the  underlying  functional  relation- 
ships were  unknown,  output  response  func- 
tions were  estimated  in  both  linear  and  log- 
linear  (Cobb-Douglas)  forms.  Within  each  func- 
tional form,  estimates  were  obtained  relating 
to  two  different  assumptions  about  the  inter- 
cept term.8  The  output  response  functions  and 
marginal  physical  productivities  are  shown 
in  Table  2. 

Several  items  are  worthy  of  note.  First,  the 
R2  values  were  uniformly  high,  regardless  of 
functional  form.  Second,  the  marginal  pro- 
ductivity estimates  appeared  reasonable  and 
were  fairly  constant  over  the  various  functional 
forms.  The  marginal  productivity  of  one  pound 
of  food  was  about  0.58  pounds  of  salmon,  a 


Table  2.  —  Output  response  functions  and  marginal 
physical  productivities  for  the  15  Oregon  Fish  Com- 
mission salmon  hatcheries. 


8  While  output  would  logically  be  zero  if  all  input 
levels  were  zero,  an  estimate  of  the  intercept  may  be 
helpful  in  assessing  the  "constant  returns"  argument 
for  the  linear  function. 


Functional  form 

Intercept 

b 

c 

b 

l 

Variables1 
b                b 

2                      3 

ft- 

I.    (a)  Linear 

2-13,998 

.572 

-15.694 

33.324 

0.959 

3     (.10) 

(.01) 

(.10) 

(.05) 

4    .572; 

-15.694 

33.324 

I.    (b)  Linear 

0 

.563 

-16.735 

29.715 

.991 

- 

(.01) 

(.05) 

(.05) 

.563 

-16.735 

29.715 

II.  (a)  Log-Linear 

-  23.74 

1.106 

-     .334 

.526 

.960 

(.01) 

(.01) 

(20) 

(.05) 

.620 

-11.281 

31.618 

II.  (b)  Log-linear 

(t 

1.047 

-     .450 

.332 

.999 

- 

(.01) 

(.10) 

(.20) 

.588 

-15.217 

19.958 

1  Variables: 

y    =    pounds   of   output  of  salmon   (released  e 

fingerlings  or  swim-up  fry) 
X   =    pounds  of  food  fed  (Oregon  Moist  Pellet). 

ther  as 

X^  =    aver 

ige   cumulat 

ve   tem| 

jerature 

units   (CTU's)  of 

water  from  May  through  October  (warm  season). 
X,  =    average  CTU's  of  water  from   November  through 
April  (cold  season). 

2  Regression  coefficients. 

3  Significance  level. 

4  Marginal  physical  productivities. 


137 


figure  that  seems  consistent  with  the  literature 
in  fisheries  biology  (Paloheimo  and  Dickie, 
1966).  Adding  one  day  with  water  temperature 
one  degree  in  excess  of  32 °F  (i.e.,  one  CTU) 
during  the  cold  season  would  add  about  30 
pounds  to  total  output;  one  additional  CTU 
during  the  warm  season  would  reduce  output 
by  about  15  pounds.  Third,  the  high  R2  values 
support  the  hypothesis  of  fixed  factor  propor- 
tions, although  we  recognize  that  another 
analysis,  covering  several  agencies  and  systems 
of  management,  might  well  yield  different 
results.  Fourth,  the  evidence  appears  to  sup- 
port the  "constant  returns"  hypothesis,  al- 
though this  is  somewhat  conjectural.  Summing 
the  coefficients  for  Cobb-Douglas  forms  is 
hindered  by  the  negative  coefficient  on  warm 
season  water  temperatures.  One  might,  as  we 
did,  view  the  water  temperature  variables  as 
"shifters"  of  food-input  relationship.  If  so, 
the  coefficients  on  the  food  variable  do  not 
differ  significantly  from  unity.9 

Our  estimates  of  marginal  productivities 
thus  enabled  us  to  ask,  "What  would  be  the 
change  in  hatchery  output  if  one  were  to  in- 
crease (or  decrease)  water  temperatures  by  a 
given  amount?"  A  10%  reduction  in  CTU's 
during  the  warm  season  would  reduce  average 
water  temperature  from  52.97°F  to  50.87°F 
and  cause  output  to  increase  by  5,684  pounds, 
or  about  4.36%  of  the  mean  hatchery  output. 
Raising  cold  season  water  temperatures  from 
43.99°F  to  45.19°F  would  add  6,218  pounds 
of  output,  or  about  4.77%  of  mean  hatchery 
output. 

Factor  Substitution 

If  controlled  inputs  are  combined  in  fixed 
proportions,  as  evidenced  above,  the  data  ob- 
viously do  not  allow  estimation  of  substitution 
possibilities.  On  the  other  hand,  our  analysis 
does  permit  us  to  identify  degrees  of  sub- 
stitution between  the  fixed  proportion  input, 
using  food  as  a  proxy  variable,  and  changes 


9  The  negative  intercept  on  the  linear  model  was 
significantly  different  from  zero  at  the  0.10  level.  This 
gives  some  evidence  of  increasing  returns,  and  is 
consistent  with  the  bi  estimates  of  1.106  and  1.047 
for  the  log-linear  models.  Acceptance  or  rejection  of 
"constant  returns"  thus,  depends  partly  on  one's  pref- 
erence for  significance  levels. 


in  the  noncontrolled  water  temperature  vari- 
ables. The  marginal  rates  of  factor  substitu- 
tion, as  estimated  from  both  linear  and  log- 
linear  functions,  are  shown  in  Table  3.  Al- 
though log-linear  models  no  doubt  conform 
more  closely  to  biological  reality,  linear  rates 
of  substitution  may  be  appropriate  for  some 
decisions.  The  degree  of  isoquant  curvature  is 
largely  a  matter  for  the  judgment  of  fisheries 
biologists;  experimental  work  in  this  area 
should  be  useful  in  checking  and  refining  our 
estimates.  Our  confidence  in  the  linear  rates 
would  be  greatest  in  the  neighborhood  of  mean 
CTU  values  (e.g.,  Figure  2). 

Table  3.  —  Linear  rates  of  factor  substitution  between 
inputs.1 


3  [Food  (Xt)] 

3  [Food(X,)l 

Functional  form  3  [Summer  CTU's  (-X2) 

3  [Winter  CTU's  (X3)\ 

I.    (a)  Linear                          -27.462 

I.  (b)  Linear                          -29.714 

II.  (a)  Log-linear                  -18.186 
II.  (b)  Log-linear                   -25.867 

-58.309 
-52.762 
-50.972 
-33.935 

1  Estimates  are  based  on  mean  values.  The  sign  on  the  X   vari- 

2 

able  (warm  season  water  temperatures)  is  reversed  here  for 
convenience  since  decision  makers  would  attempt  to  reduce 
summer  temperatures  and  increase  winter  temperatures. 


Increased  environmental  control,  as  through 
controlling  water  temperature,  is  in  fact  one 
means  that  Pacific  Coast  fishery  agencies  are 
now  considering  for  output  augmentation.  Thus 
far,  the  agencies  have  primarily  adapted  to, 
rather  than  controlled,  this  aspect  of  the  en- 
vironment. The  hatching  of  fry  is  concentrated 
to  some  degree  in  those  hatcheries  which  have 
water  temperatures  most  conducive  to  this 
operation;  other  hatcheries  tend  to  specialize 
in  the  rearing  of  fingerlings.  Control  of  tem- 
peratures would  allow  both  food  and  transport 
costs  to  be  lowered,  although  empirical  data 
on  factor  price  ratios  were  not  available.  It 
was  our  thinking  that  the  estimates  of  factor 
substitution  in  Table  3,  together  with  a  step- 
by-step  presentation  of  "output  maximization, 
given  budget  constraints"  would  aid  agencies 
in  increasing  efficiency  at  the  hatchery  level.10 


10  Specific  attention  was  directed  to  the  problem  of 
determining  factor  prices  when  there  is  significant 
unused  capacity  at  existing  hatcheries.  As  mentioned 
earlier,  seasonal  low  water  flows  often  force  non-use 
of  some  rearing  ponds. 


138 


2  - 


.  Actual  observa 

Q  ~  70,500 

Q  ~ 130,481 

m  — 

-  \^ 

Q  .  2,210, 000 

^\Qi 

\  ^v 

'•' 

1    ' 

1      i  Ql 

Cumulative 
Temperature 


(Thousands) 

Figure  2.  —  Observed  relationships  between  food  and  cumulative 
temperature  units  (November  through  April). 


This  information  will  be  made  available  to 
hatchery  management  through  an  Oregon 
State  University  Marine  Economics  publication. 

Concluding  Comments  on  the 
Hatchery  Production  Function 

Several  strengths  and  qualifications  of  our 
research  became  clearer  as  the  work  progress- 
ed. The  principal  strength  is  that  our  conven- 
tional cross-sectional  analysis  of  "firms"  can 
be  useful  to  public  decisionmakers  in  spite 
of  their  "unconventional"  incentive  frame- 
works. Our  principal  lesson  in  methodology 
has  been  that  differences  within  frameworks 
of  the  various  agencies  may  be  more  crucial 
than  differences  between  those  of  private  and 
public  firms  if  the  researcher's  objective  is  to 
provide  a  substantial  empirical  input.  In  retro- 
spect, had  we  included  a  number  of  agencies 
in  our  study,  it  may  have  been  possible  to  esti- 
mate additional  substitution  relationships.  If 
our  limited  empirical  results  are  useful  to 
management  agencies,  however,  we  may  have 
opened  the  door  for  a  data  system  reorganiz- 
ation which  will  both  allow  for  improved  eco- 
nomic analysis  and  facilitate  consideration  of 
a  broader  range  of  production  alternatives. 


Our  policy  advice  is  accordingly  limited  by 
the  methodological  constraints  of  this  study. 
Constant  returns  from  hatchery  operations 
may  exist,  ceteris  paribus,  but  the  latter  may 
not  be  a  very  legitimate  assumption  when  un- 
certainty exists  as  to  downstream  environ- 
mental conditions.  Agencies  could,  for  example, 
spread  production  over  many  small  hatcheries 
located  on  different  streams,  but  it  may  be 
more  desirable  to  construct  fewer  and  larger 
hatcheries  if  environmental  protection  can  be 
assured  on  specific  streams. 

SOME  IMPLICATIONS  OF  INCREASED 

HATCHERY  PROPAGATION  FOR 

COMMERCIAL  FISHERIES  MANAGEMENT 


Associated  Harvesting  Costs 

The  principal  limitation  on  policy  advice 
stemming  from  our  research  is,  of  course, 
whether  or  not  increased  efficiency  at  the  hatch- 
ery level  necessarily  leads  to  increased  effic- 
iency at  the  fishery  level.  The  problems  of 
open-access  in  U.S.  commercial  fisheries  are 
well  known  to  this  group  and  will  not  be  re- 
peated   here    (Christy    and    Scott,    1966    and 


139 


Crutchfield  and  Pontecorvo,  1969).  Let  it  be 
sufficient  to  say  that  there  is  both  theoretical 
ambiguity  and  a  lack  of  empirical  information 
on  the  private  and  public  costs  associated  with 
harvesting  open-access  resources  (Bromley, 
1969). 

Two  lines  of  thought,  however,  would  prob- 
ably receive  acceptance  by  this  group.  The 
first  is  that  in  the  short  run,  hatchery  produc- 
tion could  increase  output  in  most  salmon 
fisheries  with  only  minor  increases  in  associ- 
ated harvesting  costs,  since  excess  capacity 
does  exist.  The  Crutchfield-Pontecorvo  re- 
search supports  this  for  the  Pacific  salmon 
fisheries.  The  second  argument  is  that  the 
open-access  tradition  insures  that  resource 
augmentation  through  publicly  operated  hatch- 
eries will  induce  additional  effort  into  the  fish- 
ery, especially  when  the  additional  inputs  are 
provided  without  cost  to  the  fishermen.  The 
resultant  equilibrium  levels  of  factor  returns, 
output  prices,  and  excess  capacity  may  differ 
from  initial  equilibrium  levels,  but  a  priori 
speculation  about  empirical  magnitudes  is  just 
that.  Furthermore,  the  time  pattern  of  adjust- 
ment and  the  distribution  of  benefits  and  costs, 
over  both  time  and  space,  can  be  discerned 
only  vaguely. 

We  would  maintain,  however,  that  resource 
augmentation  efforts  should  be  placed  in  per- 
spective with  the  total  institutional  setting.11 
Hatchery  contributions  to  fish  stocks  may  per- 
petuate the  tendency  toward  excess  harvesting 
capacity,  but  it  should  not  have  to  bear  the 
entire  burden  of  responsibility  for  economic 
and  social  ills  of  the  fishery.  The  tendency 
toward  excess  capacity  pervades  open-access 
fisheries,  most  of  which  do  not  rely  on  hatchery 
propagation.  It  would  be  our  guess  that  the 
magnitude  of  inefficiency  associated  with  the 
larger  issue  probably  overshadows  any  un- 
desirable effects  of  hatchery  production,  if  the 
latter  in  fact  exist. 

Having  confessed  that  we  do  not  have  all 
the  answers,  we  hasten  to  add  that  we  do  have 
some  empirical  observations  on  entry  and  exit, 
over  time,  of  salmon  harvesting  resources.  We 
view  these  not  as  definitive  proof  of  anything, 
but  as  a  piece  of  the  empirical  jigsaw  puzzle 


11  We    are   indebted    to   Emery   Castle   for  this   pers- 
pective. 


which  must  eventually  be  put  together  if  econ- 
omists are  to  be  looked  to  for  policy  advice. 


Entry  and  Exit  of  Resources  in  the 

Commercial  Salmon  Harvest: 

Fixed  Asset  Theory 

The  rise  and  fall  of  the  Pacific  Coast  salmon 
harvest  has  been  well  documented  elsewhere 
(Cooley,  1963  and  Crutchfield  and  Pontecorvo, 
1969).  Peak  harvest  years  were  reached  in 
the  1930's,  and  catch  has  trended  downward 
since  that  time.  The  quantity  of  resources  com- 
mitted to  the  fishery,  however,  has  increased 
over  time.  The  number  of  fishermen  and  the 
net  tonnage  of  vessels  increased  by  about  30% 
between  1947-1949  and  1964-1966  periods, 
total  landings  declined  by  about  25%,  and  the 
deflated  value  of  landings  per  fisherman  de- 
creased by  about  15%  (Table  4).  It  appears, 
however,  that  the  deflated  average  value  of 
landings  per  fisherman  has  remained  about 
constant  since  1950,  with  year-to-year  fluctu- 
ations. This  can  be  taken,  recognizing  the  limi- 
tations on  accuracy  of  the  data,  as  very  super- 
ficial evidence  of  the  open-access  phenomenon, 
i.e.,  the  dissipation  of  rents  through  entry  of 
additional  resources. 

Even  though  there  has  been  net  entry  into 
the  fishery  since  1947,  the  time  path  of  entry 
and  exit  of  harvesting  resources  has  not  been 
fully  explored.  In  particular,  is  there  a  degree 
of  symmetry  between  the  relationships  which 
explain  entry,  on  one  hand,  and  exit,  on  the 
other?  Miss  Peerarat  Aungurarat  attempted 
to  answer  this  question  in  another  portion  of 
our  Sea  Grant  research  at  Oregon  State  Uni- 
versity (1970).  Her  results  are  especially  inter- 
esting in  light  of  the  increased  reliance  on 
hatchery  programs. 

Conventional  firm  theory  suggests  that  a 
high  degree  of  symmetry  would  exist  in  ex- 
plaining entry  and  exit  of  resources.  Given  a 
constant  factor  price,  leftward  (rightward) 
shifts  in  the  marginal  value  product  function 
would  imply  a  reduction  (increase)  in  the 
utilization  of  a  factor  of  production.  Dissatis- 
faction with  the  state  of  the  arts  in  explaining 
the  inelastic  supply  of  agricultural  products 
led  Glenn  L.  Johnson  to  formulate  a  "fixed 
asset"    theory    (1958).    Johnson's    contribution 


140 


Table  4.  —  Salmon  fishing  effort,  quantity  of  landings  (pounds  and  values)  and  average  values  per  fishermen  in  Alaska, 

Washington,  and  Oregon,  1947-1966. 


Year 

Labor 
(number  of 
fishermen) 

Vessels 

(net 
tonnage) 

Landings 
(thousands 
of  pounds) 

Value  of 

i 

landings 

(thousands 

of  dollars) 

Average  landings 

per  fisherman 

(thousands 

of  pounds) 

Average  value 
1 

per  fisherman 
(thousands 
of  dollars) 

1947 

16,249 

44,003 

486.560 

47,541 

29.94 

2.92 

1948 

19,334 

59,443 

395,981 

43,222 

20.48 

2.24 

1949 
1950 

18,451 
19.241 

59,510 
63.156 

477,074 
321,575 

54,441 
42,464 

25.86 
16.71 

2.95 
2.21 

1951 
1952 

23.589 
22.318 

70.799 
71.842 

367,030 
344.999 

55,840 
46,960 

15.56 
15.46 

2.37 
2.10 

1953 

21.889 

69.231 

304,945 

38,500 

13.93 

1.76 

1954 

20.321 

66,742 

315.217 

43,925 

15.51 

2.16 

1955 

24.608 

69.268 

277.900 

39,389 

11.29 

1.60 

1956 

19.522 

63.869 

312.837 

44,651 

16.02 

2.29 

1957 

2 

2 

260.125 

38,580 

2 

2 

1958 
1959 

2 

19.990 

2 

58.099 

303.797 
194.915 

43.976 
32.221 

2 
9.75 

2 
1.61 

1960 

21.546 

53.285 

229.227 

40,146 

10.64 

1.86 

1961 

23,206 

63.060 

301,760 

45,421 

13.00 

1.96 

1962 

21.921 

62.767 

307,892 

49.649 

14.04 

2.26 

1963 

23,689 

66.553 

286.316 

41111 

12.09 

1.78 

1964 

22,384 

66.057 

342.765 

47.128 

15.31 

2.11 

1965 

23,486 

65.691 

317.068 

54.717 

13.50 

2.33 

1966 

24,987 

67.314 

378.066 

60.671 

15.13 

2.43 

1  Deflated  by  Consumer  Price  Index  (1957-59  =  100). 

2  Data  not  available. 

SOURCE:  Derived  from  Fishery  Statistics  of  the  United  States,  U.S.  Fish  and  Wildlife  Service.  Bureau  of  Commercial  Fisheries. 


was  his  recognition  of  a  particular  form  of 
imperfect  factor  markets,  and  involved  relaxing 
the  assumption  that  firms  or  industries  can 
at  the  same  price,  both  buy  and  sell  inputs. 
A  "fixed  asset",  by  Johnson's  definition,  is  not 
fixed  because  it  has  a  certain  physical  life  ex- 
pectancy, but  because  it  is  more  economical 
to  keep  it  in  production  than  to  sell  it.  Two 
factor  prices  are  involved,  i.e.,  an  acquisition 
price  and  a  salvage  value.  Applied  to  the  fish- 
ing industry,  the  acquisition  price  is  what  a 
fisherman  (or  the  industry)  has  paid  or  must 
pay  for  an  additional  productive  asset,  e.g.,  a 
vessel;  the  salvage  value  is  what  the  fisherman 
(or  industry)  could  derive  from  the  asset  if 
it  were  sold  rather  than  used.  For  individual 
fishermen,  the  difference  between  the  two 
prices  might  be  small  if  the  quality  of  assets 
is  assumed  constant.  For  the  salmon  industry 
or  even  a  particular  segment  of  the  industry, 
the  margin  might  be  substantial.  The  more 
specialized  the  gear  or  vessel,  the  less  one 
might  expect  to  derive  from  selling  it  to  an- 
other segment  of  the  industry. 


If  there  is  a  large  difference  between  ac- 
quisition price  and  salvage  value,  then,  it 
would  be  possible  for  no  change  to  occur  in 
the  aggregate  level  of  a  resource  even  if  there 
were  significant  changes  in  factor  productivity 
or  product  price.  Figure  3  illustrates  the 
variety  of  adjustments  that  could  conceivably 
take  place,  depending  upon  (a)  the  starting 
point,  (b)  the  magnitude  of  the  shift  in  the 
MVP  function,  and  (c)  the  divergence  between 
acquisition  price  and  salvage  value.  In  the 
absence  of  specific  knowledge  about  these 
factors,  the  notion  of  symmetry  between  exit 
and  entry  in  the  salmon  fishery  becomes  an 
empirical  question.  Fixed  asset  theory,  how- 
ever, does  provide  a  conceptual  framework  for 
specifying  a  statistical  model  and  interpreting 
the  results. 

Empirical  Analysis 

In  that  we  had  access  only  to  secondary 
data  (U.S.  Department  of  the  Interior,  1947- 
67),  most  of  the  variables  in  the  analysis  were 


141 


Aggregate  Input  Level 


Starting 

Point 

Parameter 

Change 

Given 

1 

MVPQ  ^ 

MVP 

PA,  PS 
0'   0 

MVP  -» 

MVP„ 
0 

pA  pS 
0   0 

2 

MVP  -» 

MVP 

„A   „S 

V  po 

MVPo- 

MVP 

PA,  Ps 

0'   0 

3 

MVP0-» 

MVP 

„A    S 

po-  po 

MVP  -T 

MVP 

PA,  Ps 

0'   0 

implies : 


Entry 


1-T  2 


3^  2 
3-*-2 


2  J»3 
2-,   3 


(no  change) 


Figure    3.    —    Expected    factor    adjustments,    given    alternative 
assumptions  on  key  parameters. 


proxy  variables.  Additionally,  the  quality  of 
Bureau  of  Commerical  Fisheries  historical 
data  on  resource  levels  in  specific  fisheries  is 
far  from  perfect.  A  major  data  limitation  of 
this  study  was  that  it  was  not  possible  to 
separate  full-time  from  part-time  commercial 
fishermen. 

The  secondary  data  precluded  any  meaning- 
ful estimation  of  marginal  factor  productivities. 
Also,  reliable  data  on  factor  acquisition  prices 
or  salvage  were  not  available.  The  first  prob- 
lem was  bypassed  by  means  of  three  assump- 
tions; the  second  was  resolved  by  the  choice 
of  units  of  observations.  Both  require  some 
explanation. 

First,  it  was  assumed  that  the  demand  for 
salmon   is  price-elastic   at  the  ex-vessel   level. 


Some  support  for  this  assumption  comes  from 
two  studies  conducted  at  Oregon  State  Uni- 
versity under  the  supervision  of  Dr.  R.  S. 
Johnston  (Charoenkul,  1970  and  Wood,  1970). 
Second,  it  was  assumed  that  for  the  salmon 
fishery  as  a  whole,  the  supply  of  factors  is  es- 
sentially fixed  prior  to  the  fishing  season.  The 
direct  implication  of  these  two  assumptions 
is  that  increases  (decreases)  in  landings  bring 
about  increases  (decreases)  in  average  short- 
run  rents  and/or  profits  over  the  industry. 
The  third  assumption  was  that  actual  rents 
in  the  year  t  equal  expected  rents  in  the  year 
t  +  1,  ceteris  paribus.  The  expectation  of  cyclical 
fish  runs  should  be  accounted  for  empirically, 
since  ceteris  paribus  is  not  a  realistic  assump- 
tion in  areas  with  pink  and  sockeye  salmon. 


142 


These  assumptions,  in  context  with  the  earlier 
discussion  of  fixed  asset  theory,  imply  a  statis- 
tical model  wherein  changes  in  resource  quant- 
ity (labor  or  vessels)  are  regressed  on  changes 
in  salmon  landings,  lagged  by  one  year.  Ideally, 
the  influence  on  resource  use  levels  of  acquisi- 
tion prices  and  salvage  values  of  the  produc- 
tive factors  should  also  be  taken  into  account. 
In  that  these  data  were  not  available,  the  units 
of  observation  were  defined  both  cross-section- 
ally  and  over  time.  Specifically,  yearly  data 
between  1957  and  1966  for  each  of  ten  NMFS 
statistical  regions  on  the  Pacific  Coast  were 
used.12  This  yielded  a  total  of  80  observations 
and  allowed  us  to  take  into  account,  in  a  rough, 
implicit  fashion,  cross-sectional  differences 
which  might  give  rise  to  a  variety  of  deviations 
between  acquisition  prices  and  salvage  values. 
The  statistical  model  is  as  follows: 

(1)  X  (,  +  | )  =  f  [L{t),  C(f+,  ),  £/(,+  d,  D] 
where 

X'  =  index  of  fishing  effort  (number 
of  fishermen  and  net  tonnage 
of  fishing  vessels), 

L     =    index  of  salmon  landings 
(pounds), 

C  =  cyclical  nature  of  the  fishery 
(dummy  variable:  1  for  all  ex- 
pected good  runs,  whether  or 
not  they  actually  materialized, 
and  0  for  all  expected  poor  runs), 

U  =  unemployment  rate  of  the  civil- 
ian labor  force  in  the  major 
labor  market, 

D  =  distance  from  the  center  of  sal- 
mon fishing  activity  in  the 
region  to  the  nearest  major 
labor  market. 

In  order  to  test  for  symmetry  between  exit 
and  entry   relationships  with   this  model,   the 


12  The  regions  were  Southeastern,  Central,  and  West 
ern  Alaska;  Puget  Sound  and  Coastal  in  Washington; 
Columbia  River  in  Washington  and  Oregon;  Coastal 
Oregon;  and  Northern,  San  Francisco  and  Monterey 
in  California  (U.S.  Department  of  the  Interior,  1947- 
1967). 


80  observations  were  divided  into  two  subsets. 
One  subset,  with  42  observations,  consisted  of 
those  years  in  which  landings  had  increased 
over  the  preceding  year.  Given  our  assump- 
tions of  an  elastic  product  demand,  fixed  factor 
supply  (in  the  short  run),  and  rent  expecta- 
tions, it  follows  that  these  observations  repre- 
sent years  in  which  the  MVP  schedule  of  factors 
had  shifted  to  the  right  and  was  expected  to 
remain  there,  ceteris  paribus.  Similarly,  the 
35  observations13  in  the  second  subset  repre- 
sented years  in  which  MVP  had  shifted  left- 
ward. 

Fixed  asset  theory  would  suggest  that  aggre- 
gate factor  levels  in  an  industry  would  either 
increase  or  remain  constant  following  years 
of  increased  landings,  and  would  either  de- 
crease or  remain  constant  following  years  of 
reduced  landings.  Table  5  indicates  a  definite 
asymmetry  between  entry  and  exit  relation- 
ships. For  example,  in  years  of  increased  land- 
ings, the  index  of  vessel  inputs  in  year  t  + 1 
increased  0.32  per  unit  increase  in  the  land- 
ings index  for  year  t.  The  coefficient  for  years 
of  decreased  landings  was  very  slightly  nega- 
tive, but  not  significantly  different  from  zero. 
Asymmetry  is  strongly  suggested  by  the  fact 
that  the  B\  coefficients  for  the  two  subsets  are 
significantly  different  from  each  other  in  both 
the  labor  and  vessel  equations. 

This  ratchet  mechanism  is  illustrated  in 
Figure  4.  Net  entry  follows  years  of  "good 
catches,"  but  net  exit  does  not  occur  following 
the  "bad  years".  This  is  not  hard  to  imagine 
for  specialized  trolling  vessels  which  may  have 
low  salvage  values  outside  of  fishing  or  even 
in  other  segments  of  the  salmon  fishery.  It 
is  somewhat  more  difficult  to  rationalize,  on 
the  other  hand,  for  the  labor  resource,  although 
the  human  resource  would  no  doubt  be  af- 
fected by  lack  of  mobility  of  the  capital 
resource. 

The  relationships  of  resource  use  levels  to 
the  other  variables  in  the  analysis  are  also 
of  interest,  and  are  summarized  here: 

(1)  Expectations  of  cyclical  runs  in  encour- 
aging entry  were  more  important  follow- 
ing years  of  declining  landings  than  fol- 
lowing years  of  increased  landings.  This 


13  Three    observations    were    not   usable   due   to   lack 
of  a  "bench  mark"  year. 


143 


Table  5.  —  Regression  analysis  of  factors  affecting  resource  use.1 


Dependent  variable 


B 


B 


(LJ 


B 


Vm> 


B 


Wt+i>  <D> 


R2 


All  years: 
Labor 


Vessels 


Years  of  Increased  Landings: 
Labor 


Vessels 


Years  of  Decreased  Landings: 
Labor 


Vessels 


85.88 


89.95 


77.33 


84.43 


102.90 


99.39 


+0.19 

+  6.48 

-1.00 

+0.009 

(3.40) 

(1.15) 

(-0.68) 

(0.51) 

+0.19 

+  6.25 

-1.89 

+0.02 

(3.25) 

(1.06) 

(-1.22) 

(0.93) 

+0.31 

-2.41 

-2.51 

+0.04 

(3.26) 

(-0.26) 

(-1.17) 

(1.24) 

+0.32 

-6.41 

-4.46 

+0.05 

(3.26) 

(-0.68) 

(-1.99) 

(1.95) 

+0.03 

+  10.77 

-1.37 

+0.005 

(0.43) 

(1.62) 

(-0.64) 

(0.25) 

-0.002 

+  1.15 

-0.19 

-0.005 

(-0.003) 

(1.89) 

(-0.09) 

(-0.25) 

0.14 


0.13 


0.24 


0.28 


0.08 


0.13 


42 


42 


35 


35 


Variables  are  as  defined  in  text.  Parentheses  contain  "^-values"  of  the  regression  coefficients. 


t+1     Resources 


(E  =  expected  MVP  in  year  t 
A   =  actual  MVP  in  year  t_  ) 


Figure    4.   —    Asymmetry    between    entry    and    exit    of 
resources. 


may   be   somewhat   spurious   due   to   the 
2-year  cycle  of  pink  salmon. 
(2)  Increased  unemployment  rates  in  major 
labor    markets    reduced    entry    into    the 


fishery,  especially  in  years  of  increased 
landings  when  the  incentive  to  enter 
would  have  been  highest. 
(3)  Increased  distance  from  major  labor  mar- 
kets had  a  positive  relationship  to  the 
index  of  resource  use,  and  was  relatively 
more  significant  in  years  of  increased 
landings.  In  retrospect,  both  distance  and 
unemployment  rates  might  contribute 
more  to  an  explanation  of  the  B\  coeffi- 
cient   which    related    resource    levels    to 

landings  r-        dXr+ 1  i  if  these  coefficients 

LB|  ~  bLt 
could  be  estimated  for  each  district,  rather 
than  the  overall  fishery.  Our  data  did  not 
permit  this  to  be  done. 

Policy  Implications 

Although  this  analysis  was  fairly  superficial 
because  of  the  reliance  on  secondary  data,  it 
did  indicate  that  entry  of  resources  is  systemati- 
cally related  to  profit  expectations  based  on  an 
increasing  level  of  aggregate  landings.  The 
same  may  be  said  for  exit  from  the  fishery  if 
"systematic"  is  interpreted  in  terms  of  consis- 
tency with  fixed  asset  theory.  The  empirical 


144 


values  by  which  entry  and  exit  are  systemati- 
cally related  to  profit  expectations,  however, 
differ  markedly. 

This  policy  implication  for  augmenting  sal- 
mon stocks  through  hatchery  programs  is  ap- 
parent; it  is  evidently  easier  to  induce  re- 
sources into  the  fishery  than  to  induce  them 
to  leave.  If  the  real  social  objective  of  hatchery 
programs  relates  to  improving  incomes  in  the 
fishery,  rather  than  producing  and  catching 
fish,  research  and  action  programs  designed 
to  increase  salvage  values  of  labor  and  capital 
resources  would  seem  to  be  of  a  high  priority. 


COOLEY,  RICHARD  A.,  1963.  Politics  and  Conser- 
vation, The  Decline  of  the  Alaska  Salmon.  Harper 
and  Row. 

CRUTCHFIELD,  J.  A.  and  G.  PONTECORVO,  1969. 
The  Pacific  Salmon  Fisheries,  The  Johns  Hopkins  Press. 

JOHNSON,  G.  L.,  1958.  Supply  Function:  Some  Facts 
and  Notions.  IN:  Agricultural  Adjustment  Problems 
in  a  Growing  Economy.  Iowa  State  College  Press. 

MATTOX,  BRUCE  W.,  1970.  A  Partial  Economic 
Analysis  of  Hatchery  Propagation  and  Commerical 
Harvest  of  Salmonid  Resources  in  Oregon.  Unpublish- 
ed Ph.D.  Dissertation.  Oregon  State  University. 


LITERATURE  CITED 

AUNGURARAT,  PEERARAT,  1970.  An  Analysis  of 
Factors  Affecting  Resource  Usage  in  the  Pacific  Coast 
Salmon  Fishery.  Unpublished  M.S.  Thesis.  Oregon 
State  University. 


McNEIL,  WILLIAM  J.,  1969.  Survival  of  Pink  and 
Chum  Salmon  Eggs  and  Alevins.  IN:  T.  G.  Northcote. 
Symposium  on  Salmon  and  Trout  in  Streams:  H.  R. 
MacMillan  lectures  in  Fisheries.  The  University  of 
British  Columbia. 

PALOHEIMO,  J.  E.  and  L.  M.  DICKIE,  1966.  Food 
and  Growth  of  Fishes  II.  Effects  of  Food  and  Tempera- 
ture on  the  Relation  Between  Metabolism  and  Body 
Weight.  Journal  of  the  Fisheries  Research  Board  of 
Canada. 


BROMLEY,  D.  W.,  1969.  Economic  Efficiency  in  Com- 
mon Property  Natural  Resource  Use;  A  Case  Study 
of  the  Ocean  Fishery.  Working  Paper  No.  28,  Division 
of  Economic  Research,  National  Marine  Fisheries 
Service,  U.S.  Department  of  Commerce. 

CHAROENKUL,  VILAILUCK,  1970.  Analysis  of  De- 
mand for  Canned  Pink  Salmon.  Unpublished  M.S. 
Thesis.  Oregon  State  University. 

CHRISTY,  F.  M.  and  A.  SCOTT,  1965.  The  Common 
Wealth   in   Ocean  Fisheries.  The  Johns  Hopkins  Press. 


U.S.  Department  of  the  Interior,  1947-1967.  Fish  and 
Wildlife  Service.  Bureau  of  Commercial  Fisheries. 
Fishery  Statistics  of  the  United  States. 

WAHLE,    ROY    J.,    1970.    Salmon    Hatcheries,    An    En- 
couraging    Supplement    to    a     Pacific     Coast     Fishery. 
Paper  presented  to  the  Western  Division  of  the  Ameri- 
can Fisheries  Society,  Victoria,  British  Columbia. 

WOOD,  WILLIAM  R.,  1970.  A  Demand  Analysis  of 
Processed  Salmon  from  the  West  Coast.  Unpublished 
M.S.  Thesis.  Oregon  State  University. 


145 


Limited  Entry :  The  Case  of  the  Japanese  Tuna  Fishery 


E.  A.  Keen1 


ABSTRACT 

Limited  entry  has  been  advocated  strongly  as  an  important  but  as  yet  usused  man- 
agement tool  for  U.S.  fisheries.  Japan  has  maintained  a  policy  of  limiting  entry  into 
its  high  seas  fisheries  since  1949  and  thus  has  considerable  experience  of  potential 
value  to  the  use  of  this  tool  in  U.S.  fisheries.  This  paper  presents  an  assessment  of  the 
limited  entry  system  as  it  has  been  developed  for  the  Japanese  tuna  fisheries.  At- 
tention is  given  to  effects  on  the  acquistion  of  capital  and  overall  allocation  of  national 
resources,  specific  effects  on  the  size  and  nature  of  the  fleet,  pressures  to  permit  ad- 
ditional entry,  and  effects  on  the  location  of  shore-based  activities.  Special  attention 
is  given  to  problems  that  were  unforeseen  at  the  time  of  the  initiation  of  limited  entry 
that,  with  experience,  could  have  been  avoided.  The  paper  is  based  largely  on  field 
research  conducted  in  1963  and  1964. 


INTRODUCTION 

Limitation  of  the  number  of  craft  in  a  fishery 
has  been  advocated  strongly  as  a  management 
tool  for  American  fisheries.  The  volume  of 
literature  in  which  its  usefulness  is  analyzed, 
primarily  by  economists,  has  become  substan- 
tial and  continues  to  grow.  A  brief  survey  of 
work  by  Crutchfield,  Scott,  Christy  and  others 
readily  convinces  the  reader  that  economic 
benefits  to  be  gained  through  its  use  more 
than  justify  its  advocates.  In  the  case  of  the 
extremely  crowded  northeastern  Pacific  salmon 
fishery,  limitation  of  entry  appears  to  be  al- 
most mandatory  if  rational  management  only 
for  maximum  sustained  yield  from  the  phys- 
ical stocks  is  to  be  attained.  Whether  one  is 
concerned  with  maximum  sustained  yield  or 
with  maximum  economic  return,  limitation 
of  entry  obviously  is  a  powerful  tool  and  one 
that  deserves  greater  use. 

As  with  all  powerful  tools,  implementation 
and  operation  of  a  limited  entry  system  just 
as  obviously  is  not  an  easy  matter.  Fisheries 
cannot  be  considered  apart  from  the  highly 
complex  human  and  physical  systems  with 
which  they  are  intertwined.  Foreseeing  all 
effects  of  a  major  change  in  regulatory  inputs 


is  extremely  difficult.  Decisions  once  made 
and  institutionalized  are  equally  difficult  to 
change.  In  light  of  the  complexity  of  fisheries 
and  of  the  difficulty  with  which  mistakes  can 
be  corrected,  it  behooves  those  who  would 
design  and  implement  a  system  of  limited  entry 
to  take  advantage  of  actual  experience  in  other 
fisheries  to  the  extent  possible. 

The  purpose  of  this  paper  is  to  explore  the 
experience  of  the  Japanese  with  limitation  of 
entry  into  one  of  their  major  fisheries,  the 
skipjack-tuna  fishery.2  Much  of  this  experience 
is,  of  course,  specific  to  this  fishery  and  is 
therefore,  only  indirectly  relevant  to  other 
fisheries  in  Japan  or  elsewhere.  Many  of  the 
problems  grew  out  of  the  needs  of  a  rapidly 
expanding  fishery,  a  condition  that  is  not 
likely  to  occur  too  frequently  in  the  future. 
However,  some  generalizations  can  be  drawn 
from  it  that  can  be  of  use  in  management  of 
a  number  of  fisheries.  A  brief  summary  of  the 
initiation  and  development  of  the  regulatory 
system  is  presented  first  to  show  the  complex- 
ity of  its  development.  This  provides  back- 
ground for  a  discussion  of  the  major  effects, 
favorable  and  unfavorable,  that  concludes  the 
paper. 


1   Associate  Professor  of  Geography,  California  State 
University,  San  Diego. 


2  The  term  "skipjack-tuna  fishery"  is  a  direct  trans- 
lation of  the  Japanese  term  "Katsuo-maguro  gyogyo." 
All  species  of  tuna  are  sought  by  those  in  the  fishery, 
not  the  skipjack  alone  as  the  translation  might  apply. 


146 


INITIATION  OF  THE 
REGULATORY  SYSTEM 

Basic  aspects  of  the  system  of  limited  entry 
were  set  by  a  series  of  administrative  ordi- 
nances and  laws  passed  during  the  Allied 
Occupation  of  Japan.  An  administrative  order 
issued  in  July  1946  required  registration  of 
all  skipjack-tuna  craft  over  20  gross  tons  in 
size  as  an  aid  to  limit  the  operation  of  these 
craft  to  areas  designated  by  the  Occupation 
Government.3  An  ordinance  issued  by  the 
Fisheries  Agency  in  July  1947  brought  these 
craft  under  a  formal  licensing  system  and 
forbade  the  construction  of  additional  craft. 
Licenses  were  issued  to  all  owners  of  craft 
over  20  tons  for  the  gross  tonnage  of  their 
existing  craft.  An  ordinance,  issued  in  May 
1949,  regularized  the  licensing  system,  made 
provision  for  building  larger  craft  by  combin- 
ation of  the  licensed  tonnage  of  two  or  more 
craft,  and  limited  the  activities  of  craft  en- 
gaged in  the  skipjack-tuna  fishery  on  a  seasonal 
basis.  The  essence  of  these  ordinances  were 
all  codified  into  a  new  basic  fisheries  law 
passed  by  the  National  Diet  in  November  1949. 
An  important  additional  measure  included  in 
the  new  law  was  that  licenses,  while  issued 
for  periods  of  5  years,  had  to  be  reissued  to 
the  original  holder  or  his  heirs  except  in  cases 
of  serious  infraction  of  laws  on  the  part  of 
the  holder.  It  also  created  a  new  category  of 
fisheries,  called  Designated  Distant  Sea  Fish- 
eries, into  which  all  skipjack-tuna  craft  of  over 
100  tons  in  size  were  placed.  A  separate  fish- 
eries protection  law  passed  by  the  Diet  in  1950 
set  a  limit  of  300  skipjack-tuna  vessels  in  the 
Designated  Deep  Sea  category. 

Conditions  were  favorable  to  establishment 
of  the  system  during  the  few  years  over  which 
it  evolved.  The  administrative  order  and  the 
basic  regulatory  law  were  established  at  a 
time  when  profits  from  the  fishery  were  low 
or  nonexistent.  In  the  first  years  of  the  Occu- 
pation, the  Japanese  were  anything  but  prone 
to  resist  rules  issued  in  the  name  of  the  con- 
quering powers.  The  fleet  had  been  heavily 
decimated  during  the  war  but  recovery,  with 
encouragement  of  the  Occupation  Government, 


3  An  excellent  treatment  of  the  regulatory  system 
as  it  developed  up  to  1962  appears  in  Masuda  (1963). 
All  tonnage  figures  used  herein  refer  to  metric  tons. 


was  rapid  afterward.  By  the  end  of  1947,  the 
fleet  had  recovered  to  its  approximate  prewar 
size  and  was  more  than  adequate  to  harvest 
resources  within  the  area  enclosed  by  the  so- 
called  MacArthur  Line.4  Catch  per  unit  of 
effort  had  fallen  off  rapidly  with  the  increase 
in  numbers  of  craft  and  little  opposition  was 
expressed  to  institution  of  the  regulatory  sys- 
tem. Those  who  already  owned  craft  in  the 
fishery,  of  course,  stood  to  profit  by  limita- 
tion of  entry  and  supported  it.  The  low  rates 
of  return  of  the  fishery  discouraged  outsiders 
from  protesting  because  entry  was  forbidden 
to  them.  The  system  imposed  no  onerous  re- 
strictions on  fishing  effort,  such  as  closed 
seasons  or  closed  areas  within  the  fishing 
grounds  available  to  the  fleet.  It  appears  to 
have  been  accepted  fairly  readily  by  the  fishing 
community  and  functioned  without  change 
until  near  the  end  of  the  Occupation  in  April 
1952. 

Several  factors  were  put  forth  to  support 
imposition  of  the  system  during  its  develop- 
ment. However,  the  main  motivations  for  estab- 
lishment of  the  limited  entry  system  centered 
on  conditions  in  the  fishery  at  the  time,  not 
on  the  condition  of  the  resource.  That  is  to 
say,  conservation  or  management  of  the  re- 
source was  not  a  real  issue.  It  was  an  issue 
and  an  important  one  in  controlling  entry 
into  the  East  China  Sea  trawl  fishery  which 
was  placed  under  a  limited  entry  system  at 
the  same  time  as  the  skipjack-tuna  fishery. 
Concern  growing  out  of  the  serious  overfishing 
by  the  East  China  Sea  fleet  undoubtedly  in- 
fluenced the  lawmakers  in  their  decision  to 
bring  the  skipjack-tuna  fleet  under  control 
and  to  limit  the  number  of  vessels  over  100 
tons  to  300.  However,  the  skipjack-tuna  fleet 
exploited  species  that  migrated  over  great 
distances  and  showed  no  signs  of  depletion 
from  year  to  year  because  of  overfishing  in 
waters  off  Japan  used  by  the  fleet.  Sufficient 
fish  might  not  be  available  to  support  the 
fleet  during  that  part  of  their  migration  that 
made    them    available    to    the    Japanese    fleet, 

4  The  MacArthur  Line,  as  the  line  bounding  the  area 
open  to  Japanese  fisheries  that  was  established  by  the 
Occupation  Government  came  to  be  known,  originally 
included  only  the  waters  within  12  miles  of  Japan. 
However,  it  was  gradually  expanded  eastward  and  south- 
ward and  by  1950,  included  most  of  the  traditional 
Japanese  skipjack  and  tuna  ground  in  the  northwest 
quadrant  of  the  Pacific. 


147 


but  little  evidence  existed  to  suggest  that  re- 
duction of  the  stocks  in  any  one  year  serious- 
ly reduced  the  runs  the  following  year.  Thus, 
the  main  reasons  were  to  prevent  overcrowd- 
ing and  conflict  on  the  fishing  grounds  and 
to  maintain  economic  viability  of  the  individual 
fishing  enterprise.  This  latter  reason  was  to 
become  clearly  the  overwhelming  one  in  sub- 
sequent years. 


DEVELOPMENT  AFTER  THE 
OCCUPATION  PERIOD 

If  the  system  was  accepted  and  proved  ade- 
quate as  it  stood  during  the  first  years  of  its 
effect,  it  patently  was  going  to  require  modi- 
fication after  Japan  regained  full  sovereignty. 
As  stated  above,  the  fleet,  both  in  reference 
to  numbers  and  size  of  craft,  was  more  than 
adequate  to  harvest  resources  in  the  area  to 
which  it  had  been  restricted  by  the  Occupation 
Government.  However,  Japanese  tuna  fisher- 
men had  begun  to  open  up  tuna  grounds  in 
the  west  central  Pacific  and  East  Indies  waters 
prior  to  World  War  II.  Catch  rates  had  been 
high,  the  resource  was  known  to  be  large  and 
many  were  anxious  to  return  to  these  grounds 


denied  them  during  the  Occupation.  To  do 
so,  larger  vessels  were  desirable;  the  resource 
could  support  a  larger  fleet  than  existed  in 
1952.  Pressures  developed  to  permit  expan- 
sion of  the  fleet  —  internal  pressure  from 
existing  license  holders  to  build  larger  vessels, 
external  pressure  from  nonlicense  holders  for 
permission  to  enter  the  fishery. 

The  following  decade  was  marked  by  con- 
tinual modification  of  the  regulatory  system 
as  the  fishery  expanded  beyond  the  most  san- 
guine anticipations  of  anyone  connected  with 
it  in  the  early  1900's  (see  Figure  1).  The  1949 
fishery  law  was  explicit  as  to  the  number  of 
craft  that  could  be  licensed,  the  1950  law  as 
to  the  number  that  could  be  larger  than  100 
tons.  The  upper  limit  of  300  craft  over  100 
tons  in  size  had  already  been  approached.  The 
only  expansion  possible  without  a  new  law 
from  the  National  Diet  was  of  tonnage  within 
the  framework  of  the  existing  law.  Subsequent 
laws  and  administrative  orders  based  on  them 
were  numerous  and  increasingly  complex.  No 
attempt  will  be  made  to  treat  all  of  these  in 
detail;  to  do  so  would  become  extremely  tedious. 
However,  the  first  two  are  covered  in  some 
detail  to  show  the  pattern  set  for  expansion 
of  the  fleet. 


THOUSAND 
TONS 


800 

700 

600 

/           Total         \ 
/ 

500 

^           — ' 

400 

/ 

/ 
s 

300 

y' 

--" 

"*                              y       Long  Line 

200 

/ — 
/ 
/ 

-"""""                               /\ 

^> 

,..--' 

_^_, .       . ^^ ' 

100 

—              Pole  and  Line 

1    1    1    1    1    1    1    1    1    1    1    1    1    1    1    1 

51   52   53   5U   55  56   57  58   59  60   61   62   63   6U   65   66  67 

Figure  1.  —  Landings  of  tuna  and  other  species  by  skipjack  pole-and  line 
craft  and  by  tuna  longliners,  1951-67.  Data  for  1951-1961  from  Masuda 
(1962,  p.  361),  and  for  1962-1967  from  Japanese  Tuna  Fisheries  Federa- 
tion (1968  and  1969). 


148 


The  first  measure  for  expansion  was  con- 
tained in  an  administrative  order  from  the 
Fisheries  Agency  issued  in  March  1952.  This 
order  permitted  enlargement  of  vessels  by  a 
combination  of  free,  additional,  licensed  tonnage 
and  licensed  tonnage  from  decommissioned 
existing  craft.  The  owner  of  a  Designated  Dis- 
tant Seas  craft,  i.e.,  one  over  100  tons  in  size, 
could  build  a  vessel  40  tons  larger  than  the 
existing  one  without  withdrawing  additional 
tonnage  from  another  license.  If  the  new  vessel 
were  between  40  and  100  tons  larger  than  the 
original,  a  50-ton  vessel  had  to  be  withdrawn 
from  the  fleet;  if  a  new  vessel  100-200  tons 
larger  than  the  original  were  desired,  two 
50-ton  or  one  50-  to  100-ton  vessel  had  to  be 
withdrawn.  A  similar  system  was  set  up  for 
the  "medium-sized"  vessels  as  vessels  in  the 
20-  to  100-ton  category  had  come  to  be  called. 
The  legal  requirement  that  these  craft  be  less 
than  100  tons  cramped  measures  to  enlarge 
them  but  a  graduated  system  of  free  and  de- 
commissioned tonnage  was  instituted.  Any 
vessel  could  be  enlarged  up  to  10  tons  with 
no  restriction  but  half  of  any  enlargement 
over  this  had  to  come  from  vessels  withdrawn 
from  the  fleet.  Any  permitted  enlargement  as- 
sumed, of  course,  that  the  new  vessel  was  to 
be  less  than  100  tons  in  size.  This  technique 
of  granting  limited  free  tonnage,  to  be  com- 
bined with  tonnage  withdrawn  from  other 
vessels,  became  integral  to  the  regulatory  sys- 
tem during  the  ensuing  decade. 

The  March  1952  measure  was  inadequate 
to  meet  pressures  for  enlargment  of  vessels 
in  the  existing  fleet  and  did  nothing  to  meet 
pressure  to  permit  additional  entry.  This  latter 
pressure  was  especially  strong  from  fishermen 
in  the  offshore  trawl  fisheries,  the  resources 
for  which  were  judged  to  be  exploited  excessive- 
ly. The  expanding  tuna  fishery  appeared  to 
offer  an  opportunity  for  relief  for  these  fisheries. 
The  apparent  need  for  additional  tuna  vessels 
could  be  met  by  permitting  transfer  to  the 
tuna  fishery. 

These  conditions  led  rather  rapidly  to  modi- 
fication of  aspects  of  the  1949  fisheries  law 
that  related  to  the  fishing  power  of  the  tuna 
fleet.  The  National  Diet  passed  a  law  that  be- 
came effective  in  July  1953  and  that,  for  two 
years,  set  aside  aspects  of  the  1949  laws  that 
limited  the  size  and  number  of  vessels  in  the 


fleet.  Under  the  new  law,  known  as  the  Ex- 
ceptional Measures  Law,  craft  already  in  the 
fleet  were  divided  into  four  size  categories 
based  on  their  size  as  of  December  1952.  Li- 
censed craft  between  20  to  70  tons  were  per- 
mitted to  go  to  100  tons,  those  between  70 
and  95  tons  to  135  tons,  those  between  95  and 
100  tons  to  150  tons,  and  those  over  100  tons 
to  enlarge  with  no  limitations.  Owners  of 
licenses  for  the  medium-sized  craft  complained 
strongly  that  the  permitted  increases  were  not 
adequate.  In  April  1954,  the  upper  limits  for  70- 
to  90-ton  craft  and  for  90-  to  100-ton  craft  were 
rasied  to  160  and  180  tons  respectively.  The 
2-year  moratorium,  however,  was  not  extended 
beyond  its  original  July  1955  termination  date. 

Pressure  for  additional  entry  was  also  vented 
somewhat  by  the  2-year  law.  Originally,  it 
permitted  issuance  of  100  full-time  and  240 
part-time  skipjack  tuna  licenses.  This  aspect, 
too,  was  revised  further  in  April  1954.  New 
licenses  were  granted  for  120  skipjack-tuna 
craft  up  to  85  tons  in  size,  for  10  craft  be- 
tween 85  and  100  tons  in  size,  and  for  150 
part-time  licenses  of  less  than  85  tons.  These 
licenses  were  granted  to  craft  owners  in  cer- 
tain fisheries  deemed  to  be  overcrowded,  pri- 
marily the  offshore  trawl  and  purse-seine 
fisheries.  Recipients  in  all  cases  had  to  agree 
to  give  up  their  right  to  fish  in  their  original 
fishery  and  to  withdraw  their  craft  from  it. 

The  Exceptional  Measures  Law  resulted  in 
a  much  larger  and  greatly  changed  fleet.  Be- 
tween December  1953  and  December  1955, 
the  number  of  licensed  craft  increased  from 
1,154  to  1,372  or  19%  ;  gross  tonnage  increased 
from  112,945  tons  to  176,026  tons  or  57% ; 
and  craft  over  100  tons  in  size  increased  from 
290  to  621  (Masuda,  1963,  p.  354).  The  1950 
limitation  to  300  craft  of  over  100  tons  had 
obviously  been  abandoned. 

Fundamental  changes  had  also  taken  place 
in  the  nature  of  many  of  the  craft.  If  defined 
by  fishing  method,  the  skipjack-tuna  fishery 
is  actually  two  fisheries,  the  skipjack  live  bait 
pole-and-line  fishery  and  the  tuna  longline 
fishery.  Historically,  the  pole-and-line  fishery 
is  the  older  of  the  two.  It  developed  to  exploit 
the  large  runs  of  skipjack  and  to  a  lesser 
extent,  albacore,  that  appear  off  Japan  during 
the  spring  and  summer  months.  The  longline 


149 


fishery  developed  as  on  offseason  activity  for 
craft  in  the  former  and  remained  subordinate 
to  it  until  the  end  of  the  Allied  Occupation. 
Equipment  and  crew  requirements  for  the 
two  bear  little  similarity.  The  maximum  sized 
craft  that  could  be  used  efficiently  in  the  pole- 
and-line  fishery  was  about  150  tons  at  the 
time.5  Live  bait  wells  are  an  absolute  essential 
for  the  pole-and-line  fishery  but  are  unneces- 
sary for  the  longline  fishery.  Crew  size  for 
the  former  is  usually  a  little  more  than  double 
that  needed  for  the  longline  fishery  with  con- 
sequent additional  space  required  for  quarters. 
The  world  market  for  tuna  grew  rapidly  after 
World  War  II  and  tuna  soon  provided  a  higher 
return  than  did  skipjack.  Larger  craft  could 
operate  year  round  on  the  new  longline  grounds 
being  opened  up  in  the  southern  Pacific  and 
Indian  Oceans.  As  a  consequence,  most  of  the 
craft  built  when  the  Exceptional  Measures 
Law  was  in  effect  and  afterward  were  special- 
ized vessels  for  the  longline  fishery  only.  Lack 
of  a  live  bait  well  alone  effectively  denied 
their  use  in  the  pole-and-line  fishery. 

Landings  of  the  fishery  increased  propor- 
tionately along  with  the  tonnage  of  the  fleet. 
Tuna  longliners  landed  117,000  tons  in  1952; 
in  1955  this  had  increased  to  197,000  tons 
(Japanese  Tuna  Fisheries  Federation,  1961, 
p.  16).  The  value  of  the  landings  fell  rapidly; 
the  average  price  of  yellowfin  tuna  at  Yaezu, 
Japan's  most  important  tuna  port,  dropped 
from  $289  per  ton  in  1953  to  $192  in  1955 
(Yaezu  Fishery  Cooperative,  1963,  p.  25). H 
Lingering  effects  of  the  Bikini  nuclear  weapon 
incident  of  1954  that  had  greatly  reduced  de- 
mand for  fresh  tuna  in  Japan  accounts  in 
part  for  the  lower  price.  However,  the  main 
reason  was  excessive  supply.  The  world  mar- 
ket for  tuna,  limited  at  the  time  largely  to 
Japan  and  the  United  States,  was  not  able  to 
absorb  the  added  catch  at  the  1953  price  levels. 

The  Fisheries  Agency  policy  with  the  end 
of  the   Exceptional   Measures   Law   called   for 


5  A  vessel  of  about  150  tons  is  the  minimum  sized 
vessel  needed  to  operate  from  Japan  on  the  west-central 
Pacific  grounds  to  which  the  pole-and-line  fishery  ex- 
panded in  the  mid-1960's.  In  1967,  forty-one  vessels  in 
the  200-500  ton  category  were  used  in  the  newly  de- 
veloped distant  seas  pole-and-line  fishery  (Japanese 
Tuna  Fisheries  Federation,  1969,  p.  13). 

H  Conversions  from  yen  to  dollars  was  made  at  the 
rate  of  360  to  1. 


absolute  restrictions  on  new  entry.  However, 
it  did  continue  the  policy  of  permitting  and 
encouraging  enlargement  of  craft.  In  a  few 
cases,  slight  enlargements  were  permitted 
without  abolishment  of  licensed  craft.  The 
heart  of  the  policy,  however,  was  to  permit 
use  of  licensed  tonnage  for  medium-sized 
vessels  for  combination  with  other  licenses  to 
build  larger  craft.  The  net  effect  of  this  was 
to  reduce  the  total  number  of  craft  but  to  in- 
crease the  number  of  larger  craft  for  operation 
on  distant  grounds.  The  rapid  increase  in 
vessels  over  200  tons  at  the  expense  of  those 
under  that  size  is  shown  graphically  in  Figure 
2.  The  total  number  of  licensed  craft  decreased 
from  1,380  in  1956  to  1,243  in  1957. 

Landings  continued  to  grow  at  about  50,000 
tons  annually  into  the  early  1960's.  The  mar- 
ket also  began  to  recover  after  the  lows  of 
1955  and  prices  began  a  steady  upward  trend. 
By  1962,  the  average  price  of  yellowfin  at 
Yaezu  had  risen  to  $328.  Small  fortunes  were 
being  made  by  the  end  of  the  decade.  It  be- 
came apparent  that  craft  of  at  least  250  tons 
in  size  were  needed  to  operate  efficiently  from 
Japan  on  the  south  Pacific  and  Indian  Ocean 
grounds  as  well  as  from  bases  on  the  newly 
opened  Atlantic  grounds.  The  value  of  licenses 
for  supplementary  tonnage  increased  rapidly. 
Supplementary  tonnage  could  be  purchased 
for  about  $100  per  ton  in  1955,  rose  to  about 
$500  in  1959,  and  in  1960  approached  $1,000 
per  ton  (Masuda,  1963,  p.  556). 7  In  1960,  ad- 
ditional free  tonnage  was  permitted  for  craft 
of  less  than  240  tons  in  size  if  they  were 
wooden  craft  over  6  years  old  or  steel  craft 
over  12  years.  Also,  restrictions  on  the  use 
of  the  licenses  for  the  less  than  100-ton  vessels 
issued  after  1953  as  supplementary  tonnage, 
licenses  that  previously  could  not  be  used  for 
this  purpose,  were  relaxed.  Another  building 
boom  was  underway  and  the  average  size  of 
the    vessels    in    the    fleet    grew    with    it    (see 

Figure  3). 

i 
Pressure  for  additional  entry  into  the  tuna 

fishery,  never  quiescent,  began  to  rise  marked- 
ly with  the  rise  in  profits  from  the  fishery. 
Pressure  was  especially  strong  after  1956  from 


7  Precise  figures  on  sale  value  of  licenses  are  difficult 
to  obtain  since  profits  from  their  sale  is  subject  to  capital 
gains  tax.  Underreporting  to  avoid  taxes  appears  to  have 
been  the  rule. 


150 


900 

800 

700 

'Number   600 
of 

Craft 

500 

400 

300 

200 

100 


40-100  Tons 


100-200  Tons 


""-  / 


=5^ 


"7 


s 


>^  / 


over  500  Tons 


1951 


i  i  i  i  i — i — i — i — i — i — i — i — i — r 

1955  1960  1965 


Figure  2.  —  Trends  in  numbers  of  licensed  distant  sea  skipjack-tuna  craft  by  size 
category.  Data:  (Japanese  Tuna  Fisheries  Federation,  1969,  p.  6). 


325 


Figure  3.  —  Annual  construction  of  skipjack-tuna  craft  over  fifty  gross 
tons  in  size.  Data  for  1951-52  from  (Masuda,  1963,  p.  542),  for  1963-67 
from  (Japanese  Tuna  Fisheries  Federation,  1969,  p.  9). 


151 


the  salmon  fishery  as  a  result  of  restrictions 
on  that  fishery  growing  out  of  the  USSR-Japan 
agreement  concerning  it.  An  attempt  to 
relieve  this  pressure  was  made  in  June  1957 
by  raising  the  lower  limit  for  licensed  skipjack- 
tuna  vessels  from  20  to  40  tons.  The  result 
was  the  almost  instantaneous  creation  of  a 
39.9-ton  tuna  vessel  fleet.8  A  fairly  large  num- 
ber of  "39-tonners"  were  built  by  owners  in 
the  traditional  salmon  ports  of  northern  Japan 
but  a  majority  of  these  new  "free  entry"  vessels 
appeared  in  the  traditional  skipjack-tuna 
ports.  The  measure  thus  did  provide  some 
relief  for  the  depressed  salmon  and  other  fish- 
eries but  the  main  effect  appeared  to  be  in- 
creased investment  by  those  already  in  the 
skipjack-tuna  fishery.  Pressure  from  the  salmon 
fishermen  continued  and  some  fifty  new 
"medium-sized"  tuna  licenses  were  given 
craft  owners  in  this  fishery  between  1960  and 
1962  in  exchange  for  their  abandonment  of 
the  salmon  fishery. 

A  demand  to  permit  increased  use  of  mother- 
ships  also  began  to  develop  in  the  late  1950's. 
Large  motherships  operating  with  independent 
licensed  tuna  vessels  had  been  authorized 
since  1948.  Fairly  stringent  restrictions  had 
been  placed  on  the  annual  catch  and  on  place 
of  fishing  of  those  "independent  vessel  mother- 
ships"  as  they  came  to  be  called.9  However,  in 
the  late  1950's,  the  larger  tuna  longline 
vessels  began  to  carry  "portable  catcher  boats" 
on  board.  Once  on  the  fishing  ground,  these 
catcher    boats    proved    almost    as    efficient    in 


H  Accurate  records  were  not  kept  on  the  number  of 
such  craft  until  a  centralized  licensing  system  was  estab- 
lished in  1964.  However,  one  study  by  Fisheries  Agency 
personnel  in  which  an  attempt  was  made  to  trace  the 
growth  of  this  fleet  showed  only  three  such  craft  were 
launched  in  1957,  23  in  1958,  117  in  1959,  and  194  in 
1960  (Japanese  Fisheries  Agency,  May  8,  1963,  p.  6). 
No  data  are  available  on  the  number  of  salmon  longline 
craft  under  40  tons  that  switched  to  tuna  longlining  but 
the  number  probably  was  substantial. 

H  Motherships  were  limited  in  place  of  operation  to 
designated  areas  in  the  central  and  southern  parts  of 
the  Pacific  and  always  under  a  catch  quota  system.  The 
maximum  number  of  motherships  used  in  any  one  year 
was  six,  each  with  up  to  50  independently  licensed  tuna 
long-liners.  In  the  early  1950's,  Antarctic  whaling  mother- 
ships  were  used  as  tuna  longline  motherships  in  the 
offseason.  However,  salmon  motherships  came  to  be 
used  with  restrictions  on  that  fishery  imposed  by  the 
Japanese-Soviet  agreement  in  1956.  Each  mothership 
fleet  was  granted  a  maximum  catch  quota  before  leaving 
port.  The  total  quota  for  all  mothership  fleets  reached 
a  high  of  28,000  tons  in  1958. 


terms  of  catch  rates  per  day  as  the  independent 
vessels.  A  new  category  of  licensing  was  estab- 
lished for  these  craft  in  April  1961  and  re- 
vised in  September  1962.  Two  classes  of  these 
"catcher  boat  carrying  motherships,"  as  they 
came  to  be  called,  were  created  —  less  than 
2,000  ton  craft  where  the  mothership  was  per- 
mitted to  fish,  and  over  2,000  ton  craft  where 
the  mothership  was  not  permitted  to  fish.  A 
complex  system  of  computing  licensed  tonnage 
was  established  for  the  catcher  boats.  In  gen- 
eral, it  required  that  regular  licensed  craft 
be  decommissioned  in  considerable  larger  ton- 
nage for  the  catcher  boat  than  the  maximum 
size  of  20  tons  established  for  each  skiff.  Re- 
strictions were  also  placed  on  area  of  operation 
of  these  two  new  classes  of  motherships.  Regu- 
lations as  to  place  of  operation  were  designed 
generally  to  limit  them  to  the  southwestern 
Pacific,  Indian,  and  Atlantic  Oceans. 

The  regulatory  system  had  become  some- 
what outmoded  and  unwieldy  by  the  early 
1960's.  The  basic  fisheries  law  was  inadequate 
for  proper  regulation  of  the  new  motherships 
and  the  need  for  regulation  of  the  new  "39- 
ton"  fleet  was  becoming  apparent.  The  former 
medium-sized  vessels  that  had  been  allowed 
to  expand  to  over  100  tons  but  held  below  200 
tons  in  size,  about  150  in  number,  were  proving 
to  be  uneconomical.  Not  large  enough  to 
operate  effectively  on  grounds  south  of  the 
equator,  they  were  too  large  to  compete  ef- 
fectively with  the  large  number  of  "39-ton" 
"free-entry"  craft  and  less  than  100-ton  li- 
censed craft  on  grounds  adjacent  to  Japan. 
The  price  of  licenses  continued  to  rise  to  a 
peak  of  about  $1,200  per  vessel  ton  in  1962. 
Few  owners  of  these  "in  between"  craft  could 
afford  to  purchase  supplementary  tonnage  for 
craft  enlargement  at  these  prices.  For  these 
and  other  reasons,  the  realization  became 
general  that  a  new  legal  framework  for  ad- 
ministration of  the  fishery  was  needed,  a  con- 
dition that  was  true  of  other  fisheries  as  well. 

A  revision  of  the  basic  fisheries  law  by  the 
National  Diet  in  August  1962  provided  a  new 
framework.  In  reference  to  the  tuna  fleet,  the 
new  law  codified  the  system  for  motherships 
described  above,  rationalized  a  number  of  com- 
plexities that  had  developed  in  the  licensing 
system,  and  lowered  the  age  at  which  a  vessel 
could  be  replaced  to  4  for  wooden  vessels  and 
8  for  steel  vessels.  The  only  aspect  of  the  new 


152 


law  that  specifically  permitted  additional  ton- 
nage to  the  fleet  concerned  the  "in  between" 
craft  between  100  and  180  tons.  These  were 
granted  permission  to  enlarge  to  240  tons, 
about  the  smallest  sized  vessel  that  could 
operate  effectively  south  of  the  equator  from 
Japanese  ports. 

Landings  from  the  longline  fishery  peaked 
in  1962.  Declines  in  catches  from  that  year, 
increased  competition  in  international  markets 
from  the  Taiwanese  and  Korean  fisheries,  and 
sudden  rises  in  labor  costs  greatly  reduced 
pressure  for  further  expansion  of  the  fleet.  The 
"39-ton"  fleet  was  brought  back  into  the 
limited  entry  system  in  1964  with  a  passage 
of  a  law  that  established  a  "near  seas"  skipjack- 
tuna  industry.  The  law  limited  the  number  of 
licenses  for  20-  to  50-ton  craft  engaged  in  the 
skipjack-tuna  fishery  to  1,850  vessels,  a  number 
selected  primarily  because  it  was  sufficiently 
large  to  cover  all  craft  of  this  size  range  al- 
ready in  the  fishery.  In  1964,  1,708  craft  were 
licensed  and  registered  under  this  law  but 
the  number  has  declined  slightly  since. 

Changes  in  the  regulatory  system  since  the 
near  seas  fleet  was  established  have  been 
relatively  few  in  number  compared  to  earlier 
years.  As  longline  catches  declined,  the  pole- 
and-line  live  bait  fishery  received  increased 
attention.  The  more  substantial  changes  in 
regulations  have  been  designed  to  permit  or 
encourage  decommissioning  of  large  vessels 
to  build  smaller  vessels  for  this  fishery.  Strong 
pressure  has  developed  since  the  mid-1960's 
for  reduction  in  the  size  of  the  fleet.  Agreement 
appears  to  be  general  that  this  should  be  done 
but  as  yet  an  acceptable  method  to  do  so  has 
not  been  devised. 

EFFECTS  ON  DIFFERENT  ASPECTS 
OF  THE  FISHERY 

As  can  been  seen  from  the  above  overly 
simplified  description,  measures  for  regula- 
tion of  the  Japanese  skipjack-tuna  fishery  center 
strongly  on  limitation  of  the  size  and  number 
of  craft.  Only  minor  use  has  been  made  of 
catch  quotas  and  restrictions  on  place  of  fish- 
ing, measures  that  tend  to  reduce  the  efficiency 
of  use  of  vessels  and  equipment.  The  fleet  as 
it  developed  is  very  much  a  result  of  regulation 
through  use  of  limited  entry  and  controls  on 
size   of  vessels.    Discussion   will   now   turn   to 


the  major  effects,  some  obvious  and  foreseen, 
some  less  obvious  and  forseen  dimly  if  at  all, 
that  the  regulatory  system  had  on  the  fishery. 

Capital  Acquisition  and  Resource  Allocation 

One  of  the  more  striking  aspects  of  the  fishery 
was  the  rapidity  with  which  the  fleet  was  ex- 
panded after  the  Allied  Occupation  ended. 
Vessels  used  in  the  fishery  are  not  extraordi- 
narily large  as  fishing  vessels  go  nor  were 
construction  costs  in  Japan  high  by  any  stand- 
ard. However,  they  do  represent  a  sizeable 
capital  investment  and  requirements  for  operat- 
ing capital  are  substantial.  Owner-operator 
enterprises  dominated  the  fishery  in  the  early 
days.  This  meant  that  most  were  small  enter- 
prises headed  by  individuals  with  poorly  estab- 
lished lines  to  sources  of  capital.  Two-  and 
three-boat  enterprises  became  common  by  the 
early  1960's  but  the  fishery  continues  to  be 
made  up  largely  of  small  enterprises.  The 
large  fishing  corporations  of  Japan  have  played 
and  continue  to  play  a  relatively  minor  role 
in  the  fishery. 

The  effect  the  system  as  applied  had  on 
acquistion  of  capital  is,  of  course,  obvious.  Li- 
censes from  the  beginning  became,  for  all 
practical  purposes,  the  personal  property  of 
the  recipient.  As  such  they  were  sold,  traded, 
or  used  as  security  for  loans.  Even  at  the  de- 
pressed tuna  prices  of  the  mid-1950's,  license 
values  ranged  from  10%  to  20%  of  construction 
costs  for  a  vessel.  At  1962  earning  levels,  the 
value  of  the  license  almost  equaled  that  of 
the  vessel.  With  security  of  this  nature  to 
offer,  no  license  holder  had  any  difficulty  in 
gaining  loans  for  either  fixed  or  operating 
capital.  Without  the  limited  entry  system  and 
property  characteristics  of  the  licenses,  the 
fishery  possibly  would  have  expanded  more 
slowly,  paradoxical  though  this  may  sound. 
Enlargement  of  craft  also  would  have  been 
more  dificult  had  these  valuable  licenses  not 
been  available  to  use  as  security  for  loans. 
One  could  postulate  that  the  fleet  would  have 
come  to  consist  of  a  much  larger  number  of 
smaller  craft  without  it,  although  larger  craft 
constructed  and  owned  by  large  corporations 
may  have  come  to  dominate  the  fishery. 

Licenses  decreased  in  value  rather  precipi- 
tously after   1962  to  a  low  of  about  $330  in 


153 


1965  (Commercial  Fisheries  Review,  1966, 
p.  73).  Rates  of  indebtedness  at  the  peak  of 
license  values  in  1962  had  been  much  higher 
than  in  other  Japanese  fisheries.  Debts  on  the 
fixed  capital  alone  of  craft  over  200  tons  in 
1962  averaged  72% ,  almost  an  inverse  ratio  to 
the  30%  rate  in  the  East  China  Sea  trawl  fish- 
ery (Masuda,  1963,  p.  539).  Debts  on  smaller 
licensed  vessels  averaged  over  50% .  Improve- 
ment in  the  earning  position  of  tuna  vessels 
in  the  late  1960's  with  the  rapid  increase  in 
price  of  tuna  in  Japan  stabilized  the  economic 
picture  for  most  owners  after  1965.  However, 
many  marginal  enterprises  were  forced  out 
of  the  fishery  during  the  mid-1960's. 

It  can  also  be  argued  that  the  licensing  sys- 
tem as  it  evolved  also  led  to  a  misallocation 
of  resources  within  the  national  economy  as 
a  whole.  From  the  standpoint  of  the  national 
economy,  investment  in  the  tuna  fishery  ob- 
siously  was  profitable  at  least  through  1962. 
However,  the  high,  and  at  times  unrealistic, 
value  of  the  licenses  in  the  tuna  fishery  gave 
this  fishery  an  extremely  favorable  competi- 
tive position  within  financial  institutions 
specializing  in  fisheries,  and,  indeed,  in  the 
national  capital  market  as  a  whole.  The  total 
investment  was  substantial  and,  as  proved 
later,  was  larger  than  needed  to  harvest  the 
resource.  Where  the  investment  level  would 
have  proved  most  advantageous  is  difficult  to 
determine  and  no  effort  to  do  so  is  known  by 
the  author.  Few  would  argue,  however,  that 
a  better  allocation  of  national  resources  would 
not  have  been  obtained  had  part  of  the  invest- 
ment in  the  tuna  fleet  been  directed  to  other 
channels. 

Size  and  Nature  of  the  Fleet 

That  the  size  and  characteristics  of  craft 
in  the  fleet  was  shaped  strongly  by  the  regu- 
latory system  is  apparent  from  the  earlier  dis- 
cussion of  the  development  of  the  system.  En- 
largement of  craft  was  a  basic  and  continuing 
policy  throughout  the  period  of  expansion. 
The  most  effective  measure  used  to  fulfill  this 
policy  was  the  frequent  granting  of  additional 
free  licensed  tonnage  that  could  only  be  used 
with  the  licensed  tonnage  of  the  old  vessel 
which  was  in  turn  decommissioned.  This,  and 
the  practice  of  allowing  only  licensed  tonnage 


from  decommissioned  "medium-sized"  craft 
to  be  used  for  enlargement  under  any  circum- 
stances, hurried  the  disappearance  of  these 
smaller  licensed  craft  as  well  as  the  construc- 
tion of  larger  ones. 

The  measures  used  were  highly  effective  as 
is  shown  by  the  increase  in  average  vessel  size 
from  91  to  230  gross  tons  between  1952  and 
1962.  It  also  meant  that  many  vessels  were 
retired  well  before  their  useful  life  was  ended. 
This  wasteful  aspect  was  recognized  and  an 
attempt  made  to  minimize  it  by  placing  mini- 
mum ages  on  craft  that  could  be  decommis- 
sioned. That  this  time  was  shortened  from  6 
to  4  years  for  wooden  vessels  and  from  12 
to  8  years  for  steel  vessels  illustrates  the  pres- 
sures applied  to  take  advantage  of  grants  of 
tonnage,  grants  which  usually  carried  a  2-year 
maximum  for  use  from  the  date  they  were 
granted.  A  recognized  shortcoming  of  the  sys- 
tem, it  was  nevertheless  one  that  was  never 
solved  satisfactorily  during  the  period  of  ex- 
pansion. 

An  unforeseen  result,  or  certainly  one  that 
was  predicted  poorly,  concerned  adverse  ef- 
fects on  the  structure  of  individual  vessels. 
As  the  fishing  grounds  became  more  distant, 
a  premium  was  placed  on  hold  capacity  for 
fuel  and  fish.  Given  the  absolute  limit  on  gross 
tonnage  permitted  for  an  individual  vessel, 
the  owners  designed  around  this  limit  with 
emphasis  on  increased  carrying  capacity.  First 
started  in  the  late  1950's,  craft  with  20%  to 
30%  greater  carrying  capacity  were  soon  being 
built  with  no  increase  in  computed  tonnage 
(Masuda,  1963,  p.  546).  Crew  quarters  and 
below-deck  working  space  became  more  cramp- 
ed in  the  process  and  safety  equipment  was 
reduced  to  the  minimum  permissible  standards 
and  often  stowed  in  inaccessible  places.  Sea- 
worthiness also  often  suffered  because  of  re- 
arrangement of  storage  space  that  decreased 
stability,  a  factor  that  undoubtedly  contributed 
to  the  loss  at  sea  of  a  number  of  smaller  craft. 
Many  of  these  adverse  aspects  have  been  cor- 
rected subsequently  but  only  through  greater 
expenditure  of  administrative  time  for  inspec- 
tion, additional  tonnage  concessions  that  could 
not  be  used  for  hold  space,  and  a  weakening 
of  the  competitive  position  of  the  fishery  for 
labor  because  of  poor  working  and  living  con- 
ditions while  at  sea. 


154 


Effect  on  Other  Fisheries 

One  could  argue,  as  was  pointed  out  earlier, 
that  the  superior  competitive  position  of  the 
tuna  fishery  possibly  had  some  adverse  effects 
on  other  fisheries,  primarily  in  reference  to 
competition  for  capital.  Comparatively  high 
returns  to  labor  in  the  tuna  fishery  also  gave 
it  a  competitive  position  in  this  respect.  How- 
ever, labor  was  not  a  major  problem  for  any 
fishery  prior  to  the  early  1960's  and  since 
labor  was  generally  drawn  from  families  and 
acquaintances  of  vessel  owners,  the  tuna  fishery 
appears  to  have  had  little  effect  even  on  the 
quality  of  labor  available  to  other  fisheries. 

The  overall  effect  on  other  fisheries,  or  at 
least  the  administration  of  them,  probably 
was  positive.  Since  entry  was  controlled,  re- 
lief could  selectively  be  provided  fisheries 
creating  the  greatest  administrative  problems. 
Certainly  the  Minister  of  Foreign  Affairs  must 
have  been  happy  to  see  pressure  relieved  on 
the  East  China  Sea  and  North  Pacific  Salmon 
fisheries  in  light  of  the  adverse  reaction  of 
mainland  China  and  the  Soviet  Union  to  these 
fisheries.  Had  these  new  licenses  for  the  tuna 
fishery  been  placed  on  open  bid,  one  could 
hardly  have  expected  fishermen  from  depressed 
fisheries  to  compete  for  them  with  any  degree 
of  success. 

Effects  on  other  fisheries  may  be  somewhat 
nebulous  and  difficult  to  define  with  precision, 
but  the  effect  on  the  live  bait  pole-and-line 
fishery  is  much  clearer.  That  the  two  methods, 
or  fisheries  if  one  wishes,  were  administered 
as  a  single  fishery  meant  that  expansion  of 
the  live  bait  fishery  was  neglected  for  over  a 
decade.  Catches  by  the  live  bait  method  did 
not  decline  during  expansion  of  the  longline 
fishery,  in  fact  the  secular  trend  was  up  slight- 
ly (see  Figure  1).  However,  resources  for  this 
fishery  were  underutilized,  a  fact  known  at  the 
time  and  borne  out  by  the  increase  in  landings 
since  the  mid-1960's.  Craft  of  sufficient  size 
to  properly  exploit  this  resource  and  permitted 
to  do  so  were  also  the  only  ones  permitted  to 
fish  with  longlines  for  tuna.  Given  the  higher 
rate  of  return  on  tuna,  the  choice  of  a  vessel 
owner  is  not  difficult  to  see.  That  most  did 
specialize  in  longlining  is  shown  by  the  fact 
that  the  number  of  licensed  craft  using  the 
live   bait   method   declined   from   737   in    1953 


to  231  in  1961;  total  tonnage  of  vessels  so 
used  declined  from  80,000  tons  at  the  peak  to 
33,000  tons  in  1961  (Masuda,  1963,  p.  358  and 
546). 

That  the  total  catch  by  the  live  bait  method 
continued  to  be  stable  throughout  expansion 
of  the  tuna  longlining  can  be  attributed  pri- 
marily to  unlicensed  craft,  including  the  "39- 
tonners"  after  1957.  These  craft  were  sufficient- 
ly large  to  exploit  the  traditional  grounds 
adjacent  to  Japan.  However,  craft  of  over  100 
tons  in  size  are  needed  to  exploit  the  large 
skipjack  resources  in  more  distant  southern 
waters.  By  1960,  nearly  all  craft  of  this  size 
had  been  rebuilt  without  live  bait  wells.  With 
the  decline  in  longline  catches,  a  distant  seas 
live  bait  fishery  developed  fairly  rapidly.  In 
1964,  only  138  craft  over  100  tons  in  size 
used  the  live  bait  method;  by  1967,  the  num- 
ber had  increased  to  224  (Japanese  Tuna  Fish- 
eries Federation,  1969,  p.  13).  Had  craft  using 
the  live  bait  method  been  administered  sep- 
arately, it  can  be  assumed  that  craft  would 
have  been  available  to  develop  these  distant 
grounds  during  the  1950's.  That  this  was  not 
done  can  be  regarded  as  a  loss  to  the  national 
economy  during  the  period. 

Effects  on  Location  of 
Shore-Based  Activities  in  Japan 

The  regional  pattern  of  economic  activities 
connected  with  the  fishery  changed  consider- 
ably during  the  period  of  rapid  expansion. 
Fishing  ports  and  the  fleet  were  distributed 
fairly  evenly  between  the  southern  tip  of  the 
island  of  Kyushu  and  the  northeastern  port 
of  Honshu  when  the  live  bait  method  dominat- 
ed the  fleet's  activities.  Most  of  the  fleet  would 
gather  in  the  south  in  early  spring  to  pick 
up  the  annual  runs  of  skipjack  and  to  a  lesser 
extent,  albacore,  and  follow  them  northward 
along  the  Pacific  Coast  until  they  disappeared 
in  late  summer  off  northeastern  Honshu.  Land- 
ings were  made  at  the  nearest  port,  nearly  all 
of  which  had  a  dried  skipjack  stick  process- 
ing industry,  the  main  use  for  most  of  the 
catch.  Craft  would  then  be  converted  for  tuna 
longlining  on  winter  tuna  grounds  adjacent 
to  Japan.  The  main  market  for  tuna  was  in 
the  Tokyo  region  and  catches  from  the  winter 
fishery  were  landed  at  ports  in  that  area. 


155 


As  tuna  longlining  increased  in  importance 
and  became  a  year  round  activity,  one  could 
easily  have  predicted  that  activities  would 
concentrate  in  a  smaller  number  of  ports. 
Grounds  for  the  year  round  tuna  fishery  were 
so  distant  from  Japan  that  no  port  had  a 
locational  advantage  of  any  significance  in 
reference  to  the  grounds  as  was  the  case  with 
the  live  bait  fishery.  The  main  markets  for 
tuna  were  the  canneries,  export  companies, 
and  the  large  urban  population  in  the  Tokyo 
area.  As  craft  became  larger,  smaller  markets 
were  unable  to  handle  the  full  load  of  most 
vessels  expeditiously,  a  factor  that  further 
favored  concentration.  Concentration  of  eco- 
nomic activities  of  the  longliners  in  a  few  ports 
thus  would  have  been  expected  quite  apart 
from  the  regulatory  system. 

The  regulatory  system  as  applied  did,  how- 
ever, influence  the  regional  pattern  signifi- 
cantly. Among  the  more  readily  apparent  in- 
fluences perhaps  was  that  it  hastened  enlarge- 
ment of  craft  and  thus  increased  tendencies 
toward  concentration  in  the  central  ports. 
Conversely,  in  another  aspect,  it  tended  to 
favor  continued  dispersion  of  economic  activi- 
ties other  than  landing  of  the  fish.  This  de- 
rived from  the  fact  that  ownership  of  the  fleet 
was  dispersed  at  the  time  licenses  were  issued. 
Ties  of  Japanese  fishermen,  both  economic 
and  social,  to  their  home  port  are  strong.  A 
man's  boat  is  his  livelihood  and  sale  of  the 
right  to  use  it  is  restricted  by  strong  pressures 
of  tradition.  That  the  value  of  the  license  in- 
creased steadily  during  most  of  the  period  of 
expansion  meant  that  most  holders,  even  in 
more  remote  areas,  were  able  to  fund  new 
craft  and  expand  along  with  the  fishery.  With- 
out this  source  of  funding,  the  longliners  would 
almost  certainly  have  been  concentrated  in 
all  respects  in  the  centrally  located  ports  where 
capital  was  more  readily  available  and  where 
attention  to  the  fishery  would  have  been  much 
stronger.  However,  having  been  given  the  li- 
censes, owners  in  outlying  ports  generally 
kept  pace  with  the  switch  to  longlining;  with- 
out the  license  as  security,  lack  of  capital  alone 
probably  would  have  been  a  major  deterrent 
to  so  doing.  Landing  and  most  resupplying 
of  vessels  might  be  carried  out  in  centrally 
located  ports  such  as  Yaezu,  Misaki,  or  Tokyo 
but  the  economic  stimulation  from  other  activi- 


ties such  as  management,  labor  recruitment, 
and  expenditures  by  management  and  labor 
largely  accrued  to  the  ports  where  the  owner 
of  the  license  resided.  As  such,  the  fishery  con- 
tinued to  contribute  to  regional  economies  to 
a  larger  extent  than  if  the  regulatory  system 
had  not  existed.  Thus,  the  net  effect  of  the 
regulatory  system  appears  to  have  been  a 
conservative  one  working  against  an  expected 
tendency  toward  concentration  in  the  major 
market  ports. 

Flow  of  Capital  to  Other  Countries 

A  predictable  effect  of  a  limited  entry  system 
in  a  profitable  fishery  such  as  the  tuna  fishery 
in  which  overall  control  of  entry  to  the  fishing 
grounds  is  impossible  would  be  a  flow  of 
capital  to  other  countries.  This  was  recognized 
early  in  the  period  of  expansion  and  fairly 
effective  controls  were  developed  to  control 
it,  at  least  through  1963.  The  method  used 
was  to  restrict  export  of  tuna  longliners.  The 
craft  themselves  are  not  particularly  complex 
nor  is  the  equipment  used  on  them.  However, 
countries  that  had  the  industrial  establish- 
ment to  build  them,  by  and  large  were  not 
able  to  compete  with  the  Japanese  in  the 
fishery  because  of  labor  costs.  Countries  that 
desired  to  enter  the  fishery  and  were  in  a  favor- 
able competitive  position  in  reference  to  labor 
costs  were  not  able  to  build  the  vessels.  Given 
these  conditions,  strict  controls  on  export  of 
longliners  were  used  to  prevent  Japanese  entre- 
preneurs from  transferring  registration  to 
other  countries  and  using  Japanese  or  foreign 
crews  and,  at  the  same  time,  retard  the  de- 
velopment of  the  fishery  by  other  countries. 
Some  transfer  of  registration  was  permitted 
for  operation  by  joint  Japanese  and  foreign 
companies  from  ports  in  the  country  of  the 
latter.  However,  conditions  under  which  this 
could  be  done  were  restricted  severely;  in  a 
1965  survey  by  the  Fisheries  Agency,  only  17 
vessels  were  found  to  be  so  operated  (Com- 
mercial Fisheries  Review,  1966,  p.  85).  Pres- 
sures to  permit  export,  especially  by  shipyard 
owners  in  Japan,  were  great,  but  were  con- 
tained until  1964.  By  this  time,  other  nations, 
especially  Korea,  were  developing  a  capacity 
to  build  longliners  and  the  restrictions  were 
relaxed. 


156 


Japanese  capital  has  played  an  important 
role  in  the  development  of  foreign  fleets  since 
the  early  1960's.  Large  Japanese  trading  com- 
panies handle  most  of  the  tuna  exported  from 
overseas  bases,  bases  originally  established 
to  serve  Japanese  vessels.  As  other  countries, 
namely  Taiwan  and  Korea,  began  to  develop 
fleets,  they  also  used  these  bases  and  sold 
their  catches  to  the  Japanese  companies.  In 
return,  vessels  from  these  countries  have  re- 
ceived financial  assistance,  largely  operating 
capital,  from  these  large  companies.  A  new 
base  opened  recently  by  a  large  Japanese  com- 
pany in  Mombasa,  Kenya  reportedly  is  to  be 
used  almost  entirely  by  Taiwanese  vessels 
(U.S.  Bureau  of  Commercial  Fisheries,  Febru- 
ary 24,  1969).  However,  this  Japanese  invest- 
ment must  be  attributed  primarily  to  the  higher 
labor  costs  of  Japanese  vessels  not  to  restric- 
tions on  their  number.  Under  conditions  in 
the  Japanese  fishery  since  the  mid-1960's,  it 
is  doubtful  that  any  significant  increase  of 
Japanese  vessels  operating  from  these  bases 
could  be  expected  even  if  the  fishery  were 
opened  to  unlimited  entry. 


CONCLUSION 

In  retrospect,  no  one  in  Japan  or  elsewhere 
would  consider  the  regulatory  system  develop- 
ed for  the  Japanese  skipjack-tuna  fishery  to 
be  a  complete  success.  However,  few  would 
argue  that  the  fishery  and  the  country  were 
not  served  better  by  limitation  of  entry  than 
they  would  have  been  had  no  controls  been 
imposed  on  the  number  of  craft.  The  system 
did  have  a  goodly  measure  of  success  in  refer- 
ence to  its  main  goal,  that  is,  to  maintain  a 
high  level  of  economic  viability  of  enter- 
prises in  the  fishery.  Without  it,  a  gross 
over-investment  in  small  vessels  is  almost 
certain  to  have  taken  place  in  the  early  1950's. 
Depression  of  the  market,  strained  financial 
condition  of  enterprises,  and  a  loss  of  all 
economic  rent  from  the  fishery  likely  would 
have  occurred  long  before  the  resource  ap- 
proached full  exploitation.  Conflicts  on  the 
fishing  grounds,  international  incidents,  and 
disasters  at  sea  also  would  have  been  more 
numerous.  Thus,  a  second  major  goal,  harmony 
within   the  fleet  and   on  the  fishing  grounds, 


was  at  least  partially  achieved.  If  the  system 
has  been  less  successful  since  the  early  1960's, 
the  fault  can  hardly  be  laid  at  the  feet  of  the 
fishery  policy  makers  and  administrators. 
Their  control  over  entry  of  fishermen  of  other 
countries  ended  with  Japanese  ability  to  con- 
trol the  technology  of  the  fishery.  Had  fisher- 
men from  other  nations  had  the  wherewithal 
to  enter  the  fishery  from  1950,  acceptance 
of  the  system  by  the  Japanese  fishermen 
would  have  been  far  more  difficult  to  attain. 

Mistakes  were  made,  many  of  them  avoid- 
able. Perhaps  the  largest  was  to  raise  the 
minimum  size  of  licensed  vessels  to  40  tons. 
That  it  was  done  appears  to  have  resulted 
from  an  inadequate  assessment  of  technolog- 
ical developments.  Less  than  40-ton  craft  in 
existence  at  the  time  were  patently  too  small 
to  operate  on  distant  grounds  but  could  re- 
lieve the  need  for  more  vessels  to  exploit  the 
annual  runs  of  skipjack  and  albacore  on  near 
seas  grounds.  Vessels  of  19.99  tons  could 
never  be  designed  for  effective  operation  on 
distant  grounds.  However,  redesign  of  vessels 
of  39.99  tons  led  to  craft  with  the  fishing  power 
of  a  70-ton  vessel  designed  by  standards  used 
in  the  mid-1950's.  At  the  catch  rates  and 
prices  of  tuna  in  the  late  1950's,  these  vessels 
could  operate  profitably  on  distant  grounds 
although  the  large  number  of  disasters  sug- 
gest they  should  not  have  attempted  to  do  so. 
The  problem  of  safety  was  corrected  only  by 
granting  permission  to  increase  size  of  these 
vessels  to  50  tons  with  the  provision  that  the 
additional  tonnage  would  be  used  only  to  in- 
crease crew  comfort  and  safety  and  limiting 
their  use  to  waters  adjacent  to  Japan.  How- 
ever, the  number  of  such  vessels  far  exceeds 
needs  and  the  problem  of  overcapitalization 
has  been  far  more  intractable. 

Some  lawmakers  and  administrators  were 
troubled  also  by  the  tremendous  value  that 
the  licenses  came  to  have  at  no  cost  to  the 
holders  of  the  licenses.  Had  the  tremendous 
expansion  of  the  fishery  and  its  profitable- 
ness been  foreseen  at  the  time  the  fishery  was 
brought  under  regulation,  some  means  pos- 
sibly could  have  been  devised  to  siphon  off 
at  least  part  of  the  economic  rent  represented 
by  the  licenses  into  the  public  coffers.  How- 
ever, to  have  worked  out  an  acceptable  scheme 
for  the  fishery  after  the  basic  system  was  al- 


157 


ready  operating  would  have  been  extremely 
difficult.  Certainly  it  would  have  added  com- 
plexities to  an  already  overly  complex  struc- 
ture that  possibly  would  have  caused  the 
entire  system  to  break  down.  Also,  a  national 
law  that  singled  out  one  fishery  for  such  treat- 
ment probably  would  not  be  acceptable  to  the 
lawmaking  body.  Values  of  licenses  in  more 
stable  Japanese  fisheries  have  never  reached 
levels  considered  to  be  a  problem;  to  impose 
controls  on  these  fisheries  would  create  more 
administrative  problems  than  could  possibly 
be  justified  by  gains  resulting  from  the  controls. 
In  short,  to  have  solved  this  problem,  if  it  was 
one,  in  the  political  arena  of  Japan  or  any  other 
country  with  representative  government  would 
have  been  extremely  difficult.  Possibly  ignoring 
it  was  the  wiser  route  to  follow. 

The  problem  of  overcapitalization  of  the 
world  tuna  fleets  appears  to  be  approaching 
rapidly  if  it  has  not  already  been  reached. 
The  Japanese  were  able  to  limit  entry  to  the 
fishery  and  maintain  economic  viability  of 
enterprises  in  it  during  the  period  that  they 
controlled  longline  technology.  Beyond  ques- 
tion, limited  entry  could  also  be  used  to 
control  excessive  fishing  power  and  the  ex- 
cessive pressure  on  world  tuna  stocks  that 
it  is  certain  to  bring.  The  Japanese  experience 
illustrates  many  of  the  problems  that  would 
attend    the    far    more    complicated    problems 


foreseeable  in  establishment  of  an  international 
system.  It  also  suggests  the  benefits,  in  refer- 
ence to  stock  management  as  well  as  eco- 
nomic viability  of  the  fishing  enterprise, 
could  be  well  worth  the  effort  required  to 
establish  the  system. 


LITERATURE  CITED 

Commercial  Fisheries  Review,  January  1966.  Vol.  28, 
No.  1,  p.  85. 

Commercial   Fisheries  Review,  July   1966.  Vol.   28,   No. 

7,  p.  73. 

Japanese  Fisheries  Agency.  May  8,  1963.  Katsuo-Maguro 
Gyogo,  No  -  40-ton  Munar  Gyosen  ni  Kansuro  Shirgo. 
p.  6. 

Japanese  Tuna  Fisheries  Federation.  1968.  Statistics 
of  the  Japanese  Tuna  Fishery. 

.     1969.    Statistics    of 


the  Japanese  Tuna  Fishery. 

MASUDA,    SHOICHI,   ed.    1963.   Katsuo-Maguro   Soran 
(Skipjack  tuna  Overview),  Tokyo:  Suisanska,  p.  758. 

Yaezu  Fishery  Cooperative.   1963.  Mizuage-daka  Tokei, 
no.  11. 

U.S.    Bureau    of    Commercial    Fisheries.    February    24, 
1969.  Foreign  Fishery  Information  Release  69-7. 


158 


A  Study  of  the  Socioeconomic  Impact  of  Changes 

in  the  Harvesting  Labor  Force  in  the 

Maine  Lobster  Industry1 


A.  M.  Huq2 


ABSTRACT 

The  basic  question  of  the  mobility  of  the  labor  force  in  the  Maine  lobster  fishery 
is  investigated  with  particular  emphasis  on  the  productivity  of  control  groups  within  a 
sample  and  their  social,  educational,  economic,  and  demographic  characteristics.  Under 
various  assumptions  which  would  lead  to  exit  from  the  fishery  of  these  groups  certain 
consequences  are  enumerated,  both  with  regard  to  those  leaving  and  those  remaining 
as  well  as  the  impact  on  and  role  of  the  local  communities  involved.  A  preliminary 
assessment  of  the  impact  of  certain  types  of  management  programs  upon  the  labor 
component  of  the  harvesting  sector  is  presented. 


INTRODUCTION 

In  any  discussion  of  alternative  manage- 
ment strategies  (e.g.,  limited  entry)  that  might 
affect  the  labor  force  in  the  lobster  fishery 
in  Maine,  it  is  important  to  examine  the  socio- 
economic repercussions  of  the  contemplated 
change.  In  some  circumstances  this  may  in- 
volve the  dislocation  of  labor.  In  this  case 
one  must,  for  example,  investigate  whether 
alternative  employment  would  be  available 
to  those  fishermen  who  will  be  excluded  be- 
cause of  limited  entry;  their  employability 
(and  trainability)  relative  to  the  local  labor 
market,  their  geographical  and  occupational 
mobility  patterns,  the  adaptability  of  their 
skills,  alternative  income  earning  possibili- 
ties ("salvage  value"  of  displaced  labor),  the 
potential  for  upgrading  their  existing  skills 
and  for  the  acquisition  of  new  skills,  the 
barriers  to  their  mobility  including  sociolog- 
ical, psychological,  and  economic  variables 
are  some  of  the  crucial  elements  to  be  care- 
fully considered. 

Furthermore,  the  policy  maker  has  to  evalu- 


1  This  paper  is  based  upon  a  study  sponsored  by 
the  National  Marine  Fisheries  Service.  In  addition  to 
the  author,  the  research  team  consisted  of  Harland 
I.  Hasey  and  Anita  Wihry,  Research  Associates. 

2  Director,  Manpower  Research  Project,  University 
of  Maine,  Orono,  Maine. 


ate  the  potential  impact  on  the  local  and 
regional  economy  in  terms  of  shifts  in  income 
and  employment  and  associated  fiscal  conse- 
quences including  welfare  expenditures  and 
changes  in  tax  revenue.  Finally,  it  would  be 
important  to  examine  how  limited  entry  in 
a  given  fishery  such  as  the  lobster  fishery 
might  affect  other  fisheries  such  as  shrimp 
and  scallop  fisheries.  In  a  comprehensive  study, 
all  these  questions  need  to  be  investigated 
before  any  definitive  conclusions  can  be  reach- 
ed. However,  the  present  study  is  of  much 
more  limited  scope  and  pertains  to  only  some 
of  these  questions  bearing  on  limited  entry. 

This  study  focuses  on  the  possible  socio- 
economic impact  of  hypothetical  reduction  in 
the  harvesting  labor  force  in  the  Maine  lobster 
fishery.  As  to  how  this  reduction  is  or  can  be 
brought  about  is  outside  the  scope  of  the 
study.  The  study  utilizes  the  data  obtained 
from  a  sample  survey  of  131  fishermen  from 
three  selected  communities.  The  problem  posed 
for  investigation  was  simply  this:  if  a  group 
of  fishermen  from  this  sample  is  excluded 
from  lobster  fishing  based  on  some  specified 
criterion,  what  sort  of  socioeconomic  impact 
can  be  expected:  Can  certain  indicators  be 
developed  to  measure  such  impact  in  order  to 
consider  alternative  management  strategies? 
For  this  purpose,  it  was  considered  desirable 
to  (a)  introduce  the  notion  of  a  target  group 
composed  of  fishermen  regarded  as  candidates 
for  limited  entry  and  (b)  to  develop  alternative 


159 


criteria  for  the  construction  of  a  set  of  target 
groups  rather  than  singling  out  one  specific 
target  group. 

Constrained  by  time  and  resources  avail- 
able for  this  project,  the  study  addressed  it- 
self only  to  selected  dimensions  of  socioeco- 
nomic impacts  of  limited  entry  into  the  Maine 
lobster  fishery.  It  is  to  be  clearly  understood 
that  some  of  the  findings  of  this  study,  be- 
cause of  its  very  limited  scope,  are  essentially 
for  illustrative  purposes  rather  than  for  use 
as  supportive  materials  for  or  against  any 
implicit  management  strategy  that  may  be 
suggested  by  the  format  of  the  target  groups. 

OBJECTIVES 

The  major  objective  of  the  study  is  to  present 
an  evaluation  of  the  socioeconomic  impacts 
of  limited  entry  into  the  Maine  lobster  fishery. 
A  complete  evaluation  may  include  but  not 
be  limited  to  the  income  and  employment 
effect  on  the  displaced  fishermen,  income 
effect  on  the  surviving  fishermen,  income  and 
fiscal  effect  on  the  local  and  regional  economy, 
effect  on  other  fisheries  and  so  on.  However, 
for  reasons  stated  above,  the  limited  objectives 
of  this  study  are: 

1.  To  make  an  appraisal  of  the  employ  ability 
and  alternative  income  earning  possibilities 
of  displaced  labor. 

2.  To  derive  some  measures  of  social  impact 
in  terms  of  (a)  income  effects  and  (b)  income 
maintenance  burden  associated  with  dis- 
placement because  of  limited  entry. 

RESEARCH  DESIGN 

The  study  was  designed  as  a  small-scale 
pilot  effort,  concentrating  on  three  typical 
communities  rather  than  encompassing  the 
entire  Maine  lobster  fishery.  These  communi- 
ties are  Phippsburg,  Beals,  and  Corea.  The 
selection  was  made  in  consultation  with  the 
Maine  Department  of  Sea  and  Shore  Fisheries 
and  the  National  Marine  Fisheries  Service. 
The  existence  of  some  contrasts  in  the  struc- 
ture of  the  local  economy  and  the  relative 
importance  of  the  lobster  fishery  in  their  econ- 
omy weighed  heavily  in  the  selection  process. 
Corea  represents  a  highly  specialized,  isolated 


economy  where  lobstering  is  the  predominant 
economic  activity.  Beals  is  also  highly  special- 
ized but  less  isolated  than  Corea.  Phippsburg's 
economy  is  more  diversified  and  in  close  prox- 
imity to  sources  of  alternative  job  opportuni- 
ties. Each  of  the  areas  has  one  feature  in 
common:  the  lobster  fishery  is  a  major  eco- 
nomic activity. 

It  is  difficult  to  say  how  representative  these 
three  communities  are  of  the  entire  lobster 
fishery.  Sufficient  information  is  not  readily 
available  to  identify  the  economic  character- 
istics of  the  population  of  lobster  fishermen 
in  Maine  and  relate  them  to  those  of  the 
sample  fishermen  in  these  communities. 

For  the  purpose  of  the  study  the  following 
hypotheses  were  formulated  for  investigation: 

1.  Limited  entry  could  eventually  exclude 
a  certain  fraction  of  the  lobster  harvesting 
labor  force  that  will  be  otherwise  unemploy- 
able. (Alternative  hypothesis:  a  significant 
fraction  of  labor  displaced  because  of  limited 
entry  will  be  employable,  given  the  conditions 
in  the  local  labor  market,  the  type  of  skill 
possessed,  the  potential  for  adapting  skills 
to  job  market  requirements,  the  availability 
of  retraining  opportunities,  motivation  for 
training,  and  mobility  and  so  on). 

2.  Displacement  of  labor  because  of  limited 
entry  may  adversely  affect  the  local  economy 
because  of  loss  of  income  from  lobstering  not 
being  compensated  for  by  income  from  alterna- 
tive jobs  and  from  additional  lobstering  by 
surviving  fishermen,  and  because  of  loss  of 
income  from  lobstering  on  the  part  of  those 
who  are  not  in  the  labor  force. 

To  generate  the  information  needed  for  this 
investigation,  a  stratified  random  sample  of 
131  fishermen  was  selected.  The  size  of  the 
sample  depended  essentially  on  the  estimated 
cost  per  interview  and  the  budgetary  con- 
straint. The  allocation  to  each  stratum  was 
strictly  according  to  proportion  of  fishermen 
in  each  community  to  the  total  number  of 
fishermen  of  all  three  communities.  The  survey 
data  were  supplemented  by  information  on 
the  local  labor  market  obtained  through  the 
cooperation  of  the  regional  offices  of  the 
Maine  Employment  Security  Commission. 

For  the  survey,  a  structured  questionnaire 
was  developed  and  pretested.  Using  the  modi- 
fied   questionnaire    and    personal     interviews, 


160 


the    survey    was    completed    in    6    weeks.    The 
response  rate  was  better  than  90% . 

The  survey  resulted  in  a  large  volume  of 
information  on  the  sampled  fishermen.  The 
following  broad  categories  of  information 
may  be  identified: 

Categories         Types  of  Information 

Demographic    Age 

Family  Size  and  Composition 

Mobility 

Marital  status 

Socioeconomic  Income 

Employment  history 
Education  and  training 
Monetary  return 
Parental  occupation 
Housing 

Operational       Gear  types 

Investment  in  boat  and  gear 
Operating  expenses 
Maintenance  and  repair  ex- 
penditures 
Size  of  operations 
Seasonal  patterns 
Rate  of  capacity  utilization 

Behavioral-       Reasons  for  lobstering 
Attitudinal   Job  interests 

Attitudes  towards  leaving  the 

lobster  industry 
Job-seeking 

Attitudes  toward  training,  views 
on  excess  capacity 

ANALYSIS 
The  Maine  Lobster  Fishery:  Some  Basic  Facts 

The  lobster  industry  in  the  State  of  Maine 
landed  19.8  million  pounds  of  lobsters  worth 
$16.1  million  in  1969.  This  accounted  for 
10.4%  of  the  quantity  and  58.3%  of  the  value 
of  the  total  fish  and  shellfish  landings  for 
that  year  (Maine  Landings,  1968-70,  p.  3). 

There  were  5,750  lobster  licenses  issued  in 
the  State  in  1969.  These  5,750  lobstermen 
fished  a  total  of  805,375  traps  or  approximately 


105.7  million  trap-days  during  the  year  1969. 
The  gross  earnings  per  unit  of  effort  was 
$0.18  per  trap-day.  This  value  is  arrived  at 
by  adjusting  Maine  landings  up  by  16%  to 
include  landings  not  reported.  This  produced 
total  landings  of  18.7  million  which  were 
divided  by  total  trap-days  yielding  the  re- 
turn of  $0.18  per  trap-day.  The  average  gross 
income  was  approximately  $3,000.  The  total 
investment  in  gear  (i.e.,  boats,  traps,  buoys, 
etc.)  is  about  $10  million.3 

There  have  been  fluctuations  in  the  number 
of  licenses  issued  over  the  past  10  years.  Table 
1  illustrates  a  seemingly  cyclical  pattern  of 
lobster  licenses,  showing  a  high  of  6,472  in 
1961,  a  low  of  5,425  in  1962,  and  another 
high  of  6,316  in  1970. 

The  communities  chosen  for  study  —  Phipps- 
burg,  Corea,  and  Beals  —  represent  277  fisher- 
men or  4.4%  of  the  6,316  fishermen  licensed 
in  1970.  A  sample  of  131  of  the  fishermen  was 
randomly  selected  by  community  as  shown 
in  Table  2.  The  geographical  locations  of  these 
three  communities  are  shown  in  Figure  1. 

Economic  Profile  of  the  Sample  Communities 

Beals  is  an  island  community  of  658  persons 
located  across  Mossabec  Reach  from  Jones- 
port,  Maine,  population  1,337  (1970  Census  — 
Preliminary  Report,  Population  Counts  for 
States).  The  two  communities  —  Beals  and 
Jonesport  —  are  integrated  as  a  labor  market 
but  have  separate  political  identities.  The  only 
administrative  connection  between  the  towns 
is  a  shared  high  school. 

Employment  opportunities  are  limited  to 
the  fishing  industry  and  service  industry  oc- 
cupations. The  Department  of  Sea  and  Shore 
Fisheries  issued  142  lobster  licenses  to  the 
residents  of  Beals  in  1969.  Other  licenses  in- 
clude worms  —  52,  and  clams  —  89.  Many 
of  the  fishermen  hold  more  than  one  license. 
No  license  is  needed  for  shrimping. 

Businesses  on  Beals  include  seven  lobster 
pounds,  most  of  which  are  family  owned  and 
operated.  The  pounds  are  used  to  store  lob- 
sters   until    market    prices    increase    and    the 


3  Information    supplied     by     Robert    Dow,     Research 
Division,  Maine  Department  of  Sea  and  Shore  Fisheries. 


161 


Calais 


Ellsworth  _         ^""Machiasport 
teals 
Gouldsboro 
(Corea) 


Figure  1.  —  Maine  —  selected  geographic  locations. 


Table   1.  —  Number  of  lobster  licenses  issued  in  Maine 
1961-1970. 


Table    2.    —    Distribution   of   the   sample   fishermen   by 
Communities. 


Year 


1961 
1962 
1963 
1964 
1965 


Number  of  licenses  Year 


6,472 
5,658 
5,695 
5,803 

5,802 


1966 
1967 
1968 
1969 

1970 


Number  of  licenses 


5,613 

5,4  25 
5,489 
5,750 
6,316 


Source:    Maine  Department  of  Sea  and  Shore  Fisheries. 


Communities 


Beals 
Corea 

Phippsburg 


TOTAL 


Total  fishermen 


137 
73 
67 


277 


Sample 


61 
27 
44 


131 


162 


pound  may  be  filled  by  the  family  owning  it 
or  the  pound  operator  may  become  a  dealer 
for  part  of  the  year,  buying  from  fishermen 
until  he  has  the  pound  stocked.  A  third  use 
of  the  pound-  is  leasing  to  a  full-time  dealer 
for  his  own  stocking  activities.  If  the  family 
does  not  operate  the  pound  on  a  part-time 
basis,  the  employment  provided  rarely  ex- 
ceeds one  job.  The  two  full-time  lobster  dealers 
on  Beals  employ  between  two  and  four  labor- 
ers each.  The  12  boatyards  are  father  and 
son  operations  although  occasionally  one  non- 
family  employee  may  be  hired.  The  two  clam 
shops  on  the  island  employ  a  total  of  between 
25  and  30  persons  together  —  mainly  women 
who  shuck  clams  for  shipment  outside  the 
area.  The  service  industry  employment  avail- 
able on  Beals  consists  of  jobs  in  three  general 
stores,  one  garage,  one  oil  company,  one 
television  and  radio  sales,  the  local  elementary 
school,  and  various  part-time  jobs  available 
in  the  town  government  (mostly  elective  posi- 
tions) (Table  3). 


Table  3.  —  Occupational  distribution  of  the  work  force 
in  Beals,  1960. 


Male 

Female 

Total 

Professional 

8 

8 

16 

Clerical 

15 

4 

19 

Craftsmen 

28 

28 

Operatives 

17 

17 

Service 

4 

4 

Laborers  (farm) 

11 

1  1 

Laborers 

77 

77 

Total 

156 

16 

172 

Source:  1960  Census  Special  Report  for  Maine  Employment 
Security  Commission.  Approximately  90%  of  the 
"laborers"  may  be  classified  as  lobster  fishermen. 


In  Jonesport  employment  opportunities  are 
in  much  the  same  industries  as  they  are  in 
Beals.  Ninety-nine  lobster  licenses,  60  worm 
licenses,  and  81  clam  licenses  were  issued  by 
the  Department  of  Sea  and  Shore  Fisheries. 
Employment  opportunities  available  in  Jones- 
port  include  jobs  in  one  restaurant,  one  bank, 
one  sardine  factory,  two  grocery  stores,  one 
clothing  store,  one  drug  store,  four  gas  stations, 
three  gas  or  oil  companies  (total  employment 
each    is    no    more    than    three),    one    dentist's 


office,  one  doctor's  office,  two  lobster  dealers 
and  a  lobster  cooperative  which  has  four  em- 
ployees. Other  firms  in  the  area  providing  sub- 
stantial employment  are  two  sardine  factories 
—  one  in  Milbridge  and  one  in  Machiasport. 
This  employment  is  part-time  and  seasonal. 

The  1969  value  of  product  given  by  the 
Census  of  Maine  Manufacturers  for  Beals  is 
$283,258,  the  total  gross  wages  are  $70,856, 
and  average  gross  $2,443.  These  figures  are 
for  manufactured  products  only  and  do  not 
include  income  from  lobstering,  shrimping, 
or  other  fishing  unless  the  catch  has  been 
processed  in  some  manner.  Total  employment 
in  these  industries  is  given  as  29.  For  Jones- 
port  the  corresponding  figures  are  value  of 
product  —  $681,509,  gross  wages  —  $192,495, 
and  average  gross  wage  —  $2,406.  Total  em- 
ployment was  80. 

Total  assessed  value  of  property  on  Beals 
in  1969  was  $237,560.  The  town  budget  shows 
total  receipts  of  $99,376,  and  total  expendi- 
tures of  $73,910,  of  which  about  $55,000  was 
for  wages  distributed  to  inhabitants  of  the 
town. 

Table  4.  —  Occupational  distribution  of  the  work  force 
in  Gouldsboro,  1960. 


Male 

Female 

Total 

Professional 

4 

4 

Managers 

21 

14 

35 

Clerical 

4 

4 

Sales 

8 

9 

17 

Craftsmen 

50 

50 

Operatives 

9 

17 

26 

Private  household 

8 

x 

L6 

Service 

5 

5 

Laborers 

137 

137 

No  information 

33 

9 

42 

Total 

275 

61 

336 

Source:  1960  Census  Special  Report  for  Maine  Employment 
Security  Commission.  Approximately  90%  of  the 
"laborers"  may  be  classified  as  lobster  fishermen. 


Corea  (Gouldsboro):  The  community  in 
Corea  is  part  of  the  township  of  Gouldsboro. 
The  1970  population  of  Gouldsboro  is  1,270, 
an  increase  of  170  people  over  the  1960  figure 
of  1,100.  In  1960  there  were  363  households. 
There  were  420  males  over  14  years  of  age 
and  406  females. 


163 


Corea's  major  industry  is  lobster  fishing, 
providing  some  70-80  jobs.  Other  types  of 
fishing,  which  are  part-time  or  supplemental, 
include  seining,  clamming,  and  worming. 
There  are  some  nine  stores,  a  boatyard  which 
employs  six-seven  people  year  around,  fish 
cannery,  a  naval  tracking  base,  and  eight 
teachers  employed  by  the  town's  elementary 
school.  These  activities  employ  109  full-time 
and  part-time  workers. 


Table  5. 


Occupational  distribution  of  the  work  force 
in  Phippsburg,  1960. 


Male 


Fema 


Total 


Professional 

8 

4 

12 

Farmers  and  farm  managers 

4 

4 

Managers 

16 

1  1 

27 

Clerical 

4 

20 

24 

Crafts 

68 

68 

Operatives 

60 

73 

Private  household 

20 

Services 

12 

12 

Farm  labor 

12 

12 

Laborers 

71 

71 

Others 

27 

8 

35 

Total 

282 

70 

358 

Source:  1960  Census  Special  Report  for  Maine  Employment 
Security  Commission.  Approximately  80%  of  the 
"laborers"  may  be  classified  as  lobster  fishermen. 


Phippsburg:  In  1970  the  population  of 
Phippsburg  was  1,180,  an  increase  of  59  people. 
Of  the  1,121  people  listed  in  April  of  1960, 
397  were  in  the  labor  force;  358  were  employed, 
and  39  were  unemployed.  Of  those  over  14 
years  of  age,  394  were  men  and  403  were 
women.  There  were  335  households. 

Phippsburg's  major  industry  is  the  summer 
tourist  and  summer  resident  trade.  At  Phipps- 
burg there  are  several  large  tenting  grounds, 
a  state  park,  and  many  summer  residences 
located  on  its  several  miles  of  ocean  frontage. 
Other  local  industries  include  fishing,  which 
consists  of  a  fish  factory,  several  large  offshore 
fishing  boats,  and  a  fleet  of  lobster  boats. 
There  are  also  two  small  construction  com- 
panies that  build  and  repair  summer  homes. 
The  bulk  of  Phippsburg's  employed  popula- 
tion, however,  commute  to  other  towns  and 
cities  for  employment.  Probably  the  largest 
employer  of  Phippsburg  people  is  Bath  In- 
dustries located  in  the  adjacent  city  of  Bath. 


Selected  Socioeconomic  Characteristics 
of  the  Sample  Lobstermen 

Average  age  of  the  lobstermen  in  the  sample 
is  42.6  years.  There  are  15  below  the  age  of 
19  and  18  in  the  age  bracket  of  65  and  over. 
The  median  income  for  the  group  is  $5,280 
and  average  income  in  $6,213.  There  are  13 
fishermen  with  income  less  than  $1,000  and 
15  with  income  over  $14,000.  Of  the  118  fisher- 
men who  gave  reasons  for  lobstering,  33 
(which  includes  3  students)  responses  may 
be  categorized  as  "economic"  and  the  rest 
"non-economic"  including  home  consumption, 
preference  for  the  particular  way  of  life,  in- 
fluence of  family,  and  so  on. 

Of  the  109  fishermen  who  supplied  informa- 
tion on  number  of  traps,  slightly  over  50% 
owned  less  than  300  traps;  23  fishermen  owned 
more  than  500  traps.  Of  the  93  fishermen  who 
gave  information  on  investment  in  trap  gear, 
approximately  50%  had  investment  of  less  than 
$2,000;  only  3  had  investment  of  $8,000  and 
over.  The  average  years  of  education  was  9.8. 
Approximately  40%  had  less  than  9  years  of 
education.  Of  131  fishermen,  41  indicated  that 
they  received  some  type  of  formal  vocational 
training  in  areas  including  carpentry,  metal 
working,  mechanic,  professional  and  clerical 
work.  Of  81  fishermen  asked  about  preference 
for  receiving  vocational  training,  63  indicated 
no  preference.  Only  a  small  fraction  express- 
ed preference  for  training  in  electrical,  pro- 
fessional, and  carpentry  work. 

Among  the  109  fishermen  who  supplied  in- 
formation on  income  from  part-time  jobs,  77 
indicated  that  they  had  little  or  no  income 
from  this  source.  Only  7  indicated  that  they 
received  more  than  50%  of  their  income  from 
alternative  jobs.4 

Analysis  of  Target  Groups 

In  order  to  analyze  the  potential  socioeco- 
nomic impact  of  limited  entry,  it  is  necessary 
to  identify  the  possible  candidates  who  might 
be  considered  targets  for  limited  entry  or  any 


4  More  detailed  information  on  these  and  other  aspects 
of  the  study  may  be  found  in  the  complete  final  project 
report,  available  from  the  Economic  Research  Labora- 
tory, National  Marine  Fisheries  Service. 


164 


other  management  strategy  that  might  affect 
the  harvesting  labor  force. 

For  the  purpose  of  this  study,  four  groups 

have  been  constructed,  using  alternative 
criteria.  It  is  not  intended  that  the  groups  be 
mutually  exclusive. 

The  variables  chosen  for  this  analysis  in- 
clude the  following:  income,  investment,  effort, 
and  earnings/effort  ratio.5  It  should  be  noted 
that  with  the  exception  of  one  target  group, 
combinations  of  variables  were  used  to  define 
the  target  groups.  Admittedly,  similar  groups 
could  be  constructed  using  different  criteria. 
Groups  selected  appeared  to  be  quite  meaning- 
ful for  the  purpose  of  this  study. 

Target  Group  I  was  chosen  on  the  basis  of 
a  combination  of  two  criteria:  (a)  low  earn- 
ings/effort ratio,  and  (b)  low  number  of  trap- 
days  serving  as  a  proxy  for  low  income.  It 
was  arbitrarily  decided  that  to  be  eligible 
for  this  group  a  fisherman  had  to  have  an 
income/effort  ratio  of  less  than  0.3  and  had 
to  fish  less  than  30,000  trap-days  per  year. 
Those  fishing  over  30,000  traps  were  not  in- 
cluded because  they  earned  sufficient  income 
for  subsistence.  Table  6  was  especially  con- 
structed for  this  purpose. 

Forty  fishermen  met  the  conditions  set  for 
this  group.  As  it  turned  out,  this  group  had 
an  average  earnings/effort  ratio  of  0.182 
compared  to  0.230  for  the  entire  sample  and 
they  fished  an  average  number  of  12,570  trap- 


5  The  earning/effort  ratio  was  calculated  by  dividing 
the  number  of  trap-days  into  gross  income  reported 
by  the  sample  fishermen. 


days  compared  to  30,707  trap-days  for  the 
sample  as  a  whole.  Their  average  income  was 
only  $2,061  compared  to  an  average  income 
of  $6,213  for  the  sample  as  a  whole.  The 
fishermen  in  this  group  fish  fewer  number  of 
days  and  have  invested  small  amounts  of 
capital  in  gear  and  boat. 

In  any  discussion  of  deliberate  or  planned 
changes  in  the  harvesting  labor  force  in  the 
lobster  fishery,  this  group  with  a  low  earnings/ 
effort  relationship  and  low  absolute  level  of 
income  would  warrant  consideration.  Pre- 
sumably, the  economic  status  of  the  remain- 
ing fishermen  would  improve  the  terms  of  a 
higher  ratio  of  income  to  effort  and  higher 
absolute  level  of  income,  if  this  group  is  elimi- 
nated. Of  course,  one  has  to  look  at  the  social 
cost  of  such  a  change  and  the  political  feasi- 
bility of  such  a  change.  Some  measures  of 
social  cost  are  developed  later  in  this  paper. 

An  alternative  approach  to  the  problem 
would  be  to  consider  only  low  levels  of  pro- 
ductivity as  measured  by  the  low  income/ 
effort  ratio,  regardless  of  the  absolute  size 
of  income.  Here  one  could  argue  that  shifting 
away  from  lobstering  in  this  case  may  be 
socially  gainful,  given  possibilities  for  im- 
proving the  income/effort  ratio  in  alternative 
employments.  From  such  a  reallocation  of 
effort  as  an  economic  resource,  both  the  dis- 
placed fishermen  as  well  as  the  surviving 
fishermen  might  benefit,  as  the  marginal  pro- 
ductivity of  both  groups  is  likely  to  increase. 

On  this  premise,  Target  Group  II  has  been 
constructed.  Those  fishermen  who  recorded 
an   income/effort   ratio   of  less   than   0.2   were 


Table  6.  —  Distribution  of  sample  lobstermen  according  to  income/effort  ratio  and  trap-days. 


Trap-days  fished 

per  year 

Earning  effort 
ratio 

5,000 

5,001- 
10,000 

10,001- 
20,000 

20,001- 
30,000 

30,001- 
40,000 

40,001- 
50,000 

50,001- 
60,000 

60,000+ 

N/I 

TOTAL 

0.100 
.100-.199 
.200-.299 

! 

2 
5 

2 
3 

1 

1 

7 
8 

8 
2 

5 

4 

1 
4 

2 
4 
6 

8 

1 

- 

7 
41 
27 

.300-.399 

- 

2 

2 

2 

2 

1 

1 

- 

- 

ID 

.400-.499 

.500  + 

N/I 

2 
6 

2 
1 

1 
1 

2 

2 
1 

- 

1 
1 

1 

5 

19 

4 

5 
37 

TOTAL 

18 

10 

20 

14 

14 

7 

15 

14 

19 

131 

Source:      University  of  Maine  Survey  Data,  1970. 


165 


considered  eligible  for  this  group  (See  Table 
6).  There  will  be  some  overlap  between  this 
group  and  Target  Group  I. 

Different  combinations  of  investment  and 
effort  suggest  other  possible  approaches  to 
management  alternatives.  For  instance,  one 
could  identify  a  group  that  represents  rela- 
tively high  effort  and  low  investment  input 
combination;  another  group  may  represent 
relatively  higher  investment  and  lower  effort 
input  combination.6  The  reasoning  for  at 
least  considering  these  groups  as  possible 
target  groups  may  be  explained  as  follows: 
in  the  absence  of  any  precise  knowledge  about 
the  optimum  combination  of  effort  and  invest- 
ment, two  contrasting  groups  —  high-effort 
low-investment  versus  low-effort  high-invest- 
ment —  might  suggest  alternative  goals  for 
management  strategies.  For  instance,  one 
might  consider  eliminating  excessive  capital 
versus  eliminating  excessive  effort  as  possible 
goals.  As  a  minimum,  the  differences  in  socio- 
economic impact  of  such  changes  should  be 
examined. 

It  is  reasonable  to  assume  that  excess 
capacity  exists  in  the  lobster  fishery,  although 
it  is  difficult  to  establish  whether  such  excess 
capacity  is  due  to  excessive  effort  or  excessive 
investment   or   both.    Under   these   conditions, 


6  This  approach  was  suggested  by  Dr.  Adam  A. 
Sokoloski,  National  Marine  Fisheries  Service  in  per- 
sonal correspondence  dated  December  16,  1970. 


it  seems  meaningful  to  isolate  for  analytical 
purposes,  two  cases,  one  showing  evidence 
of  excessive  effort  and  the  other  of  excessive 
investment.  Admittedly,  the  state  of  the  art 
does  not  provide  absolute  measurement  of 
excess  capacity  either  in  terms  of  effort  or  in 
terms  of  investment. 

Target  Group  III  has  been  constructed  to 
reflect  excessive  effort  in  the  sense  that  these 
fishermen  supply  a  large  amount  of  labor  to 
their  operation  relative  to  their  investment. 
They  fish,  on  an  average,  150  days  per  year 
compared  to  109  days  for  the  entire  sample; 
their  average  investment  amounted  to  $4,410 
compared  to  $7,575  for  the  entire  sample. 
As  a  practical  device,  the  criteria  of  those 
fishing  over  100  days  per  year  with  investment 
of  less  than  $8,000  in  gear  were  used  to  select 
the  candidates  for  this  group  of  28  fishermen. 

Target  Group  IV  represents  excessive  capi- 
tal in  the  sense  that  the  fishermen  in  this 
group  have  substantial  investments  in  gear 
relative  to  the  number  of  days  per  year  fished. 
On  the  average  they  have  invested  $12,410 
compared  to  $7,575  for  the  entire  sample  and 
they  fish  an  average  of  78  days  per  year  com- 
pared to  109  days  per  year  for  the  sample. 
This  group  of  six  fishermen  included  those 
who  have  invested  more  than  $8,000  and  who 
fish  less  than  100  days  per  year. 

Table  7  provides  the  basic  information  from 
which  Target  Groups  III  and  IV  have  been 
derived. 


Table  7.  —  Distribution  of  sample  lobstermen  by  investment  and  number  of  days  fished. 


Investment 

in  gear 

Days  fished 

2,001- 

4,001- 

8,001- 

12,001- 

16,001- 

20,001- 

per  year 

2,000 

4,000 

8,000 

12,000 

16,000 

20,000 

24,000 

24,000+ 

N/I 

Total 

dollars 

50 

HI 

3 

_ 

1 

_ 

_ 

_ 

_ 

3 

17 

51-100 

16 

7 

8 

2 

2 

1 

- 

- 

2 

38 

101-150 

3 

7 

8 

1 

4 

3 

2 

4 

- 

32 

151-200 

- 

2 

5 

6 

2 

2 

1 

- 

- 

18 

201-250 

- 

1 

2 

- 

1 

- 

1 

1 

- 

6 

N/I 

- 

- 

- 

- 

1 

- 

1 

- 

18 

20 

TOTAL 

29 

20 

23 

10 

10 

6 

5 

5 

23 

131 

Source:      University  of  Maine  Survey  Data,  1970. 


166 


Table  8.  —  Distribution  of  lobstermen  in  target  groups  by  trap-days,  gross  income,  and  capital  invested. 


Trap-days 


I 


502,799 


II 


Target  groups 
III 


1,753,287 


973,198 


IV 

185,560 


Total 
Sample 

3,470,000 


% 

*(No.),  % 

14.5 
(40)  32.0 

50.5 
(48)  38.4 

28.0 
(28)  22.4 

5.3 
(6)4.8 

(113) 

Income 

$82,450 

$250,233 

$161,583 

$61,000 

$596,500 

% 
*(No.),% 

13.8 
(40)41.7 

41.8 
(48)  50.0 

27.0 
(26)  27.1 

10.2 
(5)5.2 

(96) 

Capital 
% 
*(No.),  % 

$97,043 

11.6 
(40)  36.4 

$332,566 

39.9 
(48)43.6 

$123,485 

14.8 
(23)  25.5 

$74,465 

8.9 
(6)5.5 

$833,209 
(110) 

*The  number  in  parentheses  refers  to  the  total  number  of  fishermen  relevant  to  a  particular  category;  the  other  number  is  the  relevant 
number  of  fishermen  expressed  as  a  percentage  of  the  sample. 
Source:      University  of  Maine  Survey  Data,  1970. 


Distribution  by  Trap-days,  Income, 
and  Capital  Invested 

Table  8  presents  a  distribution  of  the  lobster- 
men  in  each  of  the  target  groups  by  trap-days, 
gross  income  and  capital  invested  in  boat  and 
gear.  Target  Group  I  emerges  as  a  critical 
group  in  that  its  share  in  trap-days,  income 
and  capital  investment  is  the  lowest  relative 
to  its  size  in  the  total  sample.  Target  Group 
II  contributes  more  trap-days,  more  capital, 
and  more  income  compared  to  Group  I.  How- 
ever, relative  to  its  size,  its  share  in  income 
and  capital  investment  is  less  than  in  propor- 
tion. Target  Group  III  contributes  relatively 
more  in  trap-days  and  relatively  less  in  capital 
and  its  income  share  corresponds  closely  to 
its  size.  Target  Group  IV  accounts  for  more 
capital  relative  to  size  and  to  number  of  trap- 
days  and  substantially  more  income  relative 
to  size.  For  this  reason,  this  group  can  hardly 
be  considered  as  a  target  group  for  limited 
entry  on  the  basis  of  income-effort  relation- 
ship. However,  if  the  income-capital  ratio  is 
considered,  this  group  does  not  appear  to 
be  equally  efficient. 

Socioeconomic  Characteristics  of  the 

Fishermen  in  Each  of  the 

Four  Target  Groups 

Beals  will  be  most  affected  if  Target  Group 
II  is  eliminated,  and  Corea  the  least.  If  Target 


Group  I  is  considered,  the  impact  on  the  three 
communities  is  comparable.  Corea  will  be  af- 
fected in  the  least  if  one  focuses  on  Target 
Group  III.  The  effect  on  the  other  two  com- 
munities is  about  the  same.  Target  Group  IV 
does  not  affect  Phippsburg  but  will  affect 
the  other  two  communities  equally  (Table  9). 
Table  10  provides  average  values  for  certain 
socioeconomic  characteristics  of  the  lobster- 
men  in  each  of  the  Target  Groups. 

Table  9.  —  Geographic  distribution 

Target  groups 


1  II 

Community       No.        %       No.       % 


III  IV 

No.        %       No.      % 


Beals1 
Corea2 
Phippsburg3 


18      29.5       31      50.8       16      26.2       4       6.5 
7      26.9         3      11.5         3      11.5        2       7.7 
15      34.1       14      31.8        9      26.5 


Total 


41) 


4S 


'Beals  61. 

2Corea  26. 

3Phippsburg  44,  includes  10  from  Bath. 

Source:      University  of  Maine  Survey  Data,  1970. 


The  average  income  of  Group  I  is  the  lowest 
attributable  both  to  low  labor  and  low  capital 
intensity  in  its  operation.  In  constrast,  Group 
IV  has  the  highest  average  income  primarily 
due  to  high  capital  intensity  in  its  operation 
in  spite  of  low  labor  intensity.  Group  II  ranks 
second  in  average  income  which  can  be  ex- 
plained in  terms  of  relatively  more  effort  and 


167 


Table  10.  —  Comparative  average  value  for  selected  socioeconomic  variables  in  the  sample  of  lobstermen  and  the  four 

target  groups. 


Target  groups 

Socioeconomic  variable 

Sample 

1 

II 

III 

IV  . 

Family  size 

3.2(122) 

2.9  (38) 

3.6(46) 

2.9(28) 

3.6(5) 

Age 

42.4(131) 

42.5  (40) 

44.0  (48) 

49.4  (28) 

31.7(6) 

Education:   years 

9.8(126) 

9.7  (40) 

9.7  (48) 

10.0(28) 

11.0(6) 

Investment  (gear  &  boat) 

$7,575  (110) 

$2,426  (40) 

$6,949  (48) 

$4,410(28) 

$12,410(6) 

Gross  income 

$6,213    (96) 

$2,061  (40) 

$5,213  (48) 

$6,214(26) 

$12,200(5) 

Months  per  year  fished 

7.2(113) 

5.7  (40) 

8.0  (48) 

8.5  (28) 

6.6(5) 

Trap-days  per  year 

30,707  (113) 

12,570(40) 

36,526  (48) 

34,757  (28) 

30,927  (6) 

Days  per  year  lobstered 

109.2(113) 

87.0  (40) 

132.2(47) 

147.9(28) 

78.0(6) 

Earning-effort  ratio 

.230    (96) 

.182(40) 

.140(48) 

.183(26) 

.355  (5) 

*The  number  in  parentheses  refers  to  the  total  number  of  fishermen  relevant  to  a  particular  category. 
Source:      University  of  Maine  Survey  Data,  1970. 


capital  used  compared  to  Groups  I  and  III. 
Group  III  ranks  third  in  average  income. 
Here  the  high  level  of  labor  intensity  offset 
the  effect  of  low  capital  intensity.  Its  income/ 
effort  ratio  is  almost  the  same  as  that  of 
Group  I. 

Socioeconomic  Impact  of  Changes 
in  Harvesting  Labor  Force 

As  pointed  out  earlier,  the  different  target 
groups  were  constructed  on  the  basis  of  differ- 
ent criteria  such  as  low  earnings/effort  ratio, 
low  level  of  both  effort  and  investment,  high 
labor  and  capital  input  combination.  The 
rationale  for  this  procedure  is  simply  to  facili- 
tate comparative  analysis  of  alternative  man- 
agement strategies.  For  instance,  one  might 
consider  limiting  entry  on  the  basis  of  low 
earnings/effort  ratio  combined  with  low  level 
of  income  (Group  I);  one  might  also  focus 
on  low  earnings/effort  ratio  regardless  of 
the  level  of  income  (Group  II);  alternatively, 
one  might  emphasize  high  labor-low  capital 
input  combination  associated  with  low  income 
as  an  indicator  of  inefficiency  (Group  III); 
finally,  high  capital-low  labor  input  combina- 
tion regardless  of  a  relatively  higher  level  of 
income  may  be  construed  as  an  indicator  of 
excess  capacity  (Group  IV). 

It  should  be  noted  that  it  was  not  the  pur- 
pose of  this  study  either  to  advocate  or  repudi- 
ate any  particular  management  strategy  and 
its  implicit  goal.  The  intent  here  is  simply  to 
analyze    the    potential    socioeconomic    impact 


of  a  change  in  the  harvesting  labor  force  in 
the  Maine  lobster  fishery  if  such  a  change 
amounts  to  reducing  inefficient  inputs  from 
given  target  groups. 

For  the  purpose  of  this  study  such  impact 
is  analyzed  primarily  in  terms  of  employment 
effects  and  income  effects  relative  to  the  target 
group  populations  and  the  local  economy. 


Employment  Effects 

Taking  into  consideration  the  employment- 
related  variable  such  as  skills  either  from 
currently  held  part-time  jobs  or  alternative 
jobs  held  in  the  past,  level  of  education,  and 
age,  a  simplified  profile  of  labor  market  par- 
ticipation potential  of  the  target  groups  is 
shown  in  Table  11. 

The  category  "potentially  employable"  in- 
cludes those  individuals  who  have  market- 
able skills  acquired  from  formal  vocational 
training  and/or  alternative  job  experience. 
This  survey  information  was  supplemented 
by  information  on  the  local  labor  market 
through  the  cooperation  of  the  regional  offices 
of  the  Maine  Employment  Security  Commis- 
sion. If  there  was  a  match  between  the  kinds 
of  skills  in  demand  in  the  local  labor  market 
and  the  skills  possessed,  an  individual  was 
considered  eligible  for  the  category  "potential- 
ly employable." 

The  category  "possibly  trainable"  includes 
those  who  on  the  basis  of  age  and  level  of 
education  would  be  likely  to  benefit  from  and 


168 


Table  11.  —  Labor  market  participation  potential  of  target  groups  I-IV. 


Target 
group 

Total 
number 

Potentially 
employable1 

Possibly 
trainable2 

Po 
core 

ential  hard- 
unemployed3 

Not  in  the 
labor  force4 

I 

40 
'       100.0% 

14 
35.0% 

4 
10.0% 

8 
20.0% 

14 

35.0% 

IJ 

48 

100.0% 

18 
37.5% 

4 
8.3% 

17 

35.4% 

9 

18.7% 

III 

28 
100.0% 

11 

39.3% 

2 
7.1% 

10 

35.7% 

5 
17.9% 

IV 

6 
100.0% 

4 

66.7% 

1 

16.7% 

1 

16.7% 

- 

'Those  having  marketable  skills. 

2Those  having  no  skill  but  less  than  35  years  of  age. 

3Those  having  no  skill  and  in  the  age  bracket  35-65  years. 

4Students  and  those  over  65  years. 

Source:      University  of  Maine  Survey  Data,  1970. 


be  capable  of  participating  in  a  training  pro- 
gram. Admittedly,  this  is  only  a  first  approxi- 
mation. 

The  category  "potential  hard-core  unemploy- 
ed" includes  those  fishermen  who  have  no 
marketable  skills  other  than  lobstering  and 
who  fall  into  the  critical  age  bracket  by  labor 
market  criteria,  35-65.  In  all  likelihood,  these 
individuals,  if  excluded  from  lobstering,  will 
find  it  extremely  hard  to  make  any  vocational 
readjustment. 

The  last  category,  "not  in  the  labor  force" 
is  self-explanatory.  This  includes  those  fisher- 
men who  are  either  students  or  over  65  years 
of  age  and  are  not  likely  to  participate  in 
the  labor  market  as  active  job  seekers,  barring 
purely  part-time  or  seasonal  jobs. 

It  should  be  emphasized  that  the  above 
classification  is  only  a  preliminary  step  in 
identifying  the  differences  in  labor  market 
participation  potential  of  various  subgroups 
within  each  of  the  target  groups.  To  be  sure, 
potential  employability,  trainability,  and  hard- 
core unemployability  require  considerably 
more  in  depth  analysis  than  was  possible  in 
the  present  study. 

It  is  apparent  from  Table  11  that  a  sub- 
stantial proportion  of  the  fishermen  in  each 
of  the  target  groups  is  potentially  employable 
(ranging  from  35%  to  67%).  Of  those  who  are 
classified  under  "potentially  employable,"  some 
already  have  full-time  jobs  and  others  have 
marketable  skills.  However,  Target  Groups 
II  and  III  are  likely  to  result  in  more  hard- 
core unemployment.  Paradoxically,  the  group 


that  has  a  high  earnings/effort  ratio  (Target 
Group  IV)  also  happens  to  be  the  one  with  a 
relatively  larger  proportion  of  potential  em- 
ployability. With  the  exception  of  this  group, 
other  groups  include  several  fishermen  not 
in  the  labor  force,  students,  and  those  65  years 
and  over.  The  question  of  their  employability 
is,  therefore,  irrelevant  in  the  present  context. 
In  analyzing  the  expected  socioeconomic 
impact  of  limited  entry,  the  survey  data  on 
each  of  the  fishermen  in  each  of  the  target 
groups  were  examined  in  depth  by  communi- 
ties. In  this  investigation,  attention  was  focus- 
ed on  such  socioeconomic  variables  as  age, 
family  size,  level  of  education,  types  of  skill, 
alternative  job  experience,  alternative  source 
of  income,  and  so  on.  On  the  basis  of  informa- 
tion from  survey  data  combined  with  informa- 
tion on  local  labor  market,  Table  12  is  recon- 
structed to  reflect  the  differences  in  labor 
market  participation  potential  by  communities. 


Income  Effect  and  Expected 
Socioeconomic  Impact 

To    perform    the    necessary     analysis,     the 
following  procedures  were  adopted: 

1.  Assume  each  target  group  to  be  a  candi- 
date for  exclusion  from  lobstering. 

2.  Estimate  private  loss  of  gross  income  due 
to  non-participation  in  lobster  fishery. 

3.  Assume  that  50%  of  the  lost  gross  income 
would  be  subsequently  earned  by  the  re- 
maining fishermen.  The  survey  date  did 


169 


Table  12.  —  Labor  market  participation  potential  of  target  groups  I-IV  by  geographic  location. 


Target  aroup 

'  by 
communities 

Total 

number 

Potentially 
employable1 

Possibly 
trainable2 

Po 

core 

tential  hard- 
unemployed3 

Not  in  the 
labor  force4 

Phippsburg 
I  Corea 
Beals 

15 

7 
18 

7 
3 

4 

2 
1 
1 

3 

5 

3 
3 
8 

40 

14 

4 

8 

14 

Phippsburg 
II  Corea 
Beals 

14 

3 

31 

8 

1 
9 

1 
3 

4 
13 

2 
1 
6 

48 

18 

4 

17 

9 

Phippsburg 
III  Corea 
Beals 

9 

3 

16 

5 

2 
4 

1 

1 

3 

7 

1 

4 

28 

11 

2 

10 

5 

Phippsburg 
IV  Corea 
Beals 

2 

4 

1 

3 

1 

1 

- 

6 

4 

1 

1 

- 

'Those  having  marketable  skills. 

2Those  having  no  skill  but  less  than  35  years  of  age. 

3Those  having  no  skill  and  in  the  age  bracket  35-65  years. 

4Students  and  those  over  65  years. 

Source:      University  of  Maine  Survey  Data,  1970. 


indicate  some  evidence  of  excess  capacity 
in  terms  of  number  of  traps  owned  and 
number  of  traps  fished  and  days  fished. 
It  was  recognized  that  the  remaining 
fishermen  may  not  be  willing  or  able  to 
capture  the  entire  amount  of  output  at- 
tributable to  the  excluded  fishermen,  at 
least  in  the  short  run.  Furthermore,  the 
purpose  here  is  to  illustrate  what  might 
happen  if  this  assumption  holds.  If  a 
different  figure  proves  to  be  more  realis- 
tic, the  results  will  change. 
Estimate  the  savings  in  effort  measured  in 
trap-days  on  the  basis  of  (3)  and  convert 
this  into  monetary  values.  For  this  pur- 
pose, we  first  calculated  how  many  trap- 
days  would  be  needed  by  the  excluded 
fishermen  in  a  given  target  group  to 
produce  the  gross  income  attributed  to 
this  group.  An  average  earnings/effort 
ratio  for  this  group  was  used  to  calcu- 
late the  number  of  trap-days  required. 
Next,  an  average  earnings/effort  ratio 
was  computed  in  the  given  target  group. 


This  average  ratio  was  applied  to  50% 
of  the  total  gross  income  of  the  group  to 
come  up  with  the  number  of  trap-days 
that  would  be  required  to  produce  this 
income  by  the  remaining  fishermen.  The 
difference  between  the  two  values  for 
trap-days  is  stated  as  saving  in  effort. 
This  quantity  multiplied  by  the  average 
earnings/effort  ratio  of  the  remaining  fish- 
ermen produced  a  monetary  measure  of 
saving  that  can  be  expected  under  the 
stipulated  conditions. 

5.  Estimate  the  sum  of  expected  new  in- 
comes generated  by  those  who  are  con- 
sidered "potentially  employable"  based 
on  information  of  types  of  jobs  available 
and  skills  needed  in  the  local  market. 
The  number  of  fishermen  in  each  target 
group  that  fits  this  category  was  identi- 
fied and  typical  wages  for  indicated  jobs 
were  applied  to  the  number  of  employ- 
able fishermen  to  produce  a  sum  of  ex- 
pected income. 

6.  Estimate  the  expected  annual  income  of 


170 


those  that  are  classified  as  "possibly  train- 
able." Assume  that  training  facilities 
and  programs  are  made  available  and 
that  individuals  are  willing  to  paticipate. 
Communication  from  people  involved 
with  Manpower  Development  and  Train- 
ing Act  (MDTA)  programs  provided  some 
information  as  to  typical  wages  MDTA 
trainees  can  expect  post-training.  These 
figures  were  used  to  derive  expected  in- 
comes that  the  "possibly  trainable"  fish- 
ermen in  each  target  group  can  expect 
if  they  receive  training  comparable  to 
those  under  MDTA  programs. 

7.  Estimate  the  training  cost  of  those  classi- 
fied under  "possibly  trainable." 

8.  Estimate  the  potential  income-mainte- 
nance burden  on  society  imposed  by  the 
loss  of  lobstering  income  of  those  who 
are  classified  under  "potentially  hard- 
core unemployed"  and  under  "not  in 
the  labor  force."  Fifty  percent  of  current 
gross  income  from  lobstering  was  used 
for  estimation  purposes.  The  rationale 
for  using  this  percentage  is  based  on  the 
consideration  that  the  net  income  from 
lobstering  is  substantially  lower  than 
reported  gross  income,  although  exact 
figures  for  net  income  were  not  readily 
obtainable.  During  the  course  of  the 
interviews,  several  fishermen  indicated 
that  although  they  could  not  provide  in- 
formation on  net  income,  roughly  50% 
of  their  gross  income  could  be  considered 
net,  after  allowing  for  business  expenses. 
The  assumed  percentage  is  considered 
reasonable  for  illustrative  purposes. 

The  reason  why  the  individuals  in  these 
categories  —  "potential  hard-core  unemploy- 
ed" and  "not  in  the  labor  force"  —  and  their 
loss  of  income  from  lobstering  are  used  as 
the  basis  for  measuring  the  income  mainte- 
nance burden  on  society  is  to  indicate  the 
upper  limit  of  the  social  burden.  This  yields 
a  relative  measure  of  income  loss  and  corres- 
ponding welfare  loss  for  a  group  of  people 
who  are  technically  outside  the  labor  force. 
At  least  in  the  short  run,  the  process  of  ad- 
justment will  be  quite  severe  for  a  bulk  of 
this  group.  Conceivably,  some  low  level,  un- 
skilled jobs  would  be  available  which  would 
moderate    the    impact.    However,    considering 


the  high   level  of  current  unemployment  and 
the  generally  depressed  conditions  of  the  local 
economies    under    consideration,    it    appeared 
reasonable  to  assume  that  alternative  sources 
of  income  would  be  unavailable   in  the  short 
run,  thereby  imposing  a  burden  on  society. 
9.  The    estimated    value    of    investment    in 
boat  and  gear  by  the  fishermen  in  each 
of  the  target  groups   is   included   in   the 
profile  of  socioeconomic  impact  of  limited 
entry  because  these  values  have  definite 
implications  for  compensation. 
Assuming  zero  salvage  value  of  such  capital 
equipment,  the  stated  figures  provide  the  upper 
limit  of  the  compensation  burden  imposed  on 
society.  It  is  reasonable  to  think  actual  com- 
pensation  will   differ  from   the  stated    figures 
because   of  some   positive   salvage   value.    For 
illustrative    purposes,    without    making    such 
allowance,  the  quoted  figures  do  serve  as  indi- 
cators of  upper  limits  of  the  cost  of  compensa- 
tion that  may  be  entailed. 

Using  the  above  procedure,  the  following 
tabulations  were  made  to  present  a  compara- 
tive picture  of  the  socioeconomic  implications 
of  limiting  entry  of  different  groups  by  using 
alternative  criteria  (Table  13). 

Group  II  is  likely  to  cause  the  largest  de- 
cline in  income  from  lobstering.  It  will  be 
partially  offset  by  additional  income  from 
lobstering  by  the  remaining  fishermen,  income 
from  alternative  jobs  for  the  displaced  fisher- 
men, and  the  savings  in  effort  measured  by 
the  fewer  number  of  trap-days  required  to 
capture  at  least  50%  of  the  gross  income  lost. 
In  absolute  terms,  this  group  may  present 
the  severest  income  maintenance  burden  on 
society.  By  comparison,  Group  I  is  likely  to 
impose  a  relatively  smaller  burden  on  society. 
On  a  per  capita  basis,  Group  III  will  impose 
the  severest  burden  on  society. 

The  proportion  of  the  "potentially  employ- 
able" and  "possibly  trainable"  among  Groups 
I-III  are  quite  comparable.  The  proportion 
of  the  same  categories  for  Group  IV  is  con- 
siderably higher.  This  accounts  for  the  rela- 
tively small  social  burden  indicated  for  this 
group.  However,  it  should  be  noted  that  this 
underestimates  the  total  real  burden  on  society 
in  that  there  will  be  a  dissaving  in  effort  and 
potential  negative  difference  between  their 
current  income  from  lobstering  and  their  ex- 


171 


Table  13.  —  Profile  of  socioeconomic  impact  by  target  groups. 


Impact  variables 


Target 

groups 

I 

II 

III 

IV 

-82,450 

-250,223 

-161,583 

-61,000 

+41,225 

+  125,116 

+  80,791 

+  30,500 

+18,574 

+  168,670 

+  31,346 

-11,083 

+  19,000 

+  41,500 

+  38,000 

+21,000 

+24,000 

+   24,000 

+   12,000 

+  12,000 

-13,800 

-   13,800 

-     6,400 

-  6,400 

-26,775 

-  64,225 

-  54,200 

-  3,500 

-97,043 

-332,566 

-123,485 

-74,465 

40 

48 

28 

6 

1.  Loss  of  income  from  lobstering   (S) 

2.  Gain  of  income  from  lobstering  ($) 

3.  Monetary  value  of  saving  in  effort  ($) 

4.  Gain  of  income  from  alternative  jobs  (marketable 
skills)  ($) 

5.  Gain  of  income  from  alternative  jobs  (post- 
training)  ($) 

6.  Training  costs  ($) 

7.  Income  maintenance  burden  on  society  ($) 

8.  Estimated  value  of  investment  in  boat  and  gear  ($) 

9.  Number  of  fishermen 


Source:      University  of  Maine  Survey  Data,  1970;  local  Manpower  Development  Training  Act  program  officials. 


pected  income  from  alternative  jobs. 

It  would  have  been  desirable  to  compute 
a  ratio  of  total  gains  and  losses.  However, 
with  the  data  in  hand,  it  does  not  appear  to 
be  feasible  and  meaningful.  First,  the  quanti- 
ties calculated  are  not  additive.  Second,  costs 
and  benefits  have  different  time  dimensions. 
For  instance,  training  costs  are  once-over 
cost  items  whereas  the  expected  income  is 
a  flow  over  time.  Finally,  the  figures  for  in- 
come maintenance  burden  on  society  do  not 
take  into  consideration  the  loss  of  income 
from  lobstering  of  those  who  are  classified  as 
"potentially  employable"  but  are  already 
employed.  Furthermore,  the  discrepancy  be- 
tween current  income  from  lobstering  and 
expected  income  from  alternative  jobs  for  those 
employable  but  currently  full-time  fishermen 
is  also  disregarded. 

Despite  these  limitations,  the  results  do 
give  certain  indicator  values  that  should  be 
considered  and  comparatively  analyzed  rela- 
tive to  alternative  management  strategies 
and  implicit  goals.  Admittedly,  these  values 
involve  many  simplifying  and  rather  arbi- 
trary assumptions,  although  hard  data  were 
utilized  when  available.  The  value  of  this 
type  of  approach  is  primarily  methodological, 
which  is  to  be  expected  in  a  pilot  study. 

CONCLUSIONS 

Several  qualifications  need  to  be  attached 
to  the  foregoing  analysis  before  any  general- 
ization is  made.  First,  some  fishermen  who 
are  considered  as  candidates  for  a  given  target 


group  may  continue  to  lobster  because  of  non- 
economic  reasons.  Second,  expected  new  in- 
comes from  alternative  jobs  for  the  displaced 
fishermen  may  not  materialize  because  of  lack 
of  motivation  and  reluctance  to  move  geo- 
graphically and/or  occupationally.  Third, 
there  is  no  assurance  that  the  additional  new 
income  earned  by  the  remaining  lobstermen 
will  exactly  equal  the  lost  income  due  to 
limited  entry.  There  is,  however,  a  strong 
probability  that  if  they  were  to  capture  the 
same  number  of  lobsters  as  attributable  to 
the  displaced  fishermen,  they  could  do  so 
more  efficiently  because  of  excess  capacity 
and  potential  economies  of  scale.  Fourth,  there 
may  be  a  significant  gap  between  the  number 
of  those  considered  trainable  and  those  who 
will  take  advantage  of  training  if  made  avail- 
able. Fifth,  a  fraction  of  those  trained  may 
still  remain  unemployed  due  to  labor  market 
conditions.  Sixth,  the  income  maintenance 
burden  may  not  be  as  severe  as  indicated  be- 
cause some  of  the  potentially  hard-core  un- 
employed may  be  absorbed  in  unskilled  jobs 
or  in  the  lobster  industry  as  "helpers."  Con- 
ceivably, jobs  may  be  redesigned  to  facilitate 
the  entry  of  these  men  into  the  labor  market. 
Finally,  some  of  those  who  are  not  in  the  labor 
force,  e.g.,  students,  will,  in  course  of  time,  par- 
ticipate in  the  labor  market  and  reduce  the 
stated  social  burden. 

It  is  important  that  in  this  kind  of  analysis 
one  takes  cognizance  of  the  time  element 
relative  to  the  process  of  adjustment.  The 
short  run  impact  may  appear  to  be  quite 
severe    because    of   the    imperfections    in    the 


172 


labor  market.  For  instance,  men  who  are 
unemployed  now  may  not  have  marketable 
skills;  men  who  have  marketable  skills  may 
not  have  information  about  available  jobs  or 
may  have  very  restricted  mobility;  job  struc- 
ture may  be  such  that  it  precludes  entry  of 
unskilled  workers;  those  who  are  trainable 
may  not  have  access  to  adequate  training 
facilities  or  programs.  Given  time,  however, 
some  of  these  market  imperfections  may  be 
reduced,  partially  through  deliberate  planning 
and  partially  through  autonomous  changes 
in  the  labor  market  itself.  For  instance,  the 
quality  of  job  information  and  job  counselling 
can  be  improved;  training  programs  may  be 
initiated;    jobs    may    be    restructured;     local 


economic  development  may  generate  new  de- 
mands for  labor;  the  lobster  fishery  itself,  if 
efficiently  managed  by  fewer  fishermen,  may 
need  additional  helpers. 

It  is  a  reasonable  expectation  that  if  a 
management  strategy  results  in  an  improved 
return  to  both  labor  and  capital  and  if  de- 
liberate efforts  are  made  to  aid  the  process 
of  adjustment,  net  social  gains  are  likely  to 
materialize  in  the  long  run.  Although  the 
present  study  did  not  consider,  nor  was  in- 
tended to  consider,  any  specific  management 
scheme  with  respect  to  its  socioeconomic 
impact,  it  did  generate  data  pertinent  to  such 
an  evaluation. 


A  U.S.  GOVERNMENT  PRINTING  OFFICE:   1973-795-774  /  9     REGION   10 


173 


349.  Use  of  abstracts  and  summaries  as  communica- 
tion devices  in  technical  articles.  By  F.  Bruce 
Sanford.     February  1971,  iii  +  11  pp.,  1  fig. 

350.  Research  in  fiscal  year  1969  at  the  Bureau  of 
Commercial  Fisheries  Biological  Laboratory, 
Beaufort,  N.C.  By  the  Laboratory  staff.  No- 
vember 1970,  ii  +  49  pp.,  21  figs.,   17  tables. 

351.  Bureau  of  Commercial  Fisheries  Exploratory 
Fishing  and  Gear  Research  Base,  Pascagoula, 
Mississippi,  July  1,  1967  to  June  30,  1969.  By 
Harvey  R.  Bullis,  Jr.,  and  John  R.  Thompson. 
November  1970,  iv  +   29  pp.,  29  figs.,  1  table. 

352.  Upstream   passage  of  anadromous  fish  through 


navigation  locks  and  use  of  the  stream  for  spawn- 
ing and  nursery  habitat,  Cape  Fear  River    N  C 
1962-66.     By   Paul    R.   Nichols  and    Darrell   E.' 
Louder.     October  1970,   iv  +   12  pp.,   9   figs.    4 
tables. 

356.  Floating  laboratory  for  study  of  aquatic  organ- 
isms and  their  environment.  By  George  R. 
Snyder,  Theodore  H.  Blahm,  and  Robert  J.  Mc- 
Connell.     May  1971,  iii   +   16  pp.,  11    figs. 

361.  Regional  and  other  related  aspects  of  shellfish 
consumption  —  some  preliminary  findings  from 
the  1969  Consumer  Panel  Survey.  By  Morton 
M.  Miller  and  Darrel  A.  Nash.  June  1971,  iv  + 
18  pp.,  19  figs.,  3  tables,  10  apps. 


UNITED  STATES 
DEPARTMENT  OF  COMMERCE 

NATIONAL  OCEANIC  &  ATMOSPHERIC  ADMINISTRATION 

NATIONAL  MARINE   FISHERIES  SERVICE 

SCIENTIFIC   PUBLICATIONS  STAFF 

BLDG.  67,   NAVAL  SUPPORT  ACTIVITY 

SEATTLE,  WASHINGTON  98115 


OFFICIAL   BUSINESS 


PENN  STATE  UNIVERSITY  LIBRARIES 


ADDDD7ED 

210