Skip to main content

Full text of "Population status of California sea otters"

See other formats


Feaecs sa : ay TRAE A IRS ne ALM St 
era 3 i Sa telaten views, 
i she Deere erates 

estes 
Tanta sczhesank 


Pegs stes 
Viana 


ance nasa roe tcnevry 
ARIA ONS MENTS PRA Sey 
Sil en tena ned Aas 

fi Ria te 


Atta 
n Grn Maieeinssin aa 
Dade Metance ets 
FO aM AMIR) 
TA Base ea ate “ 
UA Be 


De sn oko ie 
Fae ane 


a Avatmawsmeibnh chee tas 
fnew ite | meses 


Wayen aw a wos aia ahind stay eal 
Hig eios ma ernie ectntn? 
Ag ne Sk eara casa seataahans 
Spy Aca NAR GAD Baten hs 


SEs 
Salas 
aU Aan 
Sramtihaeas antes ee 


Sigh aaa 
tink nam wane 
Setanta 


Sess 
Bento tiat 
Jaaan edd DAES me At Vey MICAS, 
2 MAD wath 
Pectin 


RAs ree 
RIT Mal el are 


Cesta 


eatery 
cheated pat 
pero ime te 
iheesietet 


Happier 
LEM ierPaPre Sioa 
Ltga replete: 


peter Tks ot oaks 
Hoenn 


Co salecamesen 
Sr AE eRe 
Awienedssuccaten ye 


ERE ee 
Nat 
CA epastets 
ide missed ety ine tr tate gas 
ret ee Feat 
Speteasepgaqrcrue tetas Of ail Mekgeat ots ge 

Je a reaty tae A A oUnNe portly aay 

Powe thee nese lua ety ta s TASS (Lynde kg manne se chee 
Panties > ? elag try oy ade 


Vocus aterpeometarquirty ere 
Katoree tr evarlay eyes 
wie Pianl uy Raia ht ba Sue teed vies 
iar eelurge ae sedaielenes thay Peeled fh 
Wanaaren ne t 4 Rirarphrereitct 
Spt eat 
nt satin ya gew aus eas 
inate : reeeenees y 
Phir tee vir ; ragievindte 
tau aie 


Je tauraeemygetcr eye 
Coe Ce eae 
Hitieacae ead 
De 
very Vistar 
haateaya eaeeueu lear. 


oH 
TE SAHG THT 

Hepat ante 

itty ies at 
runs pi e ‘ i piu aay UR howe 
YESSY revty ieausee tent 1 i Rear es 


eariveny! 


‘akeivinsacuny 
Tees) 


Care ca yea 
sreaa cece nwitieys 


en Y REF ter g a 
veel ean re ra) 

SH Ate a Ansara pain) Geth 
‘ 2) igus Tieebecipee ov airnees 


oR 
Sia save 


Hea 
m pp retatea 
ie eer esu Ne ean Uses 
Dr ote : Pee AEC NER 
By WP 8: Mee Weaae 
aga gana frets 


ort age eh 
el eraeneoengeeees 
Sea uaNe undead 


a ee ce: = tra ) \ FS —_— {- y > - re ¥ = ww “7AN 
Kaw) = eu) 2 SE et 2 Get GZ? es E ty 2° 
@ J = 2 S E 2 eS Mey 2 Ene kK 
Coke S b Xe) Sy . a NO, Y/Y - & ay a Ly —- Xe wb XO g 
NDS “nse n <n4is8> m QS UG m CN.DE 2 ns 
z z 7) = o = 


}1YVYGIT_LIBRARIES SMITHSONIAN INSTITUTION NOILALILSNI NVINOSHLINS SSIYWYRIT_LIBRARIES SM 


Ly, 
NVINOSHLINS S3IYV 
Y 
Ly 
Uj 
NVINOSHLINS S3IYV 
SAC 


SMITHSONIAN 


SMITHSONIAN 


Y 
NVINOSHLIWS 
NG as 
By 
SMITHSONIAN 


NVINOSHLIWS 


SMITHSONIAN 


% 
» 


LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSNI NV 


* 


STITUTION NOILMLILSNI NVINOSHLIWS saryvagi 


2p) 4 w = ” > i!p) — 

w & a 3 hae a o 

oc = = “.¢Y ZB = oc = 

< = < a Li YY, x < = 4 

a c oc S a Yz- Ss a Cc! 

a 5 2 5 a ““ 3 a 5 

a z SOS) 2 3 4 z a 2 
1kYVUGIT LIBRARIES SMITHSONIAN INSTITUTION NOILALILSNI NVINOSHLINS S3IYVYUdIT LIBRARIES SM 

c ee Z Eee c z 

os = Es = E a = 

> = i> = = > = 

2 =. ae = - 2 Ee, 

m 2 m Zz Zz rm 2° 

n = w = = w = 
STITUTION NOILNLILSNI NVINOSHLINS S3Itu¥vudly LIBRARIES SMITHSONIAN INSTITUTION NOILOLILSNI NV 


AN 


NVINOSHLINS” S314YV4G17 LIBRARIES SM 


NVINOSHLINS S31uVvudl 


NVINOSHLINS 
SMITHSONIAN 
NVINOSHLINS 


SMITHSONIAN 
NVINOSHLIWS 


SMITHSONIAN 
TIT 


atMvuaiy LIBRARIES SMITHSONIAN INSTITUTION 


YY, f fa 
Y A Wee 4 


ISTITUTION NOILALILSNI NVINOSHLINS S3IYVYEIT LIBRARIES SMITHSONIAN 


ES 


NOILNLILSNI 
NOILNLILSNI 


LIBRARIES SMITHSONI 


LIBRARIES 
NOILNLILSNI 


NOILNLILSNI 


INSTITUTION NOIJLALILSNI NV 


S3INVUSIT. LIBRARI 


z = i = Zz 
S) Ka ° = = , = 
~ YS led 04 
= = = ke = 
= oe as “GHy, 
JIYVYE!IT_LIBRARIES SMITHSONIAN NOILALILSNI NVINOSHLINS SSIYVYEIT LIBRARIES SN 


Rey 


NVINOSHLINS S31YVUdIT_ LIBRARIES 


‘Lp 
‘Lp 
NVINOSHLINS S3IYVualT 


SMITHSONIAN 
NVINOSHLIWS 
SMITHSONIAN 
NVINOSHLIWS 
SMITHSONIAN 
iz 
4 Le 


SMITHSONIAN 
oy ey 4 
Vi 


.. _ y ; », 
ISTITUTION NOILALILSN! NVINOSHLINS S3IuUvYUdIT LIBRARIES SMITHSONIAN INSTITUTION NOILONLILSNI Nv 


o = rr) 7) 
uu a uw f ‘ Ww 
= | — Lap (RANG — 
z = E PY : ‘ 
a =| «x Ye: a 
a 5 ao i cs) 
a 2 3 i = 
J/¥vudi1 LIBRARIES INSTITUTION NOILALILSNI NVINOSHLINS SSlYVudit LIBRARIES Sh 


x 
SMITHSONIAN 


4 


8 


INSTITUTION NOILNLILSNI 


S3IINVUAIT LIBRARIES 
INSTITUTION NOILNLILSNI 


INSTITUTION NOILNLILSNI 


Sdiuvudl 
INSTITUTION 
Sadiuvaugit 


ISTITUTION NOILOLILSNI NVINOSHLINS S31YVYSIT_ LIBRARIES SMITHSONIAN INSTITUTION NOILALILSNI_ N\ 


NVINOSHLINS S31uVvudl 


22) Zz 2) P = za wo = 
= oy, = = = < = < 
‘ss SS . W Yi 2) w 2) 7) RAO AT) AD F 
oe DANS fe) Vif fis: = [e) ae a P S38 WS O i=: coe \i 
“ QF = 2 ES Saas Ss z = 
Ses 3 5 . pS 5 
J!uYvudil LIBRARIES SMITHSONIAN INSTITUTION —NVINOSHLINS Saluvuds LIBRARIES SN 
Zz e : 3 = z 
am >) & (Ee LON Wl Vi & KS LUN wi LSD B w STO a & JONG ay Ww 
LR 3 KEW = Coy = (RM) = (AM) 3 NQy = (A>) = AMD & 


iS © J k = SHaityN y 
Raby = ees D Wy Wf = Coe D > ree 
i o = ra) = 
STITUTION NOI TITSNUSNVINOSHIINS S43 luvaad rout BRARIES 
sg = = = Ke 
SS 5 = = ZNS 
~a Sz 5 z ; 2 SAS 
2 rs z =z NX, 
> j = > = 
Zz 7) 2 a > 
LIBRARIES SMITHSONIAN _ INSTITUTION NOILNLILSNI 
S o 2 to 
(22) a” — 
=! ec. =! c 
=) <x } SI < 
=I Ss s S 
fo) = fo) = 
z aj 3 a 
STITUTION _ NOILNLILSNI_NVINOSHLINS Sa luvud 7 tl BRARIES 
} aay fe S} = 
5 2 5 2 
E > E > 
5 = 5 = 
z o = o 
YVUAIT_ LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSNI 
; z ’ = z ® g Ay 
5 = 5 Yj. 
= Zz = Ze Vae 
= SX > = >” 
ow aati Dineee ds ee 
NOILNLILSNI NVINOSHLINS S3IYVYSIT LIBRARIES 
7) > 7) = 
us w uJ w 
= 4 = 4 
« = ce =! 
<x = _< = 
[aa = a =| 
m ro) es 3 
=5 Zz a) 2 
WIYvVudsiT LIBRARI ES SULIT USUI OA OMEN ES 
{ peo 
~ oO = (2) 
= w = 
2 = 2 = 
> Fe: > e 
z E a = 
D z o z 
NOILNLILSN! NVINOSHLINS SSIYVYEIT LIBRARIES 


eee 


HILIAS 
Up, 


NVINOSHLIINS 
SMITHSONIAN 
SMITHSONIAN 


uvugia LIBRARIES SMITHSONIAN INSTITUTION NOILALILSNI 


e 


RARIES 
INSTITUTION NOILALILSNI NVINOS 


STITUTION  NOILNLILSNI_NVINOSHLINS S3iuvudit LIBRARIES 


\ 
\ 


SMITHSONIAN 


SJIYVYUSIT LIBRARIES 


HuVvUaIT_LIBRARIES INSTITUTION NOILALILSNI 
“ b) @ 


NVINOSHLINS S3idvudII_LIB 


STITUTION NOILAL 


NVINOSHLIWS LIBRARIES 


ARIES SMITHSONIAN INSTITUTION NOJLOALILSNI 


ARIES SMITHSONIAN 
ILILSNI NVINOSHLINS 


ALILSNI 


} 
; 


SMITHSONIAN 


<h 
ie} 
a 

t 


0 
<niyist 


SS 


SMITHSONIAN 


INSTITU 
Salyv? 
INSTITL 


NOILALILSNI NV 


INSTITUTION 


SONIAN 


AASVAD 


NVINOSHLINS S3IYV! 


NVINOSHLIWS 


LIBRARIES SMI 


ES SMITH 


rs 
< 
a 
ro) 
Ww 
ac 
= 
= 
w 
at a — Pa, Y 
=| < al ay 
= ce Ss a % 
re) 2 fe) aaa 
2 Sage = a5 
SMITHSONIAN INSTITUTION NOILNLILSNI NVI 
z sy - S (= 
[@) = — — 
= wo = w 
E = E ss 
= = = = 
Be = i - 
z o z Oo 
NVINOSHLINS SSIYVYEIT LIBRARIES SMI 
z 7) ; Zam a) 
< = Ly < = 
z =] py, = = 
5 * yi S ; 
ae (2) SYA as (2) 
ir i Ee z 
5 z ze 
SMITHSONIAN INSTITUTION NOILN _NVI 
lJ wo uJ 7) 
z = z = 
< PS <x c 
Land 4 oa 4 
2 5 eo 3 
3 ae al =z 
NVINOSHLINS S3IYVYSIT LIBRARIES SM 
Ue z - = 
: : : : 
rad A =} a =). t 
> - > = ey 
2 = 2 i 
m me) = wn 
a) See (a) 4 
SMITHSONIAN INSTITUTION NOILALILSNI NVI 
w Zz x: (2) z 
= ae ee <= 
= z <a AE 
2 = g EX 
> = > C = 
2 7) pata g 77) 
_NVINOSHLINS Saluvudit LIBRARIES SMI 
2 St uJ = uJ 
4 AN = z a 
2 \N Ps S Ps G 
pa i co = ao 
S Oe z zi 
_ SMITHSONIAN INSTITUTION  NOILOLILSNI_NVI 
re) es , = fe) = ‘ 
5 ily, 2 5 2 
= Jala teas = 
- &, 0% e 2 
w* 7 m n m 
= 7) z ) 
NVINOSHLINS Saluvyugl oLIBRARIES, SMI 
= ats Si 
+ z a fF 
é 7s: BG 
g iS 2X 
>" = > 
Z wo 2 
SMITHSONIAN INSTITUTION _ NVI 
nf ty 2 0 NO ws = 
ss Yip 4 kK oD = a 
< oY. =i (& a) < = 


ia) 


ies 


ii 


OCS STUDY 
MMS 88-0021 


Population Status of California 
Sea Otters 


Population Status of California Sea Otters 
Pacific OCS Region 
Minerais Management Service 
U.S. Department of the Interior 


Contract No. 14-12-001-30033 


rey aD 
“i Br MADR V aye 


Ne 
ete 


Diy 
ih TURE 
ie i 


1 on 
nt it 
aT 
ean 


ti nt 
Hay 


Mi 
Orth 
eit! WM 


Population Status of California 


Sea Otters 


Edited by 


Dg Io Salbere 
Department of Ecology and Behavioral Biology 
University of Minnesota 
Minneapolis, MN 55455 


and 


K. Ralls 
National Zoological Park 
Smithsonian Institution 
Washington, D.C. 20008 


November 30, 1988 


This study was funded by the Pacific Outer Continental Shelf 
Region of Minerals Management Service, U.S. Department of the 


Interior, Los Angeles, California under Contract No. 14-12- 
001-30033 


JUN No 1693 > 


= niles 


<r" 


\ ghirnot htao to susase mokaetigen 
a1e2t0° 2ar 


wa beer cid 


gm Mee ess 


eee. theanbtiens sterue oes ties 
sive om PoneIned sehen pipmoes 


Disclaimer 


This report has been reviewed by the Minerals Management 
Service and approved for publication. Approval does not 
signify that the contents necessarily reflect the views and 
policies of the Bureau, nor does mention of trade names of 


commercial products constitute endorsement or recommendation 
for use. 


TABLE OF CONTENTS PAGE 


ACKNOWLEDGEMENTS Vv 
LIST OF TABLES AND FIGURES WALILSL 
ABSTRACT xix 
TECHNICAL SUMMARY OK 
APPENDICES 275 


CHAPTER 1. Overview of the study: 
background and general methods 
D. B. Siniff and K. Ralls a; 


CHAPTER 2. Reproduction, survival and tag loss 
in California sea otters 
D. B. Siniff and K. Ralls 13 


CHAPTER 3. Movement patterns and spatial use 
of California sea otters 
K. Ralls, T. Eagle, and D. B. Siniff 33 


CHAPTER 4. Time budgets and activity patterns of 
California sea otters 
Ke Ralls and D2 BB. sinver 64 


CHAPTER 5. Feeding patterns of California sea 
otters 
K. Ralls, B. Hatfield, and D. B. Siniff 38:4 


CHAPTER 6. Age determination of California sea 
otters from teeth 
P. Pietz, K. Ralls, and L. Ferm 106 


CHAPTER 7. Analysis of the precision and accuracy 
of radiotelemetry equipment and 
methods used in California 
A. Mercure 116 


CHAPTER 8. Movement patterns of adult female and 
weanling sea otters in Prince William 
Sound, Alaska 
C. Monnett and L. Rotterman 133 


CHAPTER 9. Sex-related patterns in the post-natal 
development and survival of sea otters 
in Prince William Sound, Alaska 
C. Monnett and L. Rotterman 162 


CHAPTER 10. A simulation model for assessing the 
risks of oil spills to the California 
sea otter population and an analysis of 
the historical growth of the population 
A. Brody 191 


ACKNOWLEDGEMENTS 


California Field Studies 


We are extremely grateful to the many people who 
contributed greatly to this research effort and made this 
report possible. Jack Ames, Robert Hardy, and Fred Wendell 
of the California Department of Fish and Game and James 
Bodkin, Brian Hatfield and Ron Jameson of the USFWS provided 
invaluable assistance by capturing otters and assisting during 
catching operations. Dr. Thomas D. Williams implanted all the 
transmitters, and was most patient and accommodating 
throughout many rather frustrating moments during field 
operations. Angela Doroff, Lisa Ferm, Brian Hatfield, Paul 
Henson, Christopher Jordon, Alan Mercure, Steve Osmek, Pamela 
Pietz, and Marian Skupski monitored otters for extended 
periods. Allan -Brody, Colleen Baggot, Ken Halama, Leslie 
Larson, Marianne Reidman and Christopher Logan also assisted 
in the field for short periods. Nancy Black, James Bodkin, 
Ron Jameson and Galen Rathbun of the USFWS and Marianne 
Reidman provided occasional reports on the location of our 
otters. Mary Faustini, Steve Mareck, Mike Henry and Stacy 
Kawa observed otters in Morro Bay. 


The following individuals gave us access to property 
along the coastline so that we could monitor our otters: Cc. 
Douglas Walling (Pfeiffer Point Mutual Water Company), Claire 
and Sybil Chappellet and their foreman Dennis Krackenberg 
(Rancho Rico), Evan Goldblatt (Big Creek Reserve), Doyle 
Danley and Wayne Titus (Cambria Radar Station), and Robert 
Smith (Diablo Canyon Nuclear Power Plant). 


Dick Rodgers, Morro Bay Harbormaster, and Brooks Bowhay, 
Monterey Harbormaster, gave us temporary docking facilities 
for our boat and Clyde Clark allowed us to store it 
temporarily in Morro Bay State Park. The McQueens of the Big 
Sur Campground allowed our small trailer to be a most 
important field camp in the Big Sur Area and also provided 
much logistic and moral support for our people in the field. 


Thomas Eagle wrote several helpful computer programs and 
assisted with data management. Lisa Ferm, Alan Mercure, and 
Marian Skupski completed most of the data entry. Dorothy 
Bromenshenkel and Rachel Ayetey worked magic with computer 
word processors, bringing this report through its many drafts 
until its final emergence into the form presented here. 
However, any errors that still exist, of course, rest with the 
editors. 


Robert L. Brownell, Jr. of the USFWS gave us access to 
facilities at the Piedras Blancas Lighthouse for our field 
base. Ron Jameson was most helpful throughout the project, 
as he provided valuable insights about field operations and 
sea otter biology from his extensive experience in California. 
Cedar Creek bioelectronics lab (L. Kuechle, D. Reichle, R. 
Schuster) provided telemetry equipment and advice about its 
use. Lynn Rathbun provided us with excellent drawings of our 
study area and a schematic drawing of the transmitter. 


Alaska Field Studies 


E. Birney, K. Ralls and D. Smith read the Alaska papers 
in this report and made many useful comments. Many thanks to 
A. DeGange, D. Garshelis, K. Schneider and T. Simon-Jackson, 
who provided access to various unpublished manuscripts. D. 
and J. Garshelis gave freely of their knowledge about the sea 
otters of Prince William Sound. The following individuals 
contributed during some phase of the field work: J. Bennett, 
D. Carlquist, A. DeGange, A. Doroff, F. Foode (pilot), P. 
Gullett (veterinarian), K. Hill (veterinarian), A. Johnson, 
F. Koecher (veterinarian), J. Nelson, P. Rosenberg (pilot), 
J. Ross, L. Rotterman, J. Sarvis, T. Simon-Jackson, M. 
Sorenson, D. Traun and T. Williams (veterinarian). 


For the Alaska studies, logistical support and/or 
equipment for field work were provided by the United States 
Fish and Wildlife Service (USFWS: M. Blenden, C. Dau, A. 
DeGange, A. Johnson, S. Lawrence, L. Pank, J. Sarvis, T. 
Simon-Jackson, D. Taylor), the Minerals Management Service 
(MMS: B. Hughes, G. Reetz, S. Treacy), the National Oceanic 
and Atmospheric Administration (NOAA: L. Jarvela, G. Lapine, 
M. Meyer), and the U.S. Forest Service (C. Nelson, K. 
Giezentanner, L. Keeler), the National Marine Fisheries 
Service, the Peter Pan Cannery (W. Bright), the Alaska 
Department of Fish and Game (H. Griese, J. Reynolds), and the 
University of Minnesota. The University of Alaska Marine 
Advisory Program (R. Steiner and G. Ference) provided office 
space and a contact point. 


Two Cordova businesses subsidized this project and 
contributed significantly to its successful completion. The 
Reluctant Fisherman Hotel (M. Johnson and R. Borer) generously 
provided lodging in Cordova at times when it was sorely 
needed. The Eyak Corporation (L. Borer) permitted us to 
establish a field camp on Native Alaskan lands. 


Model 

Dr. L. L. Eberhardt, Battelle Institute, Richland, 
Washington, developed the stochastic model for the females, 
derived the parameter estimates required for its operation and 


vi 


was invaluable in helping with the data analysis. Dr. Tony 
Starfield was also very helpful and stimulating in helping 
with, and formulating ideas about, the population model 
development. Jack Ames, Robert Hardy, and Fred Wendell of 
CDF&G gave us access to the raw data from past surveys for sea 
otters in California. CDF&G, USFWS, and the museums listed 
in Chapter 6 generously provided us with teeth for age 
determination. 


Finally we acknowledge the patience, confidence and 
general support shown to us by Gordon Reetz, our Contractor 
Officer's Technical Representative. Throughout the project 
we were faced with delays and difficult decisions as to how 
to proceed. His support at these times was greatly 
appreciated. 


LIST OF TABLES AND FIGURES PAGE 


TABLE 2.1 -- Summary of sea otters captured 15 
in California during 1984 and 1985. 


TABLE 2.2 -- A list of sea otters that were 
instrumented with implanted radio transmitters. 17 


TABLE 2.3 -- Reproductive and age data, and 
length of monitoring period for adult 
female sea otters. 21 


TABLE 2.4 -- Reproductive information on 
eight adult female sea otters and data 
on 6 otters that were used to calculate 
the inter-birth interval. 24 


TABLE 2.5 -- Annual survival rate estimates 
for the four sex/age categories of adult 
females, adult males, juvenile females and 
juvenile males. 25 


TABLE 2.6 -- Estimates of annual tag survival 
rates based upon the method of Heisey and 
Fuller (1985). 26 


TABLE 2.7 -- A comparison of the age of pups at 
separation from the female in California and 
Alaska. 27 


TABLE 2.8 -- Numbers of male and female carcasses 
in good condition compared to those expected 
if the sex ratio was equal. 28 


TABLE 3.1 == Average distance (km) between 
successive locations, recorded between 18 
and 36 hours apart for each instrumented otter. 37 


TABLE 3.2 -- Average distance (km) between successive 
locations, recorded more than 36 hours apart, 
for each instrumented otter. 38 


TABLE 3.3 - Area (ha) of daily home ranges based on 
data obtained during 24-hour watches. 47 


TABLE 3.4 - Average monthly distance deviations (km) 
from the harmonic mean center for instrumented 
California sea otters. 53 

TABLE 3.5 - The average location along the five 
fathom line for each instrumented sea otter. 55 


viii 


TABLE 3.6 - The average distance (km) between 
extreme locations. for four sex and age 
categories of California sea otters. 


TABLE 3.7 - Comparison of home range areas (ha) 
for sea otters in California and Alaska. 


TABLE 4.1 - A comparison between activity data 
obtained visually with that obtained using the 
quality of the telemetric signal. 


TABLE 4.2 - A comparison between time budgets 
calculated from observing activity visually 
and judging activity using the quality of the 
telemetric signal. 


TABLE 4.3 - A comparison among the methods of 
calculating time budgets. 


TABLE 4.4 - Analysis of variance testing for 
differences in percent of time spent feeding 
between the various sex/age classes. 


TABLE 4.5 - A comparison of time budgets between 
females with small pups and those with large 


pups. 


TABLE 4.6 - A comparison of the percent of time 
spent in each of the three activity categories, 
resting, feeding and other, for activity data 
over the entire 24-hour period. 


TABLE 4.7 - A comparison of the activity budgets 
for sea otters calculated in this study and 
those in the literature. 


TABLE 5.1 - A comparison among the average dive 
times (sec) for feeding sea otters for various 
type of prey. 


TABLE 5.2 - The average length of the feeding dives 
(sec) for sea otters for prey of different 
sizes. 


TABLE 5.3 - A comparison among the surface times 
(sec) required to consume the various prey 
items. 


TABLE 5.4 - A comparison among the average surface 
times (sec) required to consume various size 
prey. 


TABLE 5.5 - A comparison of the percent of successful 


ix 


56 


59 


73 


74 


76 


76 


UN 


77 


79 


88 


88 


89 


89 


feeding dives for various sex/age categories. 90 


TABLE 5.6 - A comparison of the average dive times 
(sec) for sea otters during feeding bouts. 91 


TABLE 5.7 - Analysis of variance testing for 
differences in the length of the dives made 
by otters belonging to the various age/sex 
classes. 94 


TABLE 5.8 - The average surface times (sec) for 
individual sea otters during feeding bouts. 95 


TABLE 5.9 - A comparison among the average surface 
times (sec) for five sex/age categories that 
were recorded during feeding bouts. 96 


TABLE 5.10 - The average lengths of the feeding bouts 
(min) for individual sea otters. 97 


TABLE 5.11 - The lengths of feeding bouts (min), 
grouped by time intervals, for five sex/age 
categories. 98 


TABLE 5.12 - The average lengths of intervals between 
feeding bouts (min) for individual sea otters. 99 


TABLE 5.13 - The frequency of the lengths of the 
intervals between feeding bouts (min), grouped by 
time intervals for five sex/age categories. 100 


TABLE 5.14 - A comparison of day and night dive 
lengths (sec) for the individual sea otters, 
for five age/sex classes. 101 


TABLE 5.15 - A comparison of the mean lengths of 
surface intervals made during the day and 
night by the individual instrumented otters 
in the various age/sex classes. 102 


TABLE 6.1 - Tooth age estimates for sea otters of 
minimum known age. 108 


TABLE 6.2 - A comparison of animal ages determined 
from tooth cementum by Gary Matson: (1) with 
animals of known age; (2) with duplicate 
determinations based on a different tooth from 
the same animal; and (3) with ages estimated 
from the degree of tooth wear (reprinted from 
Matson's Tooth Cementum Age Analysis, Progress 
Report No. 9, Spring 1987, Table 1). LL? 


TABLE 6.3 - Comparisons of sea otter age assignments 
based on counts of incremental lines in tooth 
cementum. 113 


TABLE 7.1 - Summary statistics for the 897 hand-held 
compass bearings to radio transmitters on buoys 
off the California coast. 127 


TABLE 7.2 - Summary of the calculated precision and 
accuracy of locations of radio transmitters. 128 


TABLE 7.3 - Comparisons between the field method of 
plotting data and the Andrews estimator for the 
same set of data. 129 


TABLE 9.1 - Comparison of growth rates for large vs. 
small pups. 170 


TABLE 9.2 - Error is estimation of birth dates from 
growth rate assumptions. 172 


TABLE 9.3 - Sex of dependent Prince William Sound 
sea otter pups. 172 


TABLE 9.4 - Survival rates of sea otter pups in 
Prince William Sound. 182 


TABLE 10.1 - Estimates of annual survival rates of 
telemetered sea otters in California, as 
determined by the method of Heisey and Fuller 
(1985), 1983-1986. 205 


TABLE 10.2 - Default parameters used in OTPOP and 
- LESLIE. 212 


TABLE 10.3 - CDFG and USFWS censuses used in analysis 
of California sea otter distribution. 225 


TABLE 10.4 - Analysis of variance in CDFG and USFWS 
California sea otter census data, 1968-1985. 
Dependent variable is the proportion of census 
total along a 10 km section of coast. 226 


TABLE 10.5 - California sea otter sex ratios of 
recovered carcasses (1968-1985) and as 
subjectively estimated by field biologists, by 
season and CDFG carcass recovery area. 232 


TABLE 10.6 - Parameters used in short-term otter 
movement model. AR and CE are regression 
parameters discussed in text. The symbol sd 
is standard deviation of regression errors, 


xi 


R* is given for the regressions. Vmax is mean 
maximum daily movement, derivation discussed 
Ty Cext 246 


TABLE 10.7 - Parameters giving the best fit of 
OTRANGE to historical data. See text for 


explanation of parameters. 267 
FIGURE 1.1 - A map of the study area in California. 4 
FIGURE 1.2 - A map of the study area in Alaska. 5 


FIGURE 1.3 - Schematic drawing of the upper jaw 
of a sea otter showing vestigial first pre- 
molar that was sometime removed from captured 
animals and sectioned for determination of 
age. 7 


FIGURE 1.4 - A schematic drawing of the implanted 
radio transmitter used during this study. 9 


FIGURE 2.1 - The percent of the adult females with 
pups during each month of the year. 22 


FIGURE 2.2 - A comparison of tag-loss rates between 
this study and the study of Ames, et al., 1983. 30 


FIGURE 3.1 - A comparison of the average distance 
between successive locations for four age/sex 
categories, for locations made 18-36 hours apart 
and those made more than 36 hours apart. 39 


FIGURE 3.2 - A plot of the 20 longest trips made 
between successive locations that were 18-36 
hours apart for four age/sex categories. 41 


FIGURE 3.3 - The average location of each instrumented 
sea otter along the California Coast. 42 


FIGURE 3.4 - The general north-south movement pattern 
of individual adult males. 43 


FIGURE 3.5 - The general north-south movement pattern 
of individual adult females. 44 


FIGURE 3.6 - The general north-south movement pattern 
of individual juvenile females. 45 


FIGURE 3.7 - The general north-south movement pattern 
of individual juvenile males. 46 


FIGURE 3.8 - The distribution of distances offshore 
for four age/sex categories of California sea 
otters. 


49 
FIGURE 3.9 - The average distance offshore while 


resting and feeding for individual otters for five 
age/sex categories. 50 

FIGURE 3.10 - The average distance offshore by hour 
of the day for juvenile males and females. 51 

FIGURE 3.11 - The average distance deviation from the 


harmonic mean center of monthly home ranges for 
four age/sex categories. 


52 
FIGURE 3.12 - The distance between extreme locations 


for instrumented sea otters in California. 54 
FIGURE 4.1 - The locations along the California coast 
of watches for collecting time budget data on sea 
otters instrumented with radio transmitters. 68 
FIGURE 4.2 - The percent of time that adult male sea 
otters spent in various activities at the various 
hours of the day. 70 
FIGURE 4.3 - The percent of time that juvenile sea 
otters (male and female spent resting, 


feeding and 
in other activity for various hours of the day. 


71 
FIGURE 4.4 - The percent of time that adult female sea 
otters (with and without pups) spent resting, 


feeding and in other activity for various times 
of the day. 


UZ 
FIGURE 5.1 - The distribution of dive times during 


feeding bouts for five groups of sea otters in 
California. 


92 
FIGURE 5.2 - The distribution of the length of time 


of the surface intervals during feeding bouts 
for five groups of sea otters in California. 


93 
FIGURE 6.1 - Comparison of age estimates based on 


teeth to age estimates using skull features. 


110 
FIGURE 6.2 - Distribution of age estimates based on 


incremental lines in tooth cementum. 


114 
FIGURE 7.1 - Illustration of the method used to 


determine the location of a buoy by taking 
compass bearings. 


118 


xiii 


FIGURE 7.2 - Distribution of compass bearings from 
sightings through a telescope (Questar). 123 


FIGURE 7.3 - Distribution of hand-held compass 
bearings to a prominent landmark. 124 


FIGURE 7.4 - Compass bearings to a radio transmitter 
on a buoy off the California coast. SLABS 


FIGURE 7.5 - Distribution of hand-held compass 
bearings to the signals from radio transmitters 
off the California coast. 126 


FIGURE 8.1 - Study area in Prince William Sound, 
Alaska, 1984-1987. 136 


FIGURE 8.2 - Distances between extreme locations of 
eight adult female sea otters in eastern Prince 
William Sound, Alaska. The number of fixes and 
total monitoring intervals are given: 

# fixes / # days. 139 


FIGURE 8.3 - Movements of an adult female sea otter in 
Prince William Sound, Alaska, during a 20 month 
interval, June 1984 - February 1986. Summers 
were spent in the western portion of the study 
area and winters in the eastern portion, near 
the Cordova male area. 140 


FIGURE 8.4 - Division of study area in Prince 
William Sound, Alaska, into numerically 
designated habitat zones and superzones. Zones 
correspond to major bays or passages. 141 


FIGURE 8.5 - Use of habitat zones in Prince William 
Sound, Alaska, by eight radio-instrumented adult 
female sea otters. 142 


FIGURE 8.6 - Seasonal changes in the use of portions 
of eastern Prince William Sound by eight radio- 
instrumented adult female sea otters. Superzones 
are delineated on Figure 4. 143 


FIGURE 8.7 - Distances between extreme locations of 
26 female sea otters in Prince William Sound, 
Alaska, that were accompanied by dependent pups. 
Most observations are based on females 
accompanying radio-instrumented pups. 144 


FIGURE 8.8 - Changes in the home ranges of sea otter 
female-pup pairs in Prince William Sound, Alaska, 


Xiv 


that occur as the pups approach weaning age. The 
distances between extreme locations of pairs are 
compared for the last 30 days before weaning and 
for the earlier period when the pup was younger. 147 


FIGURE 8.9 - Distances traveled from the site of 


weaning in Prince William Sound, Alaska, by male 
and female weanling sea otters. Monitoring 

interval varied from a few days to approximately 
18 months. Short monitoring intervals resulted 


when pups died during their travels. 


FIGURE 8.10 - Distance between weaning location of 
sea otters location in Prince William Sound, 


Alaska, and their first post-weaning home range. 


The distance was traveled in a single 
relatively rapid trip. 


FIGURE 8.11 - Relative size of weanling male and 
female sea otter home ranges in Prince William 


Sound, Alaska, during the first winter following 


weaning. Only weanlings with well defined home 
ranges are included. 


FIGURE 8.12 - Tendency for weanling sea otters in 


Prince William Sound, Alaska, to leave the natal 


female area after being weaned. Female 
weanlings usually do not leave the natal female 
area, whereas males usually do. Female area 
consists OF, ZONES alte e2),. 1S rate: Ole 7). Sn and: tt 
on Figure 4. 


FIGURE 9.1 - Location of study area in Alaska and 
in Prince William Sound. 


FIGURE 9.2 - Growth rates of dependent male and 
female sea otter pups. 


FIGURE 9.3 - Estimated birth dates and capture 


dates of sea otter pups in Prince William Sound, 


Alaska. 


FIGURE 9.4 - Weaning dates of instrumented sea 
otter pups in Prince William Sound, Alaska, 
1985-1986. 


FIGURE 9.5 - Chronology of dependency periods of 27 
sea otter pups in Prince William Sound, Alaska, 
1984-1986. 


FIGURE 9.6 - Sea otter pups instrumented in Prince 
William Sound, Alaska, that died or with which 


xv 


149 


151 


152 


153 


165 


173 


174 


175 


176 


radio contact was lost. 179 


FIGURE 10.1 - Schematic representation of the 
interrelation of the submodels used to predict 
the potential effects of oil spills on California 
sea otter population dynamics. 195 


FIGURE 10.2 - Hypothetical survivorship curve 
depicting the relationships of the 3 competing 
risks of Siler (1979) and Eberhardt (1985). 200 


FIGURE 10.3 - Hypothetical reproductive curve 
depicting the relationship between prime 
reproductive rate and senescence, after 
Eberhardt (1985). 202 


FIGURE 10.4 - The effect of the value of b on the 
non-linearity of the density dependence function 
used in OTPOP and LESLIE. K is the carrying 
capacity. 203 


FIGURE 10.5 - Distribution of ages of California sea 
otters estimated by tooth cementum. 207 


FIGURE 10.6 - Relative average number of small pups 
and large pups, by month, in the CDFG index 
areas, 1977-1984. 209 


FIGURE 10.7 - Age-specific female survivorship curves 
and annual survival rates at different per 
capita growth rates. 214 


FIGURE 10.8 - Age-specific reproductive rates under 
the default population parameters used in 
the population model. 215 


FIGURE 10.9 - Age-specific female California sea 
otter annual survival rates calculated from 
certain model parameters. 216 


FIGURE 10.10 - Age-specific male California sea otter 
annual survival rates calculated from certain 
model parameters. 217 
FIGURE 10.11 - Male and female California sea 
otter survivorship curves from certain 
population parameters. 218 


FIGURE 10.12 - Pattern of monthly pup abundance 
obtained by simulation. 221 


FIGURE 10.13 - Contour diagram indicating annual 


xvi 


changes in sea otter density in California 
from 1968-1985. 227 


FIGURE 10.14 - Contour diagram indicating monthly 
changes in sea otter density in California from 
1968-1985. 228 


FIGURE 10.15 - Density functions used to obtain 
locations of independent sea otters in 
California in June (dashed line) and December 
(solid line). 230 


FIGURE 10.16 - Local proportion of California sea 
otters that are female in June (dashed line) and 
December (solid line), used in the population 
models. BSL 


FIGURE 10.17a - Density functions used to obtain the 
location of male (dashed line) and female 
(solid line) sea otters in California in June. 234 


FIGURE 10.17b - Density functions used to obtain the 
location of male (dashed line) and female 
(solid line) sea otters in California in 
December. 235 


FIGURE 10.18 - Schematic representation of the 
methods to predict sea otter densities in 
expanded range. 237 


FIGURE 10.19a - Daily locations of a juvenile 
female California sea otter as determined by 
telemetry, 1985-1986. 242 


FIGURE 10.19b - Daily locations of a juvenile male 
California sea otter as determined by telemetry, 
1985-1986. 243 


FIGURE 10.20a - Daily locations of a juvenile female 
California sea otter, as in Fig. 10.19a. 244 


FIGURE 10.20b - Daily locations of a juvenile male 
California sea otter, as in Fig. 10.19b. 245 


FIGURE 10.21a - Simulated movements of sea otters 
around an oil spill in Monterey Bay beginning 
1 December and lasting 15 days. 249 


FIGURE 10.21b - Simulated movements of sea otters 
around an oil spill in Monterey Bay beginning 
1 December and lasting 15 days, as in 
Fig. 10.21la, except that in this simulation 


xvii 


movement parameters were different than in 
aleyy UO GAINS i 250 


FIGURE 10.21c - Simulated movements of sea otters 
around an oil spill in Monterey Bay beginning 
1 December and lasting 15 days, as in Fig. 
10.21la, with different parameter values. Pysyal 


FIGURE 10.22 - Age-specific reproductive values of 
female California sea otters for default 
parameter settings. 252 


FIGURE 10.23a and 10.23b - Printed output from a run 
of the model introducing a large oil spill for 
10 days along a 50km section of coast between 
Marina and Yankee Point. 254-7 


FIGURE 10.23c - Trace of the total simulated 
population size for runs in Fig. 10.23a. 258 


FIGURE 10.23d - Trace of the total simulated 
population size for the control (no oil spill) 
runs in Fig. 10.23a. 259 


FIGURE 10.23e - Trace comparing the mean values from 
Figs. 10.23c and 10.23d. 260 


FIGURE 10-23£ - Mean cumulative number of otter 
deaths due to oiling for the run in Fig. 10.23a.261 


FIGURE 10.24 - Schematic representation of the computer 
program OTRANGE. 264 


FIGURE 10.25 - Density dependence functions used in 
the computer program OTRANGE. 265 


FIGURE 10.26 - Fit of computer model to historical 
data using the "best estimate" parameters 
without density independent mortality. 268 


FIGURE 10.27 - Fit of computer model output to 
historical data using the "best estimate" 
parameters incorporating density independent 
mortality. Solid line traces population size, 
dashed line traces carrying capacity. 269 


Xviii 


ABSTRACT 


The main objective of the contract was to develop a 
simulation model to facilitate analysis of the risk of oil 
spills to the threatened California sea otter population. 
Existing data on the dynamics and demography of the population 
were reviewed and synthesized. The additional data needed for 
model development were collected through radiotelemetry 
studies of sea otters in Alaska and California. 


Our field work indicated that the California population 
had a high reproductive rate but many pups did not survive to 
weaning. /Adult females had the highest survival rates and 
adult males the lowest. Juvenile females had lower survival 
rates than adult females and spent more time foraging than 
other otters. Otters tended to stay within a small area for 
an extended period and then suddenly move for a considerable 
distance. They made more long-distance movements than 
expected. Juvenile males tended to travel more extensively 
and range farther offshore than other otters. 


The simulation model contains four interrelated 
stochastic submodels: a short-term population model, a long- 
term population model, a sea otter distribution model, and a 
sea otter movement model. This report includes a detailed 
description of the model, the data on which it is based, and 
an operating manual. The computer program for the model has 
also been provided to MMS. 


TECHNICAL SUMMARY 
Chapter 1. 


The objectives of the contract were to review and 
synthesize existing information on the dynamics and demography 
of the threatened California sea otter population, to design 
and conduct field studies to collect the data needed to fill 
data gaps identified through this process, and to develop a 
Simulation model to facilitate analysis of the risk of oil 
spills to this population. 


We computerized past survey data collected by the 
California Department of Fish and Game and the U.S. Fish and 
Wildlife Service and aged teeth from the salvaged otters found 
dead along the California beaches over the last 20 years. 


The additional data needed for model development were 
collected through radiotelemetry studies of sea otters in 
Alaska and California. Because of the sensitive nature of 
hands-on field work on the threatened California population, 
we tested procedures and equipment in Alaska before applying 
for permits to use them in California. We developed and used 
a radio transmitter that could be implanted within the 
abdominal cavity of sea otters. The use of these transmitters 
enabled us to make a number of new discoveries about sea 
otters in both Alaska and California. 


Chapter 2. 


We observed 40 California sea otters, representing all 
four major age/sex groups, that were flipper-tagged and 
instrumented with implanted radio transmitters. 


The proportion of females accompanied by a pup peaked in 
the spring, with a secondary peak in the fall. Two methods 
of estimating the annual reproductive rate gave comparable 
values of 0.88 and 0.90 pups per adult female. The average 
inter-birth interval was 416 days. Eight of the 19 pups born 
did not survive to weaning. 


Among the four major age/sex classes, adult females had 
the highest estimated survival rates and the adult males the 
lowest. Juvenile females had lower survival rates than adult 
females but juvenile males had higher survival rates than 
adult males. 


The estimated annual loss rate for the flipper-tags was 
0.26. More individuals lost two tags than would be expected 
by chance. It is unlikely that accurate estimates of sea 
otter survival rates can be derived from observations of 
tagged individuals. 


XX 


Chapter 3. 


We obtained a detailed picture of sea otter movement 
patterns in California by attempting to locate each 
instrumented otter, by radiotelemetry, on a daily basis. [In 
general, otters tended to stay within a small area (1-2 km of 
shoreline) for an extended period and then suddenly move for 
a much longer distance. Our daily monitoring revealed that / 
individual otters of all age/sex classes make a surprising | 
number of long-distance movements at all times of year. There 
was substantial variation in movement patterns among 
individuals within all age/sex classes but there were also 
Significant differences between classes. Juvenile males were 
the most extensive travelers and also ranged farther offshore 
‘than otters of the other age/sex classes. 


Chapter 4. 


Radiotelemetry is particularly useful for collecting time 
~budget and activity data on sea otters because radio signals 
are not transmitted through sea water. Three general 
categories of activity can be distinguished by listening to 
the radio signal from an otter: resting, feeding, and 
"other". Otters of all age/sex classes tended to be active 
and feed for a large proportion of the time during the late 
afternoon and early evening but there were differences in the 
activity patterns of the various groups. Juvenile females 
and adult females with pups spent more time foraging than 
other otters. Differences in the ability of members of 
different age/sex classes to compete for food resources are 
common in vertebrates. [In the California sea otter 
population, the juvenile females spent almost half of their 
time foraging, suggesting that they are poor competitors for 
food. 


Chapter 5. 


Although we collected some data on uninstrumented otters, 
we focused on the foraging patterns of individual instrumented 
sea otters as indicated by radio-telemetry. Our telemetry 
data indicated that visual observations of otter foraging 

~patterns tend to underestimate mean dive lengths. There was 
a striking degree of individual variation in foraging 
patterns, Many individuals displayed differences in diurnal 
and nocturnal dive-length patterns that may reflect a tendency 
to specialize on different prey species by day and night. 
However, there was no general tendency for longer dive lengths 
or surface intervals during the day or night. Juvenile males 
often fed far from shore where they could not be seen. 
Juvenile females had longer feeding bouts than otters of the 
other age/sex classes. 


XxXi 


Chapter 6. 


In an effort to gain insight into the age structure of 
the California population, we studied a sample of premolars 
from more than 580 dead sea otters salvaged from beaches. We 
counted bands in the cementum of the sectioned teeth to 
estimate age. We were able to examine teeth from ten otters 
of known minimum age and the age estimates based on these 
teeth compared quite favorably with those made by field 
biologists. Age estimates based on teeth also compared well 
with those based on skull features. Teeth that had been 
boiled were more difficult to interpret than those that had 
not been boiled. There was excellent agreement between 
successive age estimates by the same reader and good agreement 
across readers. 


Chapter 7. 


We evaluated the accuracy and precision of the 
radiotelemetry methods we used to locate otters in California 
with a radio transmitter on a buoy anchored off the coast. 
We established the location of the buoy with visual methods 
and took a series of compass bearings on the buoy's radio 
signal. Signal bounce was not a significant problem. The 
accuracy of our bearings compared quite favorably with that 
of those taken in other radio-telemetry studies. For otters 
located within about 800 meters from shore, precision was 
estimated at 0.03 to 0.06 hectares and accuracy at 51 to 110 
meters. The results obtained by hand-plotting points, which 
was our usual field procedure, compared well with those 
obtained with the Andrews estimator calculated by the computer 
program TRIANG. 


Chapter 8. 


This chapter focuses on relatively long-term movement 
patterns of adult female and juvenile sea otters in Alaska. 
Adult females were much more mobile than had previously been 
suspected but their movements were greatly reduced in the 
month before weaning. Male weanlings left the area in which 
they were born shortly after weaning, so that spatial 
segregation of the sexes occurred at a very young age. Sea 
otters used different portions of the available habitat for 
different purposes, such as for weaning pups and over- 
wintering. Hence, movements and habitat use varied 
seasonally. 


Chapter 9. 


Birth dates, growth rates, dependency periods, weanling 
behavior and survival of male and female otter pups in Alaska 


Xxii 


were compared. Many of these factors did not vary between 
males and females. However, dependent male pups grew more 
rapidly than dependent females and weaned females had lower 
survival rates than weaned males. 


Chapter 10. 


A stochastic simulation model of California sea otter 
population dynamics was constructed to be used in the analysis 
of the risk of oil spills to the legally threatened 
population. The model consists of four submodels: iL) 
population model that iterates on a monthly basis; 2) a 
population model that iterates on a yearly basis; 3) a 
spatially explicit population distribution model; and 4) a 
sea otter movement model. Simulated population dynamics are 
density-dependent but the model has the flexibility to allow 
investigation of density-independent reproduction and 
mortality. The monthly population submodel operates for four 
Simulated years before the simulated oil spill. At the time 
of the spill, individual animals are distributed along the 
coast by the distribution submodel. In the movement submodel, 
individual animals then either avoid or are killed by the 
spill. Population recovery can be simulated for up to 50 
years after the spill using the monthly and annual population 
models. 


Age and sex specific survival and reproductive rates are 


the core of the population submodel. These rates are 
estimated using telemetry and other data in a "competing 
risks" theoretical framework. Data from the semi-annual 


censuses of the population conducted by the California 
Department of Fish and Game and the U.S. Fish and Wildlife 
Service are incorporated in the distribution model. [In the 
movement submodel, daily movements are modeled with regression 
equations, using parameters estimated from the radiotelemetry 
data on the California animals. 


Sensitivity analysis of the population model indicated 
that the recovery time after a spill depends on the percentage 
of the female population killed, the status of the population 
in relation to its carrying capacity at the time of the spill, 
and the amount of environmental stochasticity in annual 
survival rates. 


Xxiii 


CHAPTER 1 


OVERVIEW OF THE STUDY: BACKGROUND AND GENERAL METHODS 


D. B. SINIFF AND K. RALLS 


November 30, 1988 


BACKGROUND 


The outer continental shelf of the Pacific coast is 
believed to contain extensive oil and gas reserves. The Santa 
Maria and Santa Cruz Basins, off the coast of central 
California, are potentially some of the most active areas of 
oil exploration and development. Areas to the south of Point 
Conception have already been developed into productive fields, 
and more will be developed in the future. The Minerals 
Management Service (MMS), U.S. Department of Interior, is the 
federal agency responsible for administering leases of 
submerged federal lands. Amendments to the Outer Continental 
Shelf Lands Act of 1953 set MMS objectives for managing 
development of outer continental shelf lands, including 
protection of human, marine, and coastal environments. 


A first step in making decisions about leasing, 
exploration, and development that protects the marine and 
coastal environment is risk analysis. MMS has directed and 
funded a number of studies of the risk of off-shore oil 
development, particularly of 011 spills resulting from leasing 
activities, to wildlife populations (e.g. Ford 1985, Reed, et 
al., 1986). 


One of the most sensitive wildlife species, from both 
political and biological perspectives, that could be impacted 
by an accidental spill development in the Santa Maria and 
Santa Cruz Basins is the California sea otter (Enhydra 
lutris). Commercial exploitation during the 18th and 19th 
centuries reduced the aboriginal population of perhaps 20,000 
otters along the California coast to probably less than 100 
in 1911 (USFWS 1986). Protection provided by international 
treaty and federal and state legislation allowed the 
population to recover, at a rate of about 5% per year, to its 
present size of approximately 1500 animals (Ralls, et al., 
1983). In 1976 the southern sea otter was officially listed 
as "threatened" under the Endangered Species Act; a major 
reason for the designation was the potential risk of oil 
spills to the small and geographically isolated population. 
To obtain information on the southern sea otter, MMS issued 
a request for proposal in late April 1983. It outlined a 
series of objectives for studies on "The Population Status of 
California Sea Otters". Three objectives were central to this 
request for proposal: 1) to consider the existing information 
on the dynamics and demography of the California sea otter 
population and determine what additional information would be 
necessary to predict the effects of oil spills, of various 
sizes in different parts of the sea otter range in California 
(Fig. 1.1), on the population; 2) to design and conduct 
studies needed to fill the identified data gaps; and 3) to 
develop a population model that would help to determine the 
way in which the size and productivity of the population would 


2 


be likely to be affected by oil spills in different parts of 
the range. 


In October, 1983, we received the contract to carry out 
this work and immediately started to obtain the necessary 
federal and state permits to conduct field studies of sea 
otters in California. We proposed the use of implanted radio 
transmitters to monitor otters along the California coast. 
In March, 1984, we received the necessary permits to implant 
five otters and began field work in California. We 
subsequently obtained additional permits and have implanted 
a total of 40 sea otters in California and monitored them 
through late December, 1987. In this report, we analyze the 
data collected on these individuals and the results are 
presented in Chapters 2 through 5. 


We collected teeth from the salvaged otters found dead 
along the California coast over the last 20 years. The teeth 
were sectioned to allow estimation of the ages of individuals. 
This information was used in the development of the population 
model. The methods used and the results of the tooth analysis 
is presented in Chapter 6. In the development of the model, 
the data from the field studies were used primarily for 
establishing a basis for movement patterns along the coast and 
estimates of reproduction and survival rates. Since the 
precision and accuracy of the telemetry positions along the 
California coast were important in our analyses, an evaluation 
was carried out using transmitters at fixed locations that 
simulated floating otters. This evaluation is presented in 
Chapter 7. 


Because the California sea otter population is classified 
as threatened under the Endangered Species Act, we tested 
procedures and equipment on sea otters in Prince William 
Sound, Alaska (Fig. 1.2), before using them in California. 
Data on sea otters in Alaska were collected while equipment 
and procedures were being developed. These data resulted in 
significant new findings, particularly with respect to events 
during the period of pup dependency. They are presented in 
two chapters of this report, "Movement patterns of adult 
female and weanling sea otters in Prince William Sound, 
Alaska" (Chapter 8) and "Sex-related patterns in the post- 
natal development and survival of sea otters in Prince William 
Sound, Alaska" (Chapter 9). 


The final chapter of this report describes the population 
model that we developed. A significant body of data from 
diverse sources has been integrated into this effort. These 
data have come from our current field studies as well as the 


FIGURE 1.1 - A map of the study area in California showing the 
approximate range of sea otters during the period of the study 
in 1984, 1985, and 1986. The portion of the range in which 
reproduction occurs is indicated by cross-hatching; the 
northern and southern areas occupied mostly by males are 
indicated by diagonal lines. 


San Nicolas @ 


FIGURE 1.2 - A map of the study area in Alaska showing the 
general range of the sea otters that were studied during 1984 
and 1985. 


“eee a” PRINCE WILLIAM 
SOUND 


[i] stuoy AREA 


Ati YW ALDEZ 
iQ mu 


—zZ—> 


AR Se “gl 
BG 


‘ i 
i jut 


: iat 


ng! ior 


“2 7 


i 


oi aT iy | ig a ae 


past efforts of both the California Department of Fish and 
Game and the U.S. Fish and Wildlife Service. We organized all 
these past survey data into a data base and used it 
extensively in the development of the model. 


GENERAL METHODS 


Capture and release 


We captured otters in three ways: with floating gill- 
nets, the Wilson trap developed by the California Department 
of Fish and Game (Ames, et al., 1983), and dip-nets, which 
have been used extensively by the U.S. Fish and Wildlife 
Service in California. Floating gill-nets were used mostly 
in Alaska. We used them initially in California but wind and 
fog made it difficult to check the nets frequently. The 
majority of the adult otters in California were captured with 
the Wilson trap and most of the juveniles with dip-nets. We 
also used dip-nets to capture pups and newly independent young 
in Alaska. All animals were released near the point of 
capture. 


Teeth 


Sea otters have a small, vestigial premolar directly 
behind each canine (Fig. 1.3). Schneider (1973) developed an 
aging technique based upon the number of cementum layers found 
in stained sections and Garshelis (1984) applied this 
technique to teeth extracted from anesthetized sea otters in 
Alaska. We collected this premolar from many of the 
California animals. However, our sample was incomplete due 
to lack of permission to extract a tooth during the early 
portion of the study and some breakage during extraction of 
the teeth. 


Transmitters 


The transmitters were developed by the University of 
Minnesota's Cedar Creek Bioelectronics Laboratory. The first 
models, used in Alaska in 1982 before this project began, 
measured about 6.8 x 4.8 x 1.8 cm and weighed about 70 g. 
Although it had been shown that implanting transmitters in 
the abdominal cavity had no deleterious effects in other 
species (Smith, 1980; Eagle, et al., 1984), no information on 
this point was available for sea otters. Therefore we set out 
to compare the results of implanting transmitters beneath the 
skin (subcutaneous) and within the abdominal cavity 


FIGURE 1.3 -- Schematic drawing of the upper jaw of a sea 
otter showing the vestigial first premolar that was sometimes 
removed from captured animals and sectioned for determination 
of age. 


VESTIGIAL FIRST PREMOLAR 


(intraperitoneal). Because some of the 1982 implants were to 
be subcutaneous, we used small, flat batteries on which we had 
no previous performance data. Neither method of implantation 
appeared to have significant deleterious effects on the 
otters; however, the subcutaneous implantation procedure left 
a noticeable lump so we decided to use the intraperitoneal 
method in the future. Unfortunately, the small batteries 
proved to be unreliable and most of them failed within four 
months after the transmitters were implanted (Garshelis and 
Sime, IGS) 


We then designed a new transmitter using lithium 
batteries developed by the Medtronic Corporation. These 
batteries were developed for use in medical devices implanted 
in humans and were known to be extremely reliable in these 
applications. However, at the request of the U.S. Fish and 
Wildlife Permit Office, these transmitters were subjected to 
four months of extensive testing prior to use in California. 
They were operated under simulated physiological conditions 
for this entire period, except when undergoing tests at 
extreme temperature and pressure conditions, and successfully 
passed all tests. 


These transmitters (Figure 1.4) measured about 7.6 x 5 
x 2.5 cm and weighed about 120 g in air. This weight ranges 
from about 1.8 percent (in an 18-lb juvenile) to about 0.4 
percent of sea otter body weight (in a 70-lb adult male), thus 
these transmitters were smaller, relative to body weight, than 
those used successfully in other species of mammals. For 
example, Eagle, et al., (1984) used transmitters that ranged 
up to 3.7 percent of body weight in mink and about 8 percent 
of body weight in ground squirrels. 


They were coated with medical grade Energy Technology 
Urethane to ensure that they would not produce adverse 
reactions in biological tissues. They were gas-sterilized and 
sealed in plastic surgical bags for storage until implanted. 


These new transmitters were first used in five animals 
in California, beginning in March 1984. Their reliability 
proved to be excellent and their lifespan approached the 
maximum expected battery life of 700 days. They had a rather 
limited range of approximately one mile from surface 
monitoring stations. Engineers at the Cedar Creek 
Bioelectronics Laboratory then reconfigured the placement of 
the internal components of the transmitter and redesigned the 
antenna. These’ improvements increased the range 
significantly, up to five miles from surface stations and 10 
miles from aircraft. This new design was used in subsequent 
transmitters, beginning in Alaska in the summer of 1984 and 
in California in the spring of 1985. 


FIGURE 1.4 - A schematic drawing of the implanted radio 
transmitter used during this study showing its component parts. 


Encapsulating material 


Antenna 


Lithium battery 
Urethane coating 
Lithium Expansion spacer 
battery 


Lithium battery 


Expansion spacer 


Drugs 


The otters were immobilized using the methods given in 
detail in Williams, et al., (1981). Fentanyl was given 
intramuscularly at dosages of 0.5-0.1 mg/kg of body weight in 
combination with azaperone at dosages of 0.010 to 0.053 mg/kg. 
This combination produced a safe, short-acting, and easily 
reversible immobilizing agent suitable for use under field 
conditions. The combination of anesthetic and tranquilizer 
was given to the otters while they were entangled in the gill- 
net or held in the Wilson trap or dip-net. 


Surgical procedures 


All surgery in California, and the initial surgery in 
Alaska, was carried out by Dr. Thomas D. Williams, who 
developed both the anesthetic procedures (Williams and Kocher, 
1978) and the surgical techniques (Williams and Siniff, 1983). 
In 1984, Dr. Williams trained two other veterinarians to do 
the operation and they performed some of the operations in 
Alaska. 


Surgery was carried out on a specially constructed 
operating table, either on board the capture boat or on the 
beach near the capture site. After an initial health 
screening procedure, the anesthetized otter was secured to the 
table with the ventral surface up. The status of the animal 
was monitored by capillary perfusions, color of the mucous 
membranes, respiratory rate and depth, temperature, and heart 
beat. A 50-50 mixture of KY jelly and betadine solution was 
applied to the ventral midline below the umbilicus and rubbed 
down to the skin. A comb was used to part the pelage and 
betadine solution was sprayed over the part. A sterile drape 
was placed over the ventral abdomen and thorax. Sterile 
gloves were used for each operation and all instruments were 
sterilized in benzol. 


Taggin 


In California, all instrumented animals were tagged on 
the hind flippers with colored Temple tags. In Alaska, both 
Temple tags and small button tags were used. The tagging 
methods and color/location coding system used in California 
were those used by both the California Department of Fish and 
Game and the U.S. Fish and Wildlife Service for many years 
(Ames, et al., 1983). The particular color combination used 
on each animal was selected in consultation with CDF&G 
personnel. 


Monitoring procedures 


We used different monitoring techniques and schedules in 
California and Alaska because of the climatic and geographical 
differences between the two areas. It was possible to monitor 
the animals all year in California but not in Alaska. The 
usual field season in Alaska extended from late April to mid- 
September. However, we visited Alaska occasionally during the 
winter and conducted aerial searches for instrumented animals. 


Routine daily monitoring in California was done from the 
ground. When animals could not be located from the ground, 
aerial searches were conducted using a small, ized wing plane 
with antennas attached to each wing. 


There were no roads in the Alaska study area. Some 
monitoring was done from a small boat but it was impossible 
to locate each animal every day due to the large size and 
complicated geography of the study area. Much of the 
monitoring was done from the air. 


In California, we evaluated the accuracy and precision of 
the sea otter locations that were obtained by telemetry 
triangulation. Transmitters were placed in floating buoys and 
positioned along the coast so that readings could be taken 
according to the established procedures we used in our 
monitoring of instrumented otters. The results of this 
evaluation are presented in Chapter 7. 


LITERATURE CITED 


Ames, J.A., R.A. Hardy, and F.E. Wendell. 1983. Tagging 
materials and methods for sea otters, Enhydra lutris. 
Calif. Fish and Game 69:243-252. 


=== ——————— - 1986. A simulated translocation of sea otters, 
Enhydra lutris, with a review of capture, transport, and 
holding techniques. Marine Resources Technical Report 
NOs BB, NY jyas 


Eagle, T.C., J. Choromanski-Norris, and V.B. Kuechle. 1984. 
Implanting radio transmitters in mink and Franklin's 
ground squirrels. Wildl. Soc. Bull. 12:180-184. 


Garshelis, D.L. 1984. Age estimation of living sea otters. 
J. Wildl. Manage. 48:456-463. 


Garshelis, D.L. and D.B. Siniff. 1983. Evaluation of radio- 
transmitter attachments for sea otters. Wildl. Soc. 
Bull. 11(4) :378-383. 


Schneider, K.B. 1973. Age determination of the sea otter. 
Alaska Dept. Fish and Game, Fed. Aid in Wildlife 
Restoration, Final Report, Proj. W-17-4 and W-17-5, Job 
8.10R. 23 pp. 


Smith, H.R. 1980. Intraperitoneal transmitters in suckling 


white-footed mice, Peromyscus leucopus. Biotelemetry 
Patient Monitoring 7:221-230. 


Williams, T.D., A.L. Williams, and D.B. Siniff. 1981. 
Fentanyl and azaperone produced neuroleptanalgesia in the 
sea otter, (Enhydra lutris). J. Wildl. Dis. 17:337-342. 


Williams, T.D. and F.H. Kocher. 1978. Comparison of 
anesthetic agents in the sea otter. J. Am. Vet. Med. 
Assoc. 173:1127-1130. 


Williams, T.D. and D.B. Siniff. 1983. Surgical implantation 
of radiotelemetry devices in the sea otter. J. Am. Vet. 
Med. Assoc. 183:1290-1291. 


CHAPTER 2 


REPRODUCTION, SURVIVAL AND TAG LOSS IN CALIFORNIA SEA OTTERS 


D. B. SINIFF AND K. RALLS 


November 30, 1988 


INTRODUCTION 


The California sea otter population is listed as 
threatened on the U.S. Endangered Species list and its status 
and management are of concern to several state and federal 
agencies. A population model is a basic tool for the 
understanding and management of any wildlife population. The 
development of a population model requires reliable estimates 
of reproductive and survival rates; no estimates of these 
rates are available for the California sea otter population. 


Early knowledge of the general biology of the sea otter 
reproductive cycle was gained mostly by the examination of 
reproductive tracts from animals collected in the U.S.S.R. or 
Alaska (Barabash-Nikiforov, 1947; Sinha, et al., 1966; Kenyon, 
1969; Schneider, 1973). These studies showed that the litter 
size is typically one, with maternal care extending at least 
four months after parturition, and that a birthing peak occurs 
in the spring, although birth can occur at any time of year. 
These early studies generally placed the inter-birth interval 
at about two years. However, subsequent observations of 
tagged sea otters in both California and Alaska have indicated 
that the inter-birth interval is closer to one year (Jameson 
and Johnson, 1979; Loughlin, et al., 1981; Wendell, et al., 
1984). Considerable data on tagged individuals are available 
for the California sea otter population (Estes and Jameson, 
1983; Wendell, et al., 1984) but it has not been possible to 
obtain good estimates of survival rates from these data for 


several reasons. For example, information on tag loss rate 
and difficulty in resighting tagged individuals greatly 
complicates this estimate. The only available data on tag 


loss rate is based on the resighting of tagged individuals 
(Ames, et al., 1983). It is thus an estimate of the rate at 
which tagged individuals disappear from the pool of regularly 
re-sighted animals, rather than a direct estimate of the rate 
at which individual tags are lost. 


In this chapter, we present estimates of the proportion 
of adult females accompanied by pups throughout the year; the 
inter-birth interval; the period of pup dependency; and 
reproductive, survival and tag loss rates for the California 
sea otter population. All estimates are based on observations 
of telemetry instrumented, flipper-tagged sea otters. 


METHODS 


We captured 49 otters (Table 2.1). Females known to be 
pregnant and small pups were not implanted with transmitters. 
We implanted radio transmitters in 40 otters, which were 
assigned to age/sex classes on the basis of their weight, 
estimated age (sometimes from teeth annuli), and, in the 


TABLE 2.1 -- Summary of sea otters captured in California 
during 1984 and 1985. 


OTTER CAPTURE CAPTURE SEX WT. TRANS. 

NO. DATE AREA LBS) FREQ. LEFT;RIGHT TAG 

1 7Mar84 Morro Bay 74 723 4/5 red;1/2 silver 

2 16Mar84 Morro Bay 65 955 4/5 white;1/2 silver 

3 21Mar84 Morro Bay 44 545 4/5 chartreuse;1/2 silver 
4 21Mar84 Morro Bay 53 784 4/5 1t. blue;1/2 silver 
6+ 3Ju184 San Simeon 36 807 1/2 orange;1/2 roy.blue 

7 15Feb85 Big Sur R. 54 333 4/5 pink;4/5 purple 

8 1Mar85 Rancho Rico # 49 none 1/2 1t. green;1/2 purple 
9 1Mar8s85 Wreck Beach 30 233 1/2 silver;1/2 purple 

10 1Mar85 Wreck Beach 60 041 1/2 1t. blue;4/5 purple 
11 15Mar85 Molera Point 35 417 4/5 red;1/2 purple 

12 16Mar85 Grimes Point # 45 none 4/5 chartreuse;1/2 purple 
13 16Mar85 Torre Canyon 25 461 4/5 white;4/5 purple 

14 16Mar85 Torre Canyon 41 217 4/5 1t. blue;1/2 purple 
15 20Mar85 Torre Canyon 43 842 1/2 purple;4/5 orange 

16 3Aprs5 Wreck Point 39 884 4/5 pink; 1/2 purple 

17 3Aprs5 Grimes Point 53 230 1/2 1t. green;4/5 purple 
18 3Aprs5 Torre Canyon # 57 none 1/2 silver;1/2 purple 

19 3Apr85 Rancho Rico 36 373 4/5 white; 1/2 purple 
20 3Apr8s5 Rancho Rico none 4/5 purple;1/2 purple 

21 10Apr8s5 False Sur 45 133 1/2 dk. green; 1/2 purple 
22 10Apr8s5 Big Sur R. 42 562 1/2 roy. blue; 1/2 purple 
23 10Apr8s5 False Sur 63 062 1/2 roy. blue; 4/5 purple 
24 10Aprs5 Point Sur * -- none ear tag 217 
-- 10Apr85 Point Sur -- none none 
25 13Apr8s5 Anderson Crk 56 933 1/2 orange; 1/2 purple 
-- 13Apr8s5 N. of slide -- none none 
-- 13Aprs5 N. of slide * -- none none 
26 8May85 Little Sur R. 46 680 4/5 1t. green; 1/2 purple 
27 40ct85 Dolan Rock 46 121 4/5 1t. green; 1/2 purple 
28 40ct85 Esalen 36 256 4/5 orange; 1/2 purple 

29 110ct85 Dolan Rock SiG 6S 5ieel /2iesa vex 2p anik< 

30 110ct85 Dolan Rock 34 475 1/2 white; 4/5 purple 

31 110ct85 Buck Cr. 40 904 4/5 silver; 1/2 purple 

32 180ct85 Buck Cr. # 50 none 4/5 gold; 1/2 purple 

33 180ct85 Big Slide 46 970 1/2 gold; 1/2 purple 

34 190ct85 J.P. Burns 80 625 4/5 1t. blue; 4/5 purple 
35 8Nov85 Ragged Pt. 35 960 1/2 1t. blue; 4/5 chartreuse 
36 8Nov8s5 County Line 37 433 4/5 red; 1/2 chartreuse 
37 22Nov85 Beckets Rf. 25 380 4/5 orange; 1/2 chartreuse 
38 22Nov85 Cypress Ovrl. F 25 603 4/5 pink; 1/2 chartreuse 
39 22Nov85 San Carpoforo F 30 587 4/5 yellow; 1/2 chartreuse 


ed) testes od eed Pred 1) Ta) Feed ea) Ua) 00a) aed Pee) 5p tesa) ae) Ua od og) 19 td Pood) eal ead cena) ed tee 
| 
fey) 


15 


TABLE 2.1 (continued) 


OTTER CAPTURE CAPTURE SEX WT. TRANS. 

NO. DATE AREA LBS) FREQ. LEFT;RIGHT TAG 

40 17Dec85 Piedras Blan. 31 405 4/5 1t. blue; 1/2 chartreuse 
41 17Dec85 Piedras Blan. 28 152 1/2 orange; 4/5 chartreuse 

42 17Dec85 San Carpoforo 35 273 4/5 1t. green; 1/2 chartreuse 
43 18Dec85 Piedras Blanc 353 1/2 pink; 4/5 chartreuse 

44 18Dec85 Piedras Blanc 33 534 4/5 white; 1/2 chartreuse 

45 18Dec85 Piedras Blanc 27 493 4/5 purple; 1/2 chartreuse 

46 18Dec85 San Simeon Pt 25 301 4/5 silver; 1/2 chartreuse 

47 30Dec85 Lover's Pt. 30 028 1/2 pink; 1/2 orange 


hy hy hy ey Ss oS 
uJ 
| aed 


+ Otter 5 was captured and tagged by the U.S. Fish and Wildlife 
Service 

# Pregnant female 

* Pup 


case of females, reproductive history during the monitoring 
period. Our sample consisted of nine adult males, five 
juvenile males, 16 adult females, and 10 juvenile females 
(Table 2.2). 


Adult females were located on an almost daily basis by 
radiotelemetry. We then attempted to observe them visually 
through binoculars and a Questar spotting scope (up to 80 
power) and record the presence or absence of pups and flipper 
tags. This was often impossible due to weather conditions, 
such as fog or rough seas, or difficult lighting conditions. 
Some individuals were more difficult to observe then others, 


depending upon their location along the coast. Thus the 
length of time between visual observations varied across 
individuals. 


To determine the proportion of females accompanied by 
pups each month of the year, we tallied whether or not each 
female was accompanied by a pup for every month she was 
monitored (Appendix 2.1). We considered that a particular 
female had been accompanied by a pup for a given month if we 
knew she had been accompanied by a pup for more than a 15 day 
period that month. If she had been accompanied by a pup for 
less than a 15 day period, we considered that she had not been 
accompanied by a pup that month. Months where the status of 
a particular female -was unknown were not counted. The 
variation in the number of days between sighting of the 
individual female otters created a problem in calculating the 
inter-birth interval and the number of days a pup remained 
with a female. In the case of an inter-birth interval (the 
time from birth of one pup to birth of the next pup), we 
usually did not know the exact birth dates of the two pups, 


16 


TABLE 2.2 -- A list of sea otters that were instrumented with 
implanted radio transmitters, their estimated age when 
available and other vital statistics. 


OTTER SEX WEIGHT ESTIMATED PUPPED? AGE/SEX 


NUMBER LBS AGE CLASS 

1 M 74 25 SF ADULT MALE 

2 M 65 oS QS ADULT MALE 

3 M 44 2S SS ADULT MALE 

4 M 53 2S 2S ADULT MALE 

6 F 36 SO NO ADULT FEMALE 

7 M 54 7* SO ADULT MALE 

9 F 30 5* YES ADULT FEMALE 
10 M 60 6* ae ADULT MALE 
11 F 35 == YES ADULT FEMALE 
13 M 25 SO am JUVENILE MALE 
14 F 41 6* YES ADULT FEMALE 
15 F 43 5% YES ADULT FEMALE 
16 F 39 S52 YES ADULT FEMALE 
17 M 53 6* oS ADULT MALE 

19 F 36 SS YES ADULT FEMALE 
21 F 45 15* NO ADULT FEMALE 
22 F 42 8 NO ADULT FEMALE 
23 M 63 5 aS ADULT MALE 
25 F 56 7 YES ADULT FEMALE 
26 F 46 2S YES ADULT FEMALE 
27 F 46 10 YES ADULT FEMALE 
28 F 36 13* NO ADULT FEMALE 
29 F 36 3 NO JUVENILE FEMALE 
30 M 34 2* a JUVENILE MALE 
31 F 40 11 YES ADULT FEMALE 

33 F 46 10 YES ADULT FEMALE 

34 M 80 8 == ADULT MALE 
35 M 35 2 oo JUVENILE MALE 
36 F 37 9 YES ADULT FEMALE 

37 F 25 2 NO JUVENILE FEMALE 
38 F 25 1* NO JUVENILE FEMALE 
39 F 30 2 NO JUVENILE FEMALE 
40 F 31 3 NO JUVENILE FEMALE 
41 M 28 2 =D JUVENILE MALE 
42 1p 35 2 NO JUVENILE FEMALE 
43 M 31 <1 o> JUVENILE MALE 
44 F 33 2 NO JUVENILE FEMALE 
45 F 27 1 NO JUVENILE FEMALE 
46 F 25 2 NO JUVENILE FEMALE 
47 F 30 <1 NO JUVENILE FEMALE 


*The teeth from these individuals were damaged so that only 
a minimum age could be estimated. 


although we knew that each birth occurred within some time 
period. For example, if we saw an instrumented female without 
a pup on 1 April and saw her again with a pup on 10 April, the 
pup had obviously been born between these two dates. Suppose 
we continued to monitor this female throughout the period of 
pup dependency and her next pregnancy. We last saw her 
without a pup on 15 March of the following year but she had 
a pup when next seen on 10 April. There were thus two 
intervals when her reproductive status was unknown. The first 
interval was 10 days and the second was 26 days. To calculate 
the inter-birth interval, we added one-half of each of these 
intervals to the number of days between the first sighting of 
the female with a pup one year and the first sighting of this 
female with her next pup. For our example, then, the 
inter-birth interval was 383 days -- 365 days from 10 April 
of the first year to 10 April of the second year plus five 
days from the interval between 1 April and 10 April spanning 
the birth of the first pup and 13 days from the interval 
between 15 March and 10 April spanning the birth of the 
subsequent pup. To make this calculation, we arbitrarily 
included for the interbirth interval estimates, only data for 
which neither of the unknown status intervals was more than 
51 days. 


The same problem arose in calculating the number of days 
the pup remained with the female. Again, there were two 
intervals when the status of the pup was unknown, one spanning 
its birth and a second spanning its disappearance. As before, 
we used only data in which both intervals for which the status 
of the pup was unknown were less than 52 days. We divided 
each interval in half and added it to the period between the 
first and last dates the pup was seen with the female. 


We estimated the annual reproductive rate in two ways. 
The first method was based on the total number of days all the 
adult females were monitored and the known number of pups born 
to them during this period. This calculation again included 
intervals when the adult females were not seen and their 
status was unknown. The longest such intervals were 10 months 
for one female (otter 11) and four months for another (otter 
16) (Appendix 2.1). We summed the total number of monitoring 
days across 13 adult females, and divided this figure by the 
total number of pups produced to obtain the average number of 
days required to produce a pup. This average number of days 
was then divided into 365 to obtain an estimate of the annual 
pupping rate. 


In the second method, we considered only the five females 
used to calculate the inter-birth interval and the period the 
' pup remained with the female. For this data set, we divided 
the average inter-birth interval into 365 to obtain an 
estimate of the annual pupping rate. 


18 


Our first method of estimating the annual reproductive 
rate is probably most similar to that which one would use for 
tag-resight data (Wendell, et al., 1984), where females are 
observed periodically and their reproductive status noted. 
This method also produces gaps where the status of the female 
is unknown; the length of these gaps varies with the frequency 
of the attempts to re-sight the females and the difficulty of 
observing each female. 


The estimation of survival rates from radiotelemetry data 
may be approached somewhat differently than that from data 
derived: from resighting or recapture of tagged animals. 
Instrumented animals are known to be alive or dead on a daily 
basis. Thus the procedure that has been developed for 
radiotelemetry data is to estimate a daily survival rate and 
expand it to the time period desired, usually one year, 
assuming that the daily rate remains constant over this period 
(Heisey and Fuller, 1985). The formulation recommended by 
Heisey and Fuller is: 


Annual Survival = (Transmitter days - deaths) *° 


Transmitter days 


When days are used as the basic time interval, it is 
necessary to assume that the status of each individual (dead 
or alive) is known for each day. This assumption was not 
fulfilled for our animals when individuals disappeared and we 
were unable to determine their fate. These individuals were 
classified as missing (Appendix 2.2), and these individuals 
might have died or their transmitters might have expired. 
Animals classified as transmitter failed, transmitter expired, 
or transmitting were known to be alive at the number of days 
indicated in Appendix 2.1. The way in which these missing 
animals are treated can affect the survival rate estimate. 
One way to handle this problem is to assume that transmitters 
are extremely reliable for some number of days, and animals 
lost prior to this time have died. We used 450 days for this 
decision point for missing animals. Thus, we assumed that 
otters missing in less than 450 days from the capture date had 
died and that those missing after 450 days were alive on the 
date they became missing. This criterion was based on the 
average life span, 485 days, of the five transmitters in which 
expiration was verified. We also calculated survival rates 
for each age/sex class by using the number of individuals in 


each sex and age class as the basis of calculations. This 
means following the binomial model: 
Seal - yn where: 
S = Estimate of survival 
D = number of animals that died 
n = sample size for the particular sex/age 


category. 


For this model an estimate of variance is easily calculated 
by the standard: 


s* = "47n where: 
p = survival rate estimate 
q = D/n, or the proportion dying 


It only remains to specify the time interval, which is 
normally taken to be one year. However, in this case we 
followed the same 450 day criterion. Thus the binomial model 
for an estimate of the annual survival rate becomes: 


Ss = (1 - "yn 365/450 


We estimated annual survival rate each of these ways for the 
four sex and age classes of adult females, adult males, 
juvenile females and juvenile males. 


The pup survival rate was based on the ratio of the 
number of pups that died during the period of dependency to 
the number of pups born, over the entire monitoring period. 
We assumed that pups that remained with the female less than 
150 days died, based on an estimated pup dependency period of 
six months (Payne and Jameson, 1984; Wendell, et al., 1984). 


One of the advantages of being able to locate animals by 
radiotelemetry is that they are still identifiable after tag 
loss has occurred. All of our instrumented animals were 
tagged with one Temple tag in each hind flipper, following a 
procedure developed by the California Department of Fish and 
Game (Ames, et al., 1983). In addition, the tags were drilled 
so that a small nylon or stainless steel screw could be used 
to hold the two sides of the tag together. A small amount of 
glue was dripped into the hole before the screw was inserted. 
Our data on the presence or absence of tags were similar to 
our data on reproduction. However, tags were more difficult 
to see than pups, so many of the intervals between the last 
date a tag was seen and the first date it was seen to be 
missing were longer than the corresponding intervals for pups. 
We analyzed these data using the same method we used to 
analyze the data on the survival of individual otters. 
However, we assumed that a tag had survived until the date it 
was seen to be missing and thus calculated only the maximum 
possible tag survival rate. 


20 


RESULTS 


Although some proportion of adult females were accompanied 
by pups throughout the year, this proportion peaked in the 
spring, with a secondary peak in the fall (Fig.2.1). Our 
total reproductive data set includes information on the 13 
adult females that were monitored at least 355 days (Table 
2.3). However, our best data for the determination of the 
inter-birth interval and the period the pup remained with its 
mother came from five females for which the length of all 
three intervals in which their status was unknown (those 
spanning the birth of the first pup, the disappearance of the 


Table 2.3 -- Reproductive and age data and length of 
monitoring period for adult female sea otters. 
OTTER AGE DAYS PUPS PUPS 
NUMBER YRS MONITORED BORN DIED 

6 = 355 0 (0) 

9 5* 544 2 1 
11 = 621 1 (0) 
14 6* 744 2 (0) 
15 5% 725 1 0 
16 = 608 2 (0) 
19 = 587 3 (0) 
22 8 585 (0) (0) 
25 7 555 1 (0) 
27 10 540 1 1 
31 11 637 2 2 
33 10 630 2 2 
36 9 609 2 2 
TOTALS 7740 19 8 


*Only a minimum age could be estimated from these individuals 
due to tooth damage. 


first pup, and the birth of second pup) was less than 52 days 


(Table 2.4). For this data set, the average inter-birth 
interval was 416 days. Our first estimate of the annual 
reproductive rate is based on the information summarized in 
Table 2.3. Nineteen pups were produced during 7,740 


monitoring days; thus, an average of 407 days was required to 
produce one pup, and the estimated annual reproductive rate 
is 0.90 pups per adult female. Our second estimate is based 
on the data in Table 2.4. The average inter-birth interval 
is 416 days, which corresponds to an annual reproductive rate 
of 0.88 pups per adult female per year. 


Eight of the 19 pups almost certainly died before 


21 


FIGURE 2.1 - The percent of the adult female study animals in 
California that were with pups during each month of the year. 


Percent 


O Females with Pups 


39 4 Females with Pups, 
Running Average 


28 | Oo 


Jan Feb Mar Apr May June July Aug Sept Oct Nov Dec 


‘Month 


Die 


weaning, giving an estimated pup survival rate of 0.57 (Table 
PE) B We estimated that these eight pups (Table 2.4B) 
remained with their mother for less than 65 days. Our 
estimate assumes that pups born shortly before the females' 
transmitters expired lived, when in fact we were unable to 
determine their fate, and it may exclude some pups that lived 
only a short time and were thus undetected (Appendix 2.1). 


Considering all 80 flipper tags in our sample (Appendix 
2.3), the estimated annual tag survival rate was 0.74 (Table 
2.6A). Survival rates for right and left tags were not 
Significantly different. However, more otters lost two tags 
than would be expected by chance, based on the assumption that 
tag loss follows a binomial distribution. Using the number 
of tags lost over the period of monitoring to derive the 
expected probabilities of losing one, two, or no tags and 
comparing these expected values to the data, we see that 
significantly more otters lost both tags than expected (Table 
2.6C). This result is perhaps not surprising, as some 
individuals have been seen to bite and manipulate their 
flipper tags, ultimately removing them, while other 
individuals appear to ignore the presence of tags. If the 
otters that lost two tags are excluded, the estimated annual 
tag survival rate increases to 0.91 (Table 2.6B). 


Both methods for estimating survival rates for the four 
main age/sex classes, indicated the highest survival for adult 
females and lowest for adult males (Table 2.5). Estimates for 
juvenile females were lower than those for adult females, 
while estimates for juvenile males were higher than those for 
adult males. The survival rate of juvenile females was lower 
than that of juvenile males for both methods of estimation. 
The data on which we based these estimates are presented in 
Appendix 2.2. 


DISCUSSION 


Our data on the proportion of females accompanied by pups 
over the yearly cycle, which show a peak in the spring and a 
smaller peak in the fall, are similar to data on the 
proportion of pups to independent otters recorded by the 
California Department of Fish and Game in their monthly counts 


of index areas (Brody, Chapter 10). The independent otters 
in these index areas are probably largely adult females along 
with a few territorial males and juvenile females. This 


bimodal pattern could be the result of females that 
successfully raise pups having a longer inter-birth interval 
than those that do not, thus placing the birth of their next 
pup at one year plus three to four months. Our data are in 
agreement with the emerging consensus that many female sea 
otters pup on an approximately annual basis (Loughlin, et al., 


23 


Table 2.4 -- Reproductive information on adult female sea 
otters that were known to have given birth to at least one 
pup. The table gives the number of days during which the 
reproductive status of these females was known and unknown 
(A). Calculated periods with and without pups and inter-birth 
intervals for those animals where the unknown intervals were 
not more than 51 days are shown below (B). 


A. DATA 

Otter Un- Known Un- Known Un- Known Un- 
known Days known Days known Days known 
Status With Status Without Status With Status 
(Days) Pup (Days) Pup (Days) Pup (Days) 

One Two 
9 13 36 11 330 51 

ALI a 113 1 318 

14 11 149 20 390 

15 2 81l** 

16 41 117 104 101 120 28%* 

19 68 85 80 91 57 39 5 

25 2 176 8 

27 11 24 26 

31 8 16 8 313 28 41 20 

33 10 31 2 376 5 8 19 

36 37 26 2 258 17 1 5 


B. CALCULATED VALUES 


OTTER PERIOD WITH PERIOD UNTIL INTER-BIRTH 
NUMBER PUP (DAYS NEXT BIRTH (DAYS) INTERVAL (DAYS 
g* 48 361 409 

14* 165 421 586 

25 181 

27 43 

31% 24, 65 331 355 

33% 37, 20 380 417 

36* 46, 12 266 312 

MEAN = 64 MEAN = 416 


* Five females used to calculate the inter-birth interval. 
** Still with pup when last seen. 


24 


Table 2.5 -- Annual survival rate estimates. Two estimates 
were made for the age/sex categories of adult females, adult 
males, juvenile females and juvenile males. The first was 
obtained by converting the daily survival rate of radio- 
instrumented otters to an annual rate (Heisey and Fuller, 
1985). The second used the standard binomial model (see text 
for explanation) and the status of individual sea otters as 
a basis for the estimate. The estimated pup survival rate was 
based on the pups born to the instrumented females. 


TRANSMITTER DAYS INDIVIDUALS 
AS THE BASIS* AS THE BASIS* 


: STANDARD 

SURVIVAL SURVIVAL DEVIATION 
ADULT FEMALES 0.91 0.89 0.088 
ADULT MALES 0.61 0.52 0.167 
JUVENILE FEMALES 0.80 0.75 0.145 
JUVENILE MALES 0.88 0.85 0.179 


*Animals that were classified as missing before 450 days were 
assumed to have died at the time they became missing, while 
those classified as missing after this time were assumed to 
have been alive when they became missing. 


PUPS TO WEANING 0.57 
1981; Estes and Jameson, 1983; Wendell, et al., 1984; 
Garshelis, Johnson, and Garshelis, 1984). Our average 


inter-birth interval of 416 days is clearly within the range 
of the 17 intervals recorded by Wendell, et al., (1984) based 
on observations of tagged otters. Although our sample size 
is smaller than theirs, the intervals when the reproductive 
status of the female was unknown tended to be smaller in our 
data. One interesting aspect of our data is the rather large 
degree of variation, with a range of 312 to 611 days. The 
reason for this variation may be correlated with the length 
of the dependency period. It is generally agreed that female 
otters rarely, if ever, mate when accompanied by a pup 
(Kenyon, 1969; Calkins and Lent, 1975; Garshelis, et al., 
1984). Clearly more data are needed on the relationship 
between the inter-birth interval and the period of pup 
dependency, which probably varies with the age and condition 
of the female. 


Minimum reproductive rates for sea otters in Alaska have 
been suggested by the reproductive condition of females in 
samples killed in late winter and spring (Kenyon, 1969; 
Schneider, 1973). Kenyon (1969) found that 71% of animals 
collected in this period were pregnant and 17% had recently 


25 


Table 2.6 -- Estimates of annual tag survival rates based upon 
the survival rate estimation procedures for instrumented 
animals as outlined by Heisey and Fuller (1985). The only 
difference here is that the "death" of tags rather than 
individuals is the unit of measure (A and B). A comparison 
of tag loss to expected tag loss based on the binomial 
distribution is shown in C. 


A. ALL OTTERS 


TAG ANNUAL NUMBER OF 
SURVIVAL TAGS LOST 
RATE (OF 80) 

RIGHT 0.70 13 

LEFT 0.78 10 

ALL 0.74 23 


B. EXCLUDING OTTERS THAT LOST TWO TAGS 


TAG ANNUAL NUMBER OF 
SURVIVAL TAGS LOST 
RATE (OF 64) 
ALL 0.91 ~ 6 


C. COMPARISON OF TAG LOSS TO THE BINOMIAL DISTRIBUTION 
ESTIMATED PROBABILITY OF LOSING ONE TAG = 23/80 = 0.2875 


NUMBER OF NUMBER OF OTTERS EXPECTED FROM 
TAGS LOST BINOMIAL MODEL 
NONE 25 21 
ONE 8 16 
TWO 7 3 
P<0.05, CHI-SQUARE = 61, 1 D.F. 
given birth. Schneider (1973) found that for samples 


collected in May, 59% were pregnant and 14% had recently given 
birth. Combining these values gives approximate annual rates 
of .88 and .73 for these two studies, respectively. No 
estimates are available for the California population. Our 
estimates, based on the number of pups born during the 
monitoring period and the average inter-birth interval are, 
as expected, slightly higher than the minimum rates available 


26 


for Alaska since ours include the complete annual cycle. 
Because of the sea otter's ability to produce pups throughout 
the year, and the probable relationship between the 
inter-birth interval and the period of pup dependency, it 
seems likely that estimates of annual reproductive rates in 
sea otters will be quite dependent upon conditions during the 
particular interval during which the data are collected. 


Our estimated survival rate for pups from birth to 
weaning, 0.57, while possessing the potential for bias, as 
mentioned under results, is close to the 0.50 estimate needed 


TABLE 2.7 --A comparison of the age of pups at separation from 
the female in California and Alaska, based on our data and 
other published data. 


DAYS PUP ALASKA* CALIFORNIA** 
WITH FEMALE 

0-50 0) 8 
51-100 3 3 
101-150 1 2 
151-200 4 5 


* Data from Garshelis (1983). 
** Data from Table 2 plus eight pups from Loughlin, et al., 
(1981). 


SSaaa———aBa=a=ESeeeeeee————— eee 


to produce a zero population growth rate in our population 
model when combined with our other survival rate estimates 
(Chapter 10). Surveys suggest that the population has been 
stable over the past decade (Estes and Jameson, 1983). Table 
2.7 contrasts survival patterns of dependent pups in 
California and Alaska. Early mortality appears to be more 
frequent in California. This pattern, if confirmed by 
additional data, may be a result of storm patterns in 
California, because the Alaskan data were collected in Prince 
William Sound, which provides more shelter during periods of 
inclement weather. 


The annual survival rate estimates we obtained by 
expanding the daily survival rate based on transmitter-days 
also appear reasonable. The relatively close correspondence 
between the estimates calculated by two different methods is 
encouraging, as is their general agreement with other aspects 
of the data collected on our instrumented animals, such as the 


movement patterns, time budgets, and feeding patterns of the 
different age/sex groups. 


Our adult females, which had the highest survival rates, 
were in areas of the range where human activities were minimal 
and they traveled the least of the four sex/age groups. Adult 
males had lower survival rates than adult females and 
juveniles of both sexes. A low survival rate for adult males 
is also suggested by the sex ratios in the California 
Department of Fish and Game database on dead sea otters. 
Considering only the carcasses in good condition, where the 
sex of the carcass was known in almost all cases, 
significantly more dead adult males than adult females washed 
ashore (Table 2.8). Our adult males tended to be particularly 
vulnerable during periods of travel. One was shot after he 
had moved a considerable distance from his capture location 
and two others disappeared when they were moving through areas 
in which gill-net boats were operating. Juvenile males also 
traveled extensively but tended to remain farther offshore, 
which may provide some degree of protection from human 
activities, such as shooting (Wild and Ames, 1974; Morejohn, 
et al., 1975) and incidental capture in gill-nets (Wendell, 
et al., 1986), and contribute to their higher survival rate. 
The only juvenile male that died during our study was almost 
certainly attacked by a shark. 


Table 2.8. - The California Department of Fish and Game 
maintains a data-base on the dead sea otters that wash ashore. 
Each carcass is given a condition rating, aged and sexed if 
possible. Numbers of male and female carcasses in good 
condition are compared to those expected if the sex 

ratio was equal. 


SEX 
AGE CLASS 
(after Morejohn MALES FEMALES 
et al. 1975) 
Observed Expected Obs. Expected 


Pups/immatures 52 53 54 53 ns 
Subadults 23 21.5 20 21.5 ns 
Adults 84 69.5 55 69.5 p<0.05* 
Old adults 29 28 27 28 ns 


*Chi-square = 6.1, 1 d.f. : 


28 


Juvenile females had low survival rates compared to adult 
females. They also had slightly lower survival rates than 
juvenile males, as did the juvenile females in Alaska (Chapter 
9). However, in the CDF&G database on dead sea otters, again 
considering only the carcasses in good condition, the number 
of juvenile male and female carcasses was about equal, 
suggesting similar mortality rates for the two sexes of 
juveniles in California. Juvenile females in California 
clearly foraged longer than other otters to obtain sufficient 
food (Chapters 4 and 5) and one juvenile female in Alaska 
starved to death (Chapter 9). Taken as a whole, our data 
strongly suggest that the age/sex groups are differentially 
affected by the various sources of sea otter mortality. 


Our observations of tags on instrumented individuals 
provide estimates of annual tag loss rates that exclude the 
possibilities that mortality and movements of tagged 
individuals out of the study area added to the perceived tag 
loss. The annual tag survival rate of 0.74 represents a good 
estimate for the Temple tag. The only other estimates of tag 
loss rates are those of Ames, et al., 1983, who used three 
methods of tag application and used resightings of individuals 
over time to estimate loss rate. The use of resighting 
information to estimate tag loss has the potential to include 
loss due to death and movement of individuals out of the study 
area in addition to tag loss, and thus one would expect it to 
overestimate actual tag loss. Curiously, however, using our 
annual tag survival of 0.74 (which excludes the other sources 
of loss of mortality and movement of the area), we found the 
loss rate to correspond almost exactly to the loss rate for 
double-anchored tags in Ames, et al., 1983 (Fig. 2.2). When 
we used our data and plotted the number of tags remaining on 
the otters that were still being located on a regular basis 
against time (subtracting from the original 80 tags all those 
that were on otters with failed or expired transmitters and 
on missing and dead otters plus those that were lost by 
animals with functioning transmitters), the estimated tag loss 
was of course higher and corresponded very closely to that for 
the other two methods of tag application (single anchor, 
unglued; and single anchor, glued) in Ames, et al., 1983 (Fig. 
Davie 


Ames, et al., found that the apparent loss rate of 
double-anchored tags was less than the apparent loss rate of 
single-anchored tags (whether glued an unglued). Our data 
suggest two possible reasons for this difference. The first 
possibility is that the retention of double-anchored tags was 
very high and the observed loss rate was almost entirely due 
to mortality and/or movements out of the study area. The 
second is that the small number of animals on which this 
method was used rarely, if ever, moved out of the study area 
and did not suffer significant mortality during the study 


29 


FIGURE 2.2 - A comparison of tag-loss rates between this study 
and the study of Ames, et al, 1983. (See text for explanation.) 


oD) pete tte... 

Gaal eta Sees ee ee 

s Q oS ccoe, 
i) 

Eo. 

@ 

ai: 

H 

a: 

Fo 

rs) od —— Single Anchor, No glue 

Eo --— Single Anchor, Glued 

A collage --- Double Anchor, Screwed 

a oe 

<= ®O A This Study, Log of Tags 

eu Eo Remaining Out of 80 Tags. 

Ore Seis ces ~ This Study, Using Estimated 

— Annual Survival Rate (0.74). 


0 100 200 . 300 400 500 600 


Time (Days) 


30 


period. These conditions could occur, for example, if the 
tagged individuals were mostly adult females. 


Considerable data can be obtained by following tagged 
animals and tag-resight data can potentially be used to 
estimate annual survival rates for the various age/sex classes 
of a species. However, mark-recapture techniques, using the 
appropriate models, must be used for these survival rate 
estimates (Seber, 1973). For sea otter tag-resight data, the 
combined effects of unequal probability of sighting among 
individuals, movement of animals out of the intensive study 
area, and differential mortality patterns among age/sex groups 
make the application of such methods extremely difficult. 
When tag loss is added to these complications it becomes 
rather unlikely that accurate estimates of annual survival 
rates for sea otters can be derived from such data. 
Furthermore, the comparison of our data with those of Ames, 
et al., (1983) suggests that for sea otters, even under the 
best conditions, it would be difficult to separate tag loss 
from actual mortality and movement away from the study area. 
In our study, known tag loss, added to verified mortality, 
would have produced unrealistically low survival rate 
estimates. 


LITERATURE CITED 


Ames, J. A., R. A. Hardy, and F. E. Wendell. 1983. Tagging 
materials and methods for sea otters, Enhydra _ lutris. 
Calif. Fish and Game 66: 196-209. 


Barabash-Nikiforov, I. I. 1947. Kalan. (The Sea Otter, pp. 
1-174) Soviet Ministron RSFSR. Glavhoe upravlenie po 
zapovednikam. (In Russian.) (Translated from Russian 
by Dr. A. Birron and Z. S. Cole. Published for the 
National Science Foundation by the Israel Program for 
Scientific Translations, Jerusalem, 1962). 


Calkins, D. G. and P. C. Lent. 1975. Territoriality and 
mating behavior in Prince William Sound sea otters. J. 
Mammal. 56: 528-529. 


Estes, J. A. and R. J. Jameson. 1983. Summary of available 
population information on California sea otters. 
Minerals Management Service, POCS Technical Paper No. 
83-11, 29 pp. 


Garshelis, D. L., A. M. Johnson and J. A. Garshelis. 1984. 


Social organization of sea otters in Prince William 
Sound, Alaska. Canadian J. of Zoology 62: 2648-2658. 


31 


Heisey, D. M. and T. K. Fuller. 1985. Evaluation of survival 
and cause-specific mortality rates using telemetry data. 
J. Wildl. Mgt. 49: 668-674. 


Jameson, J. R. J. and A. M. Johnson. 1979. Evidence of 
annual reproduction among sea otters. Abstract, Third 
Biennial Conference on the Biology of Marine Mammals, 
7-11 October 1979, Seattle, WA. 


Kenyon, K. 1969. The sea otter in the eastern Pacific Ocean. 
N. Amer. Fauna 68: 1-352. 


Loughlin, T. R., J. A. Ames, and J. E. Vandevere. 1981. 
Annual reproduction, dependency period and apparent 
gestation period in two California U.S.A. sea otters 
Enhydra lutris. Fish. Bull. 79: 347-349. 


Payne, S. F. and R. J. Jameson. 1984. Early behavioral 
development of the sea otter, Enhydra lutria. J. Mammal. 
65: 527-531. 


Seber, G. A. F. 1973. The Estimation of Animal Abundance and 
related parameters. Hafner Press, N.Y. 


Sinha, A. A., C. H. Conway, and K. W. Kenyon. 1966. 
Reproduction in the female sea otter. J. Wildl.Manage. 
30: 121-130. 


Schneider, K. B. 1973. Reproduction in the female sea otter. 
Federal Aid in Wildlife Restoration Project W-17-4. 
Project Progress Report. Alaska Department of Fish and 
Game. 36 pp. 


Wendell, F. E., J. A. Ames, and R. A. Hardy. 1984. Pup 
dependency period and length of reproductive cycle: 
estimates from observations of tagged sea otters, Enhydra 
lutris in California. Calif. Fish and Game 70: 89-100. 


32 


CHAPTER 3 


MOVEMENT PATTERNS AND SPATIAL USE OF CALIFORNIA SEA OTTERS 


K. RALLS, T. EAGLE, AND D. B. SINIFF 


November 30, 1988 


33 


INTRODUCTION 


Sea otters tend to be sexually segregated in "male 
areas", occupied largely by males, and "female areas", 
inhabited by adult females and their young (Kenyon, 1969; 
Schneider, 1978). Breeding males maintain territories, either 
seasonally (Garshelis and Garshelis, 1984) or all year 
(Loughlin, 1980), within the "female areas". Male territories 
are often smaller than the home ranges of adult females 
(Loughlin, 1980; Garshelis and Garshelis, 1984), although 
life-time home ranges of males may exceed those of females 
(Kenyon, 1969; Garshelis and Garshelis, 1984). 


In Alaska, home ranges consist of extensively used areas 
connected by travel corridors (Garshelis and Garshelis, 1984; 
Chapter 8). Long-distance seasonal movements between "male 
areas" and "female areas" have been documented in both Alaska 
and California: four adult males moved over 100 km in Alaska 
(Garshelis and Garshelis, 1984) and three males moved over 80 
km in California (Ribic, 1982). Movement patterns of juvenile 
sea otters in Alaska have been studied by Monnett and 
Rotterman (Chapter 8); few data are available on the movement 
patterns of juveniles in California. 


Sea otters are a coastal species, although they may be 
found quite far from shore in shallow-water areas (Kenyon, 
1969). Loughlin (1980) and Ribic (1982) reported that they 
rarely venture beyond the outer limits of the kelp canopy in 
California. 


We report here on the movement patterns and spatial use 
of 38 California sea otters representing all four major age 
and sex classes: adult females, adult males, juvenile 
females, and juvenile males. The otters were instrumented 
with implanted radio transmitters. Because these transmitters 
had a much longer life span than those used in previous 
studies and we attempted to locate each individual every day, 
our data provide a more detailed picture of sea otter movement 
patterns and spatial use than previously available. 


METHODS 


Otters were assigned to sex and age classes based on 
their weight at capture, estimated age based on the 
examination of cementum layers in the vestigial premolar 
extracted for this purpose, and, in the case of females, 
reproductive performance. All juveniles were judged to be no 
more than two years of age (Chapter 2). 


We usually attempted to locate each instrumented otter 
on a daily basis by listening for their radio signals from 
points along the shore. Individual otters were located at 


34 


various times of day, depending upon their movements from the 
previous day and our searching pattern. Sometimes, usually 
when an animal had moved a considerable distance from its 
previous location, an individual could not be located for 
several days. We searched for missing individuals from the 
air; this was generally successful, as the radio signal could 
be heard from a greater distance from the air than from the 
shore. Location data were also recorded once an hour during 
24-hour watches. 


Three methods were used to estimate the position of a 
otter once its radio signal was detected: visual observation 
of the otter, triangulation on the radio signal, and, when 
neither of these was possible, the best judgement of the 
researcher based on the direction and strength of the radio 
Signal. The accuracy of the locations determined by 
triangulation was estimated by triangulating on radio signals 
from transmitters attached to buoys at known locations 
(Chapter 7). The method by which each location was estimated 
was coded in the data, enabling us to analyze only those 
locations determined by a particular method when appropriate. 
Unless otherwise noted, all location data were included in an 
analysis. Triangulations were plotted on topographic maps of 
the study area; locations were recorded in the form of x-y 
coordinates based on the UTM grid. 


We used several measures in our analyses. The average 
distance between successive locations of each individual 
otter, separated into those recorded between 18 and 36 hours 
apart and those recorded more than 36 hours apart, was used 
to compare the usual distance traveled on a short-term basis 
among the age/sex classes. This distance was measured on the 
UTM grid and is the straight line distance between the two 
locations. The path taken by the otter was probably longer. 


The area used by individual otters on a daily basis was 
calculated from the hourly triangulated locations recorded 
during 24-hour watches, using the minimum convex polygon 
method commonly used for terrestrial mammals (Hayne, 1949). 


To portray the movement patterns of individuals over the 
entire monitoring period, we moved each daily location for an 
individual to the nearest point along the 5-fathom contour. 
We then calculated the deviation of each daily location from 
the mean location for that otter along this contour and 
plotted these daily deviations against time. 


Only daily locations determined by triangulation were 
used to estimate distance offshore. The coastline was 
digitized in UTM coordinates and a BASIC program was written 
to calculate the distance from each otter-location to the 
nearest point on the shore. 


35 


Monthly movement patterns were examined using the average 
general harmonic mean distance (Hp) (Neft, 1966). This 
measure, which is calculated from the distance between all of 
an otter's locations for a given month, is insensitive to a 
few long distance movements, and thus, may better reflect 
Burt's (1943) conception of the home range. At least seven 
locations in a month were required before Hp was calculated. 


The "distance between extreme locations" (DBEL) was used 
to compare the length of coastline frequented by individuals 
over the entire monitoring period. This measure was first 
used by Garshelis and Garshelis (1984) in their studies of sea 
otters in Prince William Sound, Alaska, and is the distance 
between the two farthest-apart locations of an individual 
otter over the period of time it was monitored. It can he 
considered an approximation of an otter's range during the 
monitoring period. It is particularly useful for comparisons 
between California and Alaska data, as field conditions in 
Alaska preclude collection of the daily locations of each 
otter that enabled us to calculate other measures for the 
California otters. This distance was measured on the UTM 
grid. 


Statistical comparisons among age and sex classes were 
performed using analysis of variance, controlling for 
variation among individuals within classes. We performed a 
log (base 2) transform on the data to reduce heterogeneity of 
variances. All statements that differences are statistically 
significant are based on the 0.05 probability level. 


RESULTS 
Distance between successive locations 


Otters of all age and sex classes were usually found 
within a comparatively short distance of their location on the 
previous day. Data on the distance between successive 
locations were divided into two categories: those recorded 
within 18-36 hours of the previous location of that individual 
and those recorded after an interval of more than 36 hours 
(Tables 3.1 and 3.2). Analysis of variance (Appendix 3.1) 
indicated that significant variation occurred among 
individuals within all the age and sex classes in both 
categories except for the juvenile males in the greater than - 
36 hour category. This suggested that individual movement 
patterns from one day to the next were different for young 
males but that their long term patterns were similar. 
However, juvenile males were the most similar in both the 18- 
36 and more than 36 hour categories. The variation among the 
sex/age classes was much greater than that within classes. 
For all these data, even though a log transformation was used 


36 


to improve the homogeneity of the within individuals 
"variance", Bartletts test for homogeneity showed Signitvean’ 
differences in all cases. 


TABLE 3.1 - Average distance (km) between successive 
locations, recorded between 18 and 36 hours apart, for each 
instrumented otter along the California coast. AF = Adult 
Female, AM = Adult Male, JF = Juvenile Female, JM = Juvenile 
Male. 


OTTER AGE/ STANDARD 
NUMBER SEX MEAN N DEVIATION 
15 AF 0.355 342 0.428 
28 AF 0.434 10 0.124 
4 AM 0.489 115 1.004 
7 AM 0.517 326 0.485 
10 AM 0.711 282 1.661 
1 AM 0.717 229 4.119 
2 AM 0.724 89 1.939 
33 AF 0.757 272 1.056 
17 AM 0.773 138 1.147 
46 JF 0.823 375 0.731 
42 JF 0.825 345 1.806 
11 AF 0.897 311 0.806 
19 AF 0.900 194 1.493 
16 AF 0.909 248 1.396 
36 AF 0.918 390 0.855 
9 AF 1.018 265 1.113 
3 AM 1.044 183 3.152 
6 AF 1.091 101 1.522 
25 AF 1.105 284 1.629 
38 JF 1.134 18 0.928 
45 JM 1.244 164 1.836 
37 JF 1.260 117 1.677 
40 JF 1.284 285 3.206 
31 AF 1.369 308 - 1.463 
27 AF 1.453 180 1.692 
47 JF 1.536 66 1.378 
39 JF 1.680 244 3.187 
44 JF 1.716 77 3.314 
34 AM 2.051 47 5.946 
29 JF 2.165 237 2.547 
43 JM 2.170 186 2.159 
41 JM 2.299 229 2.335 
22 AF 2.354 187 2.944 
26 AF 2.401 12 1.796 
14 AF 2.409 381 3.282 
13 JM 2.569 83 5.488 
35 JM 2.744 241 4.821 
30 JM 2.961 230 3.339 


TABLE 3.2 —- Average distance (km) between successive 
locations, recorded more than 36 hours apart, for each 
instrumented otter along the California coast. AF = Adult 
Female, AM = Adult Male, JF = Juvenile Female, JM = Juvenile 
Male. 


OTTER AGE/ STANDARD 
NUMBER SEX MEAN N DEVIATION 
15 AF 0.417 108 0.462 
28 AF 0.912 6 1.155 
33 : AF 0.998 149 1.364 
46 JF 1.077 119 1.115 
44 JF 1.120 36 0.971 
36 AF 1.177 105 0.977 
42 JF 1.302 107 2.301 
11 AF 1.388 104 1.301 
31 AF 15393), Tau 1.411 
7 AM 1.405 101 5.453 
38 JF 1.407 10 1.323 
9 AF 1.496 97 2.080 
25 AF 1.579 96 2.465 
1 AM 1.587 70 7.452 
37 JF 1.608 70 2.389 
19 AF 1.646 136 2.398 
2 AM 1.652 31 3.181 
10 AM tba ve) alas) 8.601 
47 JF 1.936 178 1.928 
16 AF 1.973 123 3.325 
6 AF 2.017 66 1.903 
26 AF 2.126 7 1.191 
27 AF 2.336 92 2.428 
4 AM 2.343 115 9.117 
3 AM 2-442 94 7.626 
45 JM 2.514 102 4.740 
29 JF 3.346 110 3.980 
41 JM 3.499 82 3.726 
14 AF 3.546 133 4.533 
17 AM 4.524 52 10.050 
40 JF 4.562 115 14.330 
22 AF 5.137 124 6.954 
35 JM 5.295 84 9.836 
43 JM 5.424 129 9.360 
13 JM 6.220 136 12.030 
39 JF 6.552 76 11.470 
30 JM 8.075 118 15.740 
34 AM 28.070 55 46.420 


To illustrate the relatively short-term movement patterns 
of the four age/sex classes, we plotted the average movement 
between successive locations for all eight data sets, e.g. 


38 


FIGURE 3.1 - A comparison of the average distance between 
successive locations for the age/sex categories of adult females, 
adult males, juvenile females, and juvenile males, calculated for 
locations made 18-36 hours apart and those made after more than 


36 hours. 


6.0 


5.0 


4.0 


3.0 


Distance (km) 


2.0 


1.0 


(643) 


(1409) 


Time between locations 


18-36 hours apart 


[_] >36 hours apart 


(620) 
(1036) 

(1413) 
(3143) . 


| | 
ADULT JUVENILE 
FEMALES FEMALES 


39 


(399) 


(801) 


| 
JUVENILE 
MALES 


adult males, adult females, juvenile females and juvenile 
males for both the 18-36 hour data set and the >36 hour data 
set, plus or minus two standard errors of the mean (Fig. 3.1). 
Within the 18-36 hour data set, the distance moved between 
successive locations was relatively consistent, averaging 
around 1 km, the exception being the juvenile males, who 
averaged about 2.3 km, significantly greater than the other 
age/sex classes. Adult males showed much more variation in 
the >36 hour data set than in the 18-36 hour data set. This 
variation reflects the occasional long-distance movements of 
adult males, which often resulted in the animals not being 
located for several days. Juvenile males and juvenile females 
also made these long-distance movements but there was more 
consistency within these classes with respect to the distance 
traveled on such trips. — 


In general, otters tended to stay within a small area for 
an extended period and then suddenly move for a considerable 
distance. Tables 3.1 and 3.2 do not illustrate the distance 
which an individual otter can travel from one day to the next 
during these periods of rapid movement. To give a sense of 
this distance, we plotted the 20 longest movements (again 
using the UTM grid) made within 18 to 36 hours for otters 
belonging to each age/sex class (Fig. 3.2). The longest daily 
movements were 47.5 km for an adult male, 40.1 km for a 
juvenile female, 38.8 km for a juvenile male, and 17.5 km for 
an adult female. However, movements of more than 10 km a day 
were infrequent. 


Daily convex polygon areas 


Areas used by individual otters over a 24-hour period 
ranged from 10 to over 1000 ha (Table 3.3). Adult males were 
not included in the analysis of variance (Appendix 3.2) 
because of insufficient sample size. For the other three 
age/sex classes, there was no significant variation within 
age/sex classes but variation among classes was significant. 
Juvenile males tended to travel over a larger area than 
individuals of the other age/sex classes. 


Long-term movement patterns along the coast 


The average location along the California coast for each 
instrumented otter is shown in Fig. 3.3. Figures 3.4, 3.5, 
3.6, and 3.7 illustrate both the substantial degree of 
variation between individuals within age/sex classes in the 
extent to which they travel away from this average location 
and the way in which many otters tend to remain within small 
areas for extended periods of time and then suddenly move a 
much greater distance. 


40 


FIGURE 3. 


2 - A plot of the 20 longest trips made between 


successive locations that were 18-36 hours apart for the four 
age/sex categories of adult females, adult males, juvenile 


females, 


DISTANCE (KM) 


and juvenile males. 


50 


> 
ro) 


® ADULT FEMALES 

+ JUVENILE FEMALES 
° ADULT MALES 

4 JUVENILE MALES. 


ie 
(=) 


20 


1 3 5 U 9 11 13 15 17 19 


TRIPS IN ORDER FROM LONGEST TO SHORTEST 


41 


FIGURE 3.3 - The average locations of the instrumented sea otters 
along the California coast. 


Santa Cruz 


Monterey 


42 


FIGURE 3.4 - The general north-south movement pattern of 
individual adult males, in relation to their average north-south 
location, over the monitoring period. 


80 


DISTANCE (KM) 
: 


80 
OTTER 10 
160 
SOs 
OTTER 17 
OTTER 34 
I a TET a a ae Card ) 
1 JAN 85 "1 JAN 86 1 JAN 87 —- 1: JAN 88 


43 


FIGURE 3.5 - The general north-south movement pattern of 
individual adult females, in relation to their average north- 
south location, over the monitoring period. 


OTTER 6 , : . 


or; 


1 JAN 84 1 JAN 85 1 JAN 86 
OMMERES 
OMERR I 


160 Oineieat 


OIA 1S) 


OMEN We ay Ane 


OnMER ets 


Co 
O 


(@) 


OME 22 


DISTANCE (KM) 


Co 
O 


OTTER 25 H - 
sage COMMER 27 
SOUTH i" 
OTTER 31 
$$ ry aarti igen pe tte $$ 


OMA IS) 
ANA rp 


OTTER 36 
rrr prterni atrcgereeninarncmcrnerngan Q  e———— 


Sr) (reel) beara rere) eo ray Poa eae air an «UL a a 
1 JAN 85 1 JAN 86 1 JAN 87 1 JAN 88 


44 


FIGURE 3.6 - The general north-south movement pattern of 


individual juvenile females, in relation to their average north- 


south location, over the monitoring period. 


NORTH 
160 


0.) 
(e) 


DISTANCE (KM) 
0a 


160 
SOUTH 


OMER 23 


OMER: 57 
ee Cee 10 ee Ss 


OMNES OS) 


OTTER 40 


OTTER 42 


spa a eee 


OTTER 44 


ae OE a Te lame 


OTTER 45 


OTTER 46 
pe ee eee eee ee eee 


OTTER 47 
I tempt OO err rm 


Sa Maa aE Ta AGE TCE in TG ante ieaMlLa Gla nenae hae 


1 JAN 85 1 JAN 86 1 JAN 87 1 JAN 88 


45 


FIGURE 3.7 - The general north-south movement pattern of 
individual juvenile males, in relation to their average north- 
south location, over the monitoring period. 


OTTER 13 
NORTH 
160 
OTTER 30 
= ao 
x 
4 
Lu 
OQ 
Z 
a OTTER 35 
O 80 
OTTER 41 
160 
SOURG Wi tf 
OTTER 43 


SSS ST ee ee 
1 JAN 85 1 JAN 86 1 JAN 87 1 JAN 88 


46 


Four general movement patterns were apparent from these 
figures: 1) remaining within a small area throughout the 
study (for example, otters 11 and 46); 2) generally remaining 
within a small area but making occasional long-distance trips 
(otters 1 and 10); 3) shifting of centers of activity for 
extended periods of time (otters 17 and 30); and 4) frequent 
travel over long distances (otters 22 and 34). 


Adult males captured in both "male areas" (otters 1-4) and 
"female areas" (otters 7, 10, 17, 23, and 34) sometimes made 
long-distance movements. These were often relatively brief 
"trips" to a new location, followed by a return to the 


TABLE 3.3 - Area (ha) of daily home ranges based on data 
obtained during 24-hour watches in which locations were 
recorded once per hour. The areas were determined using the 
minimum area home range method (Hayne, 1949). AF = Adult 
Female, AM = Adult Male, JF = Juvenile Female, JM = Juvenile 
Male. 


OTTER AGE/SEX AREA 
NUMBER (HA) 
16 AF 10.33 
36 AF 13.08 
16 AF 22.52 
36 AF 34.10 
27 AF 58.23 
19 AF 68.97 
27 AF 227.59 
29 AF 1166.35 
34 AM 6.88 
34 AM” 12.75 
7 AM 223.37 
42 JF 31.69 
42 JF 32.88 
40 JF 107.03 
40 JF 118.81 
45 JF 119.29 
47 JF 165.36 
39 JF 212.53 
47 JF 213.75 
41 JM 221.24 
35 JM 258.52 
43 JM 302.14 
35 JM 359.62 
41 JM 379.49 
30 JM 511.40 
35 JM 570.23 
30 JM 625.82 
35 JM 666.38 


41 JM 759.32 


47 


original location. No seasonal pattern was apparent in these 
"trips". Adult females tended to be more sedentary. However, 
two of them (otters 14 and 22) often moved distances on the 
order of 10 km within a short time. 


Juvenile females tended to move more extensively than 
adult females and two of them (otters 39 and 40) made long- 
distance trips. Juvenile males tended to travel more than the 
other age/sex classes. 


Distance offshore 


Adult males and females were usually found relatively 
close to shore (Fig. 3.8). Adult females with pups were 
particularly close to. shore. There waS no apparent 
correlation between age of pup and distance offshore. The 
three females with pups in Figure 3.9 are (from left to right) 
otters 16, 27 and 36. The potential ages of their pups were 
in the ranges of 81 to 101, 6 to 17 and 0 to 37 days, 
respectively. Juveniles ranged farther from shore, and the 
tendency for the juvenile males to be much farther off-shore 
than individuals of the other age/sex classes was particularly 
striking. About four percent of our locations of juvenile 
males were over three km from shore. 


In general, otters tended to feed slightly closer to shore 
than they rested but this was not true of all individuals 
(Fig. 3.9). Distance offshore was not closely related to time 
of day, although juvenile males were often relatively close 
to shore about 6 to 7 a.m. (Fig. 3.10). 


Monthly harmonic mean ranges 


The average monthly deviations from the harmonic mean 
center of the locations of the instrumented otters presented 
in Table 3.4 are plotted in Figure 3.11. These data reflect 
the general monthly movement patterns of the four age/sex 
classes. Monthly harmonic mean home range sizes of juveniles 
(Fig. 3.11) appeared to increase during the peak pupping 
months of February, March, and April, and during the secondary 
peak in August, September, and October (see Fig. 2.1). 
However, when monthly Hp were grouped into these six months 
of peak pupping vs. non-peak pupping seasons, no significant 
seasonal differences were detected by analysis of variance 
(Appendix 3.3). This lack of statistical significance, 
particularly among juvenile males, is probably because we 
monitored only a few juvenile males, there was great variation 
within individuals, and they had very small home range sizes 
aligl JAKES (Haale Bo alal)) o 


48 


FIGURE 3.8 - The distribution of distances offshore for the four 
age/sex categories of adult females, adult males, juvenile 
females, and juvenile males. 


Adult Females Adult Males 


Juvenile Females Juvenile Males 


Percent of Locations 


300 1300 2300 300 1300 2300 


Distance (Meters) 


49 


FIGURE 3.9 - The average distance off-shore while resting and 
feeding for individual otters partitioned by five age/sex 
categories of adult females, adult females with pups, adult 
males, juvenile females, and juvenile males. 


Resting Juvenile Males 
MM Feeding 


ine) 


q 


Distance Offshore (km) 


lh Juvenile Females aan 
_| | = Females 
dt a oly Adult 

| j Females 

=| ae eee | with Pups 


50 


JUVENILE MALES 


JUVENILE FEMALES 


(AM) GHOHSYAO AONVLSIG 


FIGURE 3.10 - The average distance off-shore for the various 
hours of the day for juvenile males and juvenile females. 


16 20 24 


12 


TIME OF DAY 


51 


FIGURE 3.11 - The average distance deviation from the harmonic 
mean center of monthly home ranges for the four age/sex 


categories of adult females, adult males, juvenile females, and 
juvenile males. 


ADULT FEMALES 
JUVENILE FEMALES 


ADULT MALES 
JUVENILE MALES, 


~) 
O 


—s 
Oo 


ty} 


DEVIATION (KM) 


Bye: 


TABLE 3.4 - Average distance deviation (km) from the harmonic 
mean center of all locations for the four sex and age 
categories of adult males, adult females, juvenile females and 
juvenile males for each month of the year. AF = Adult Female, 
AM = Adult Male, JF = Juvenile Female, JM = Juvenile Male. 


MONTH AM AF JF JM 
MEAN N MEAN N MEAN N MEAN N 

JAN 0.207 7 0.623 13 0.810 10 1.418 5 
FEB 0.079 7 0.472 13 1.170 10 2.288 6 
MAR 0.059 10 0.478 17 0.984 10 1.433 5 
APR 0.080 12 0.371 20 0.791 9 1.527 6 
MAY 0.081 12 0.454 20 0.653 10 1.677 5 
JUN 0.145 12 0.354 20 0.396 10 0.993 5 
JUL 0.138 12 0.459 21 0.735 9 1.143 5 
AUG 0.131 10 0.404 21 0.532 9 0.702 5 
SEP 0.099 9 0.328 20 0.368 9 1.638 6 
OocT 0.112 9 0.315 23 0.436 9 1.559 6 
NOV 0.336 8 0.27 21 0.389 9 1.038 7 
DEC 0.031 6 0.188 18 0.367 9 0.816 6 


There was significant variation in monthly -Hp among 
individuals of both adult sex classes but not for juveniles, 
whether seasonal effects were considered (Appendix 3.3) or not 
(Appendix 3.4). When seasonal effects were disregarded, 
variation of Hp among age/sex classes was much greater than 
variation among individuals within age/sex classes and was 
Significant (Appendix 3.4). Adult males tended to move within 
a small area; adult females utilized areas slightly larger 
than those of adult males. Juveniles of both sexes traveled 
over larger areas than adults throughout the year, and 
juvenile males used strikingly larger areas than individuals 
of the other age/sex classes. 


Distance between extreme locations 


The average and two extreme locations, and the distance 
between these two extreme locations, for each otter are shown 
in Table 3.5 and plotted in Figure 3.12. Unlike Hp, the 
average distance between extreme locations is extremely 
sensitive to a few long-distance movements. Analysis of 
variance indicated that differences among age/sex classes were 
Significant (Appendix 3.5), with juvenile males having the 
largest distances and adult females the smallest (Table 3.6). 
Distances for adult males, which occasionally take long- 
distance trips, were greater than those for adult and juvenile 
females. 


These data are consistent with our other analyses of the 
movement data, indicating that juvenile males are the most 
extensive travelers. 


53 


FIGURE 3.12 - The distance between extreme locations for all of 
the instrumented sea otters in California, partitioned by the 
four age/sex categories of adult females, adult males, juvenile 
females, and juvenile males. 


280 


aa) Juvenile Males 


200 Adult Males 


160 


Juvenile Females 
120 


Distance (km) 


Adult Females 


80 F 


40 


15 36 1133 31 19 27 6 2516 9 1422 23 17 4171034 46 374742 29444539 40 41 35 43 13 30 


Otter Number 


54 


TABLE 3.5 - The average location along the five fathom line, and the 
northern-most and southern-most location along this line, that were 
recorded during the period of monitoring for each instrumented sea 


otter. AF = Adult Female, AM = Adult Male, JF = Juvenile Female, JM = 
Juvenile Male. 
OTTER AGE/ MEAN NORTHERNMOST SOUTHERNMOST Distance 
NO. SEX LOCATION LOCATION (A) LOCATION (B) Between 
A&B(km 
15 AF COAST GALLERY GRIMES POINT PARTINGTON POINT 5.0 
36 AF SALMON CREEK REDWOOD GULCH RAGGED POINT 10.0 
11 AF FALSE SUR VENTURA ROCKS BIG SUR RIVER al GS) 
33 AF PARTINGTON PT GRIMES POINT ESALEN 15.5 
31 AF BUCK CREEK TORRE CANYON DOLAN ROCK 15.5 
19 AF WRECK BEACH FALSE SUR GRIMES POINT 18.5 
27 AF DOLAN ROCK ANDERSON CANYON LUCIA 21.5 
6 AF ARROYO LAGUNA CHINA GULCH LITTLE PICO CRK 21.5 
25 AF ANDERSON CR GRIMES POINT SQUARE BLACK ROCK 22.5 
16 AF PFEIFFER BEACH LITTLE SUR RVR GRIMES POINT 23.5 
9 AF BUCK CREEK PFEIFFER POINT DOLAN ROCK 26.5 
14 AF BUCK CREEK GRIMES POINT LUCIA 30.0 
22 AF ROCKY POINT POINT PINOS ANDERSON CANYON 86.0 
2 AM MORRO ROCK CAYUCOS POINT MORRO BAY 12.5 
3 AM MORRO ROCK ARROYO LAGUNA MORRO BAY 54.5 
1 AM MORRO ROCK ARROYO LAGUNA MORRO BAY 56.0 
7 AM POINT SUR POINT PINOS COOPER POINT 60.5 
4 AM MORRO ROCK RAGGED PT INN MONTANA DE ORO 79.5 
17 AM PFEIFFER BEACH MOSS LANDING COAST GALLERY 112.0 
10 AM PFEIFFER BEACH SOQUEL POINT TORRE CANYON 140.5 
34 AM SOBERANES PT GREYHOUND ROCK BUCK CREEK 181.0 
46 JF ARROYO LAGUNA BECKETS REEF SAN SIMEON POINT 13.0 
37 JF ARROYO LAGUNA BECKETS REEF SAN SIMEON POINT 15.0 
47 JF POINT JOE MUSSEL POINT SUNSET POINT 18.0 
42 JF RAGGED PT INN COUNTY LINE ARROYO LAGUNA 22.0 
29 JF DOLAN ROCK COAST GALLERY LUCIA 30.5 
44 JF BECKETS REEF GORDA ARROYO LAGUNA 34.5 
45 JF PLASKETT ROCK DOLAN CREEK BECKETS REEF 62.5 
39 JF SODA SPRING CR POINT SUR PIEDRAS BLANCAS 103.5 
40 JF CRUZ ROCK VENTURA ROCKS SAN SIMEON PT 120.0 
41 JM CRUZ ROCK REDWOOD GULCH SAN SIMEON CREEK 44.0 
35 JM COUNTY LINE PACIFIC VALLEY BECKETS REEF 64.0 
43 JM BECKETS REEF SALMON CREEK SAN SIMEON PT 82.0 
13 JM CARMEL BEACH SUNSET BEACH RAGGED PT INN 191.5 
30 JM JADE COVE SOBERANES POINT PIEDRAS BLANCAS 258.0 


55 


TABLE 3.6 - The average distance (km) between extreme 
locations, measured along the five fathom line, for the four 
sex and age categories of adult males, adult females, juvenile 
females and juvenile males. AF = Adult Female, AM = Adult 
Male, JF = Juvenile Female, JM = Juvenile Male. 


AGE/SEX MEAN N VARIANCE RANGE 
(KM) (KM) 
AF 23.65 13 368.28 5- 86 
AM 97.71 7 2042.204 54-181 
JF 46.56 9 1426.024 13-120 
JM 127.90 5 6840.04 44-228 


Because these data are based on the most extreme distances 
moved, they also show the tendency for adult males to make 
fairly long-distance trips. The one exception in Figure 3.12 
is male number 2. However, he was monitored for only 100 days 
before he disappeared and it seems probable that he simply 
happened not to make a long trip within this relatively short 
monitoring period. 


DISCUSSION 


Radiotelemetry studies have shown that sea otter home 
ranges in Alaska consist of several heavily used areas 
connected by travel corridors (Garshelis and Garshelis, 1984; 
Chapter 8). The results of our studies in California agree 
with this general picture: otters tended to stay within a 
small area for an extended period and then suddenly move for 
a much greater distance. 


Distance between successive daily locations 


Otters of all age and sex classes were usually found 
within one or two km of their location on the previous day. 
There are no comparable data on the locations of individuals 
from one day to the next in the literature but Ribic (1982) 
measured the distance between successive locations of the same 
individual at 3.4-5 hour intervals. Some of Ribic's sample 
sizes were extremely small. However, if we compare our data 
to those data where she had samples of at least 10 locations, 
the distances reported in the two studies are similar. 


This similarity suggests that sea otters have a tendency 
to move fairly quickly and directly between locations where 
a considerable amount of time is devoted to more sedentary 
activities such as resting or feeding. Such direct movements 
have been reported (Loughlin, 1980) and we often observed them 
during course of our study. 


56 


The longest distance between successive daily locations 
recorded in our study was about 47 km. This is similar toa 
movement of 48 km in 22 hours mentioned by Ribic (1982). 
Kenyon (1969) estimated that sea otters can swim at sustained 
speeds of about 2.5 km per hr., thus the long daily movements 
we recorded probably involved nearly constant swimming. 


Long-distance movements 


Other investigators have documented that sea otters make 
occasional long-distance movements (Kenyon, 1969; Ribic, 1982; 
Garshelis and Garshelis, 1984), but these were thought to 
largely reflect seasonal movements of males between summer and 
winter ranges. Our daily monitoring revealed that individual 
otters of all age/sex classes make a surprising number of 
long-distance movements at all times of year. 


Although some individuals remained within a small area for 
extended periods, e.g. otter 7, an adult male that remained 
within a very small area for 18 months (Fig. 3.4), and otter 
15, an adult female that visited only 5 km of coastline during 
the entire monitoring period (Fig. 3.5; Otter 15), it became 
evident to us that the longer an individual was monitored, the 
more likely it was to travel a significant distance. Only a 
few "trips" have been previously documented in the literature, 
probably because such long distance movements are less likely 
to be detected in tag-resight studies and radio telemetry 
studies of short duration. Although the reasons for these 
“trips" are unknown, it seems likely that they vary. Adult 
males may be seeking mating opportunities or areas with high- 
quality food resources. Juvenile males may be displaced by 
older males and may seek the company of other young males as 
well as high-quality food resources. Females, both juvenile 
and adult, that take extended "trips" may be looking for areas 
where they can become resident. 


Seasonal patterns 


We were unable to detect a significant seasonal pattern 
in the frequency of long-distance movements for any age/sex 
class, including adult males. We also looked for seasonal 
trends in the size of monthly harmonic mean home ranges. This 
method eliminates the possibility of greatly overestimating 
the area utilized due to a few long-distance movements. Once 
again, we failed to find any seasonal pattern in the size of 
the area used by adult males. However, we monitored only a 
small number of adult males. It also seems possible that 
predictable seasonal movements of adult males occur in some 
but not all areas. Juveniles appeared to show a seasonal 
pattern in harmonic mean home range size, with the peaks 
during the peaks of parturition period and pup dependency. 


57 


Although this effect was not statistically significant in our 
data set, it may nevertheless be a real phenomenon. 


Variation within and among age/sex classes 


We found substantial variation in movement patterns among 
individuals within all age/sex classes. For example, most 
adult females tended to be relatively sedentary but two of the 
15 we monitored often traveled for considerable distances. 
In spite of this extensive individual variation, some 
generalizations on the movement patterns characteristic of the 
different age/sex classes are possible. 


Measures that minimize the effects of long-distance 
movements, such as the distance between successive daily 
locations and monthly harmonic mean home ranges, indicate 
that, over the short-term, adult males tend to utilize smaller 
areas than adult females. This agrees with the findings of 
Loughlin (1980) and Ribic (1982) that adult females have 
larger home ranges than territorial or "resident" males. 
However, measures that are sensitive to long-distance 
movements, such as the distance between locations recorded at 
intervals of more than one day and the DBEL, show that adult 
males are more likely than adult females to make long-distance 
movements and thus tend to visit greater lengths of coastline 
over the long-term. 


On a daily basis, movements of juvenile females were 
similar to those of adult females. However, juvenile females 
were more likely than adult females to make long-distance 
movements. Juvenile males tended to move for greater distances 
than the otters belonging to other age/sex classes on both a 
short= and long-term basis. 


Distance offshore 


It has been reported that otters in California rarely 
travel far offshore (Loughlin, 1980; Ribic, 1982). Estes and 
Jameson (1988) found that 90% of the otters observed from 
shore were within 600 m of the coast and that the probability 
of a shore-based observer sighting an otter was constant over 
observation distances of 50-850 m from shore, although it 
declined to zero by 1300 m. The majority of our locations of 
adult males and females and juvenile females were within 800 
m of shore; however, the majority of our locations of juvenile 
males were over 800 meters from shore and over half of them 
were more than 1300 m (Figure 3.8). Although Estes and 
Jameson (1988) concluded that “few otters occurred at 
distances from shore beyond the observers' viewing ranges", 
our juvenile males were frequently located at distances beyond 
the viewing range of a shore-based observer. 


58 


Furthermore, our data underestimated the extent to which 
these juvenile males traveled offshore. The signal from an 
otter must be detectable from at least two shore locations to 
enable estimation of its distance offshore through 
triangulation. The signals of the juvenile males were often 
only faintly detectable from a single shore location; their 
distance offshore could not be estimated on these occasions. 
In addition, we were unable to locate the juvenile males a 
much larger proportion of the time than the other otters, 
probably because they often moved so far offshore that we 
could not receive their radio signals. 


Home range size 


Estimates of the area utilized by an individual otter are 
greatly influenced by whether or not long-distance movements 
by that individual were detected during the study period, 
whether or not they were included in the analysis if detected, 
and the specific method used to estimate the area utilized. 
The minimum convex polygon method is appropriate for measuring 
the areas used by otters in the intervals between long-- 
distance movements; we used it to estimate the area used 
within a 24-hour period. 


TABLE 3.7 - Comparison of home range areas (ha) for sea otters 
in California and Alaska calculated in this and previous 
studies. 


AGE/SEX CLASS _ LOCATION AREA (HA) SOURCE 


Nonterritorial males AK 400-1440 Garshelis&Garshelis 1984 
Territorial males CA 18-58 Loughlin 1980 

Resident males CA 80-460 Ribic 1982 
Nonterritorial males CA 29-138 Loughlin 1980 

All males CA 80-2980 Ribic 1982 

Adult males CA 7-223 This study 

Juvenile males CA 221-759 This study 

All females AK 20-960 Garshelis&Garshelis 1984 
All females CA 28-198 Loughlin 1980 

Resident females CA 470-680 Ribic 1982 

All females CA 470-2110 Ribic 1982 

Adult females CA 10-1166 This study 

Juvenile females CA 32-214 This study 


ape 


Our estimates of the area used on a daily basis overlap 
estimates of home range size over longer periods made by other 
investigators (Table 3.7). This suggests that, in the 
intervals between long-distance trips, otters tend to travel 
on a daily basis over much of the area they are currently 
using. Because of the variety of methods used to calculate 


59 


the area utilized and the different periods of time included 
in the analyses, these estimates show great variance of home 
range size. 


Early investigators suggested that there should be a 
relationship between home range size and metabolic rate 
(McNab, 1963). Because sea otters are known to have a high 
metabolic rate and are unable to fast for long, there may be 
a relationship between area used and prey availability in sea 
otters. However, such a relationship would be impossible to 
detect unless there is some standardization in methods of data 
collection and analysis. Greater consistency could be 
achieved by considering a rather short period, such as 24 
hours, and excluding those periods in which individuals spent 
much of their time traveling between locations. Such 
comparisons might not be appropriate between areas in which 
the near-shore communities differed greatly. 


The distance between extreme locations has been used as 
an index of home range size, particularly in Alaska (Garshelis 
and Garshelis, 1984; Chapter 8). In general, our values for 
adult males in California are greater than those reported for 
Alaska. There have been no reports of DBEL's over 100 km for 
instrumented otters in Alaska; three of our adult males in 
California had DBEL's of 112, 141, and 181 km. (Table 3.5). 
However, most of the Alaska data are from Garshelis and 
Garshelis (1984), who monitored instrumented otters for 
relatively short periods during the non-winter months, and it 
seems clear that this index, like measures of home range area, 
tends to increase with the length of time individuals are 
monitored. Although Garshelis and Garshelis (1984) did not 
document long-distance movements by any of their instrumented 
adult males, they observed that four tagged males moved about 
100 km between their summer territories and a "male area" 
occupied during other times of the year. This suggests that 
the greatest distances visited by males in California and 
Alaska may be more similar than current telemetry data 
indicate. 


DBEL's for adult females seem to be comparable in the two 
areas. Ours ranged from 5 to 86 km and those reported by 
Monnett and Rotterman for Alaskan females in Chapter 8 range 
from 28 to 80 km. Again, those reported by Garshelis and 
Garshelis (1984) were smaller, the largest being about 20 km. 


Our data for juveniles are also similar to those of 
Monnett and Rotterman from Alaska (Chapter 8), in that males 
tended to move greater distances than females. Our DBEL's for 
juveniles of both sexes are much greater than for Alaska but 
the data are not entirely comparable. The Alaska data 
represent the movements of individuals instrumented as 
dependent pups and monitored through the early period of 


60 


independence, for a maximum of 21 months after weaning. Our 
data represent the movements of individuals captured after 
weaning. Some of these individuals appeared to be quite young 
but others were estimated, based on the cementum lines in 
their premolars, to be up to two years of age. In Alaska, the 
longest distance moved by a juvenile female was less than 50 
km and that by a juvenile male was approximately 120 km 
(Chapter 8). Our longest distances were 120 km for a juvenile 
female and 258 km for a juvenile male. Although our juveniles 
moved farther, they were older. 


In Alaska, juvenile males tended to move relatively long 
distances, which took them out of the areas occupied by 
reproductive adults, within a few weeks after weaning. It is 
not known if young males make similar movements in California 
soon after weaning, but the juvenile males we monitored 
remained in the area occupied by breeding adults for most of 
the study period. Many of them did associate with a male 
group that formed near Ragged Point, well within the "female 
area" for several months. Two of them finally moved into the 
"male area" in the southern part of the California range 
towards the end of the monitoring period. 


Sex differences in dispersal patterns 


Although the timing may differ, sex differences in 
dispersal patterns appear similar in California and Alaska. 
Sea otters exhibit the dispersal pattern typically found in 
polygynous mammals (Greenwood, 1980): juvenile males tend to 
move farther than females. The juvenile males ultimately join 
male groups, usually within "male areas", while juvenile 
females remain within their natal "female area". 


Aggression from territorial males may play a role in the 
initial departure of the juvenile males from the kelp beds 
frequented by breeding adults. However, it is likely that, 
on the average, juvenile males ultimately benefit from this 
long distance dispersal in terms of increased reproductive 
success. 


The extensive travels of juvenile males probably enable 
them to become familiar with a large area; this may be an 
advantage later in life, when, as young adults, they return 
to a "female" area and search for an available territory 
(Loughlin, 1980, Ribic, 1982). By participating in the 
frequent social interactions, such as various forms of play- 
fighting, that occur in male groups, juvenile males may gauge 
their strength relative to other males and develop the 
fighting skills needed to acquire a territory. Because male 
groups are often located in areas that have been occupied by 
sea otters for a shorter period than the "female" areas, 
juvenile males that join these groups may tend to derive 


61 


nutritional benefits. Increased prey availability during the 
male's growth period could have important benefits later in 
life, as the sexual dimorphism characteristic of sea otters, 
with males considerably larger than females, suggests that 
large body size is likely to be an advantage to males when 
fighting with other males. 


In general, "female areas" have been occupied by otters 
for many years, and prey availability there is reduced 
compared to the "male areas". By remaining within these 
areas, juvenile females are forced to compete with larger, 
older otters for good foraging locations in an area that has 
already been exploited for some time. Our California data on 
time budgets and activity patterns (Chapter 4) and feeding 
patterns (Chapter 5) indicate that juvenile females tend to 
be poor competitors and their survival rates are rather low 
compared to other age/sex classes in both California (Chapter 
2) and Alaska (Chapter 8 and 9). 


Why then, do juvenile females remain in these "female 
areas"? Theory suggests that males will tend to behave so as 
to maximize mating opportunities and females so as to maximize 
the food resources available to them and their offspring. 
Juvenile females may ultimately benefit from acquiring a 
detailed familiarity with the distribution and availability 
of prey within a small area. In many mammals in which young 
females tend to remain within their natal area, female young 
may benefit by acquiring all or part of their mother's home 
range-~ The extent to which this occurs in sea otters is 
unknown. If adult females are intolerant of strange females, 
juvenile females that disperse from the "female" areas and 
subsequently return may have reduced chances of acquiring a 
good quality home range within these areas. 


Regardless of the reasons for its occurrence, the sex 
difference in movement patterns of juvenile sea otters may 
have important consequences for sea otter population dynamics 
by decreasing survival of juvenile females (Chapter 2). 


LITERATURE CITED 


Burt, W. E. 1943. Territoriality and home range concepts as 
applied to mammals. J. Mammal. 24:346-352. 


Estes, J.A. and R. J. Jameson. 1988. A double-survey 
estimate for sighting probability of sea otters in 
California. J. Wildl. Manage. 52: 70-76. 


Garshelis, D: L. and J. A. Garshelis. 1984. Movements and 
management of sea otters in Alaska. J. Wildl. Manage. 48: 
665-678. 


62 


Greenwood, P. J. 1980. Mating systems, philopatry, and 
dispersal in birds and mammals. Anim. Behav. 28: 1140-- 
1162. 


Hayne. 1949. Calculation of size of home range. J. Mammalogy 
30: 1-18. 


Kenyon, K. W. 1969. The sea otter in the eastern Pacific 
Ocean. North American Fauna. No. 68. Bureau of Sport 
Fisheries and Wildlife. U.S. Government Printing Office, 
Washington, D. C. 352 pp. 


Loughlin, T. R. 1980. Home range and territoriality of sea 
otters near Monterey, California. J. Wildl. Manage. 44: 
576-582. 


McNab, B. 1963. Bioenergetics and the determination of home 
range size. Am. Nat. 97: 133-140. 


Nees, Do | Siro 1966. Statistical analysis for area 
distributions. Regional Science Research Institute 
Monograph Series 2. Philadelphia, PA. 172 pp. 


Ribic, Coy Acar 1982). Autumn movement and home range of sea 
otters in California. J. Wildl. Manage. 46: 795- 801. 


Schneider, K. B. 1978. Sex and age segregation of sea otters. 
= Fed. Aid in Wildlife Restoration Project W-17-4 and W-17- 
5. Final Report. Alaska Department of Fish and Game. 


45 pp. 


63 


CHAPTER 4 


TIME BUDGETS AND ACTIVITY PATTERNS OF CALIFORNIA SEA OTTERS 


K. RALLS AND D. B. SINIFF 


NOVEMBER 30, 1988 


64 


INTRODUCTION 


The California sea otter population was reduced to a small 
number of animals by fur hunters in the 18th and 19th 
centuries. The remnant population grew at approximately five 
percent per year until sometime in the mid 1970's, when growth 
apparently ceased (Ralls, et al., 1983; Estes and Jameson, 
1983; Estes, et al., 1986). A central question relating to 
the dynamics of the California sea otter population, and hence 
the development of a model for this population, is whether or 
not the recent lack of growth is due primarily to density 
independent factors, such as entanglement in gill nets 
(Wendell, et al., 1986) or attacks by sharks (Ames and 
Morejohn, 1980), primarily to density dependent factors, such 
as competition for food or other resources (Miller, 1980), or 
to both. However, the most recent surveys suggest that the 
population may have resumed growth (Jameson and Estes, 1988). 


Several authors have proposed that time budgets might be 
useful as indicators of population status, assuming that food 
is an important limiting resource (Eberhardt, 1977; Estes, et 
al., 1982; Estes, et al., 1986). The prey available to sea 
otters varies with location and the length of time the area 
has been occupied by sea otters. Typically, most of the 
population initially consumes large prey items with high 
caloric content; as the availability of such prey decreases, 
the diet of the population diversifies to include smaller 
items and less preferred species (Estes, et al., 1981; 
Garshelis, et al., 1986). An otter should have to spend more 
time foraging to obtain a constant amount of energy in 
habitats with reduced abundance, size, or quality of prey than 
in those where high-quality food is abundant. Estes, et al., 
(1982) contrasted foraging patterns between two islands of the 
Aleutians; one where otters have existed for many years 
(Amchitka) and another that was recently colonized (Attu). 
The difference in time spent foraging (as determined by visual 
observations during the day) was dramatic as otters at 
Amchitka foraged during 55% of the daylight hours while those 
at Attu foraged only 17% of the daylight hours. Garshelis, 
et al., (1986) working in Prince William Sound, Alaska, 
assumed food was more abundant in areas that were recently 
invaded by sea otters and, under this assumption, confirmed 
- the predicted relationship between food availability and the 
proportion of time devoted to foraging: otters spent about 10 
percent more time foraging at Green Island, where sea otters 
had been present for many years, than in Nelson Bay, an area 
only recently reoccupied by sea otters where high quality prey 
items were abundant. 


Time budgets in California have been determined primarily 
by visual observations of unidentified otters. The most 
recent study (Estes, et al., 1986) concluded that food was 


65 


probably not limiting further growth of the California 
population because sea otters there apparently foraged less 
than at Amchitka, there is unoccupied habitat at both ends of 
the range, and other mortality sources (e.g. gill nets) have 
been impacting the population. The use of radio-telemetry to 
collect time budget data has a number of advantages 
(Garshelis, et al., 1986) and is helpful in evaluating the 
many sources of mortality. Data can be collected over the 
entire 24-hour period and the time spent foraging can be 
estimated for individuals and age/sex classes, and direct data 
on sources of mortality are sometimes available. 


One potential limitation of telemetry data is that the 
number of individuals monitored is often small. Thus, when 
the assumption is made that the instrumented individuals are 
representative of the population, the potential for bias, 
brought about by small sample size, exists. However, the 
ability to evaluate individual variation and compare age/sex 
groups allowed by this method greatly enhances’ the 
understanding of sources of variation. Further, visual scan 
samples also have biases because of differences in the spatial 
distribution among age/sex groups. For example, we found that 
juvenile males are often too far offshore to be seen, even 
with spotting scopes (Chap.3). 


There have been two studies of sea otter time budgets in 
California based on telemetry. However, Ribic (1982) did not 
separate feeding from other kinds of activity and Loughlin's 
(1980) sample was small. Our data provide more extensive 
information on time budgets and activity patterns of 
California sea otters based on telemetry data. 


METHODS 


Radiotelemetry is particularly useful for collecting time 
budget and activity data on sea otters because radio signals 
are not transmitted through sea water. Because of this 
characteristic, the radio signal pattern varies according to 
an otter's activity. Three general categories of activity 
have been distinguished by listening to the radio signal from 
an otter: resting, feeding, and "other" (Loughlin, 1980; 
Garshelis, et al., 1986). The radio signal from an otter 
resting on the surface of the water is constant. Most 
foraging otters alternately dive to obtain prey and return to 
the surface to consume their catch, or if they were 
unsuccessful, breathe before diving again. Thus the radio 
signal of a feeding otter is usually a characteristic pattern 
of alternating periods of signal and silence. When otters are 
engaged in activities other than resting or feeding, such as 
swimming, vigorous grooming, or social interactions, the 
signal is variable, with the strength of each signal depending 
upon the orientation of the otter's body with respect to the 


66 


surface of the water. We called this category "other". When 
the otter could be seen, specific behaviors such as swimming, 
active grooming, or social interactions were recorded. These 
behaviors were later combined into the "other" category so 
that the data collected visually would be comparable to data 
based solely on the radio signal. A potential bias in 
California was the very shallow feeding within the kelp canopy 
that might be confused with the "other" category. This was 
evaluated by comparisons with visual observations. 


In preliminary analyses, we distinguished a fourth 
activity category, "unknown", not used in other telemetry 
studies of 24-hour activity (Loughlin, 1980; Garshelis, et 
al., 1986), for those data points where the observer could 
tell from the radio signal that the otter was active but was 
not certain whether or not it was feeding, and those in which 
the radio signal was so weak that the activity could not be 
identified. 


We directly compared activity data collected visually with 
those obtained only by listening to the radio signal without 


seeing the otter. Two observers, one using each method, 
simultaneously recorded data on the same animal at five-minute 
intervals. These data were collected by several different 


pairs of observers on a variety of individual otters. 


A team of 3 to 4 people, each taking a 6 to 10 hour 
shift, monitored the activity of individual otters for one or 
more 24-hour periods. Most data-collection periods were 
either 24 or 48 hours in length (Appendix 4.1). Data were 
recorded at 10-min intervals. We developed these methods by 
trying various alternatives on otters 1-4 in Morro Bay. Our 
activity data on these otters are thus not directly comparable 
to the majority of our data and are not included in our 
analyses of time budgets. Thus, the data that were included 
in our analyses of time budgets mostly came from the center 
of the California sea otter range. Fig. 4.1 shows the 
relative locations of the watches. Considering the locations 
of these watches and the number of individuals that we were 
able to monitor for each age/sex class, we feel that we had 
a representative sample of sea otter activity in areas that 
had been occupied for a long period of time. 


RESULTS 


Activity in relation to time of day 


Otters of all age/sex classes tended to be active and feed 
for a large proportion of time during the late afternoon and 
early evening but there were differences in the activity 
patterns of various groups. For adult males in Morro Bay, we 
plotted only periods of activity and inactivity: these animals 


67 


FIGURE 4.1 -- The locations along the California coast of 
watches for collecting time budget data on sea otters 
instrumented with radio transmitters. 


HALF MOON BAY 


0310 20 40 60 
| 


SCALE IN KILOMETERS 
o.6S (10 20 30 40 
eee! 


SCALE IN MILES 
> ANOQ NUEVO ISLAND 


SANTA CRUZ 
OQUEL POINT 


MONTEREY MOSS LANDING 
BAY 
/SEASIOE 
\AMONTEREY—PACIFIC GROVE 
POINT PINOS 


POINT LOBOS 
-YANKEE POINT 


BIXBY CREEK 
POINT SUR 


—SIG CREEK 
LOPEZ POINT 


CAPE SAN MARTIN 
SALMON CREEK 


POINT PIEDRAS BLANCAS 
=<—SAN SIMEON 


Location of watch Adults Juveniles 
Females Males Females Males 


Santa Cruz - Moss Landing 
Moss Landing - Point Sur 
Point Sur - Cape San Martin 


Cape San Martin - San Simeon 


68 


had a clearly bimodal activity pattern with a second major 
activity period in the morning, peaking about 8 a.m. (Fig. 
4.2a). A similar early morning peak, in the percentage of 
time feeding, was noticeable in the juvenile males (Fig. 
4.3a), the adult females (Fig. 4.4a), and the adult females 
with pups (Fig. 4.4b). The evening feeding peak for the adult 
males in the Big Sur area was so large that their pattern was 
almost unimodal (Fig. 4.2b). The juvenile females had a much 
broader feeding peak than any of the other groups, feeding 
about 50% of the time even in the middle of the day when the 
other age/sex groups rested (Fig. 4.3b). In contrast to the 
other groups, the juvenile females often rested during the 
night, from midnight to seven a.n. 


There was a good deal of variation between individuals, 
but also variation from one day to the next, at the same time 
of day, for each individual. Our sample sizes, at a specific 
time of day, were not large enough to statistically test the 
significance of these sources of variance, since most 
individuals had only three to four data points for a given 
hour of the day. 


Comparison of visually and telemetrically collected data 


Our data provide the first formal comparison of visual and 
telemetric estimates of otter activity in which feeding was 
distinguished as a separate category, as neither Loughlin 
(1982) nor Garshelis, et al., (1986) undertook such a 
comparison. Ribic (1982) found good agreement between visual 
and telemetric data when only active and inactive were scored. 


We found the highest agreement between the methods when 
an otter was resting: when the visual observer indicated that 
an otter was resting, the telemetric data agreed 93 percent 
of the time (Table 4.1). The telemetric observer never scored 
resting as feeding or "unknown" but occasionally scored it as 
"other". This error tended to occur when sea conditions such 
as high swell caused the otter to move about even though it 
was resting. 


When the visual observer indicated that an otter was 
feeding, the telemetric data agreed 88 percent of the time 
(Table 4.1). The telemetric observer rarely scored feeding 
as resting or "other" but did sometimes indicate that the 
otter's activity was "unknown". This tended to occur when an 
otter was feeding without making regular dives. Otters in 
California can forage in kelp for prey items such as kelp 
crabs or obtain small items such as mussels from rocks without 
remaining submerged for more than a few seconds. 


When the visual observer indicated that an otter was 
engaged in "other" activities, the telemetric data agreed only 


69 


FIGURE 4.2 -- The percent of time that adult male sea otters 
spent in various activities at the various hours of the day. 
Males in the Morro Bay area of California are shown in the 
upper graph; the males located in the Big Sur area are shown 
in the lower figure. 


100 


O Inactive 
4 Inactive 


100 


GC Resting 


4 Feeding 
80 © Other 


Percent of time 


40 


0 4 8 12 jen) 20 24 
Time of Day 


70 


FIGURE 4.3 -- The percent of time that instrumented juvenile 


sea 


otters in California spent resting, feeding and in 
"other" activity at the various times of the day. 


Data for 


juvenile males are shown in the upper graph and those for 
juvenile females in the lower graph. 


Percent of Time 


100: 


80 


60 


40 


20 


100 j 
Resting 


Feeding 


80 Other 


60 


40 


20 


0 4 8 12 16 20 24 
Time of day 


71 


FIGURE 4.4 -- The percent of time that adult female sea 
otters in California spent resting, feeding and in "other" 
activity at the various hours of the day. Data for adult 
females without pups are shown in the upper graph and data 
for adult females with pups in the lower graph. 


100 
80 | 


60 | 


100 


Percent of Time 


80 oO Resting 
4 Feeding 
O Other 


Time of day 


72 


TABLE 4.1 - A comparison between activity data obtained 
visually with those obtained using the quality of the 
telemetric signal. The numbers in this table represent the 
number of 5-minute sampling periods where the activity of an 


otter was determined. These observations were taken 
simultaneously by two independent observers. AF = adult female 
without pup; AFP = adult female with pup; JF = juvenile 


female; JM = juvenile male. 


NUMBER OF FIVE-MINUTE SAMPLE PERIODS 


Visual Percent 
Observer Telemetry observer - Agree. 

DATE REST REST FEED OTHER UNKNOWN 

31-May-85 AF 38 36 2 

04-Feb-87 AF 32 31 1 

05-Feb-87 AFP 15 15 

25-Feb-87 AF 7 6 1 

27-Feb-87 JM 1 1 

27-Feb-87 JF 24 20 4 

03-Mar-87 AF 1 1 

30-Mar-87 AFP 12 12 

TOTAL 130 121 93 

FEED REST FEED OTHER UNKNOWN 

31-May-85 AF 33 33 

14-Feb-87 AFP 8 1 6 1 

25-Feb-87 AF 19 11 8 

27-Feb-87 JM 3 3 

27-Feb-87 JF 3 3 

02-Mar-87 JF 43 41 2 

03-Mar-87 AF 5 5 

07-Apr-87 AF 8 8 


TOTAL 122 107 88 


OTHER REST FEED OTHER UNKNOWN 


31-May-85 AF 2 1 i! 

14-Feb-87 AFP 2 2 

25-Feb-87 AF 2 1 1 

27-Feb-87 JM 10 9 1 

27-Feb-87 JF 7 2 5 

30-Mar-87 AFP 1 1 

TOTAL 24 ALS) 63 


63 percent of the time. Activities in the "other" category 
were rarely recorded as feeding by the telemetric observer but 
were sometimes scored as resting or "unknown". "Other" was 
most commonly recorded as "resting" when the otter was 
grooming fairly vigorously without submerging the main part 
of its body. 


73 


TABLE 4.2 - A comparison between time budgets calculated from 
observing activity visually and judging activity using the 
quality of telemetric signal. 


Number of 5-min periods Percentage of time 

Visual Telemetry Visual Telemetry 
REST 130 121 47 44 
FEED 122 108 44 39 
OTHER 24 25 9 9 
UNKNOWN (0) 16 (0) 6 


A comparison of the overall time budgets based on the 
visual and telemetric data collected simultaneously on the 
same otters (presented in Table 4.1) indicated that the 
telemetric data underestimated resting by 3 percent and 
feeding by 5 percent for this particular data set (Table 4.2). 
These data provide only an indication of the differences 
between estimates based on the two methods. The magnitude of 
the differences will vary with such factors as the sea state 
during the period of data collection, the extent to which the 
otter being studied feeds without making regular dives, and 
the individuals recording the data. Although these data 
suggest that we underestimated the proportion of time spent 
feeding, it would be inappropriate to simply increase our 
estimates. of the overall proportion of time spent feeding by 
five percent because of this variation and because our 
estimates are based on a combination of visual and 
telemetrically collected data. 


Time budgets 


Using the entire data set collected during the 24 to 72- 
hour watches (Appendix 4.1), we assessed the extent to which 
different ways of handling the "unknown" data might affect our 
results by comparing four ways of treating them. These were: 
1) including them in the "other" category, as in our previous 
reports (Ralls, et al., 1985; Siniff and Ralls, 1986); 2) 
keeping them as a separate category; 3) excluding them from 
the analysis; and 4) including them in the "feeding" category 
to get an estimate of the maximum possible feeding time, as 
our data suggest that much of the "unknown" category may be 
feeding. Because the "unknown" category was small, ranging 
from 1 to 7 percent of the total time, the various ways of 
treating this category had relatively little effect on our 
estimations of the percentages of time the different age/sex 
classes of otters spent resting and feeding (Table 4.3). We 
therefore included the "unknown" data in the "other" category 
in subsequent analyses. 


74 


We tested differences in average percent of time spent 
feeding among sex and age classes using analysis of variance 
for the percent of time feeding for each individual monitored 


for 24-hour activity data. The results of this analysis, 
using Scheffe's multiple comparison test for difference among 
sex/age classes, showed that juvenile females fed 


significantly more than adult females, adult males and 
juvenile males, but not more than adult females with pups (p 
< .05) (Table 4.4). Females with small pups fed more than 
females with large pups but this difference was not 
Significant (Table 4.5). 


To facilitate a comparison of our data to data collected 
visually (Estes, et al., 1986), we tabulated the number of 
ten-minute periods devoted to "resting", "feeding" and "other" 
for each observation period in three ways: 1) for all data 
recorded during that observation period (Appendix 4.1); 2) for 
those data recorded during daylight hours (defined as 1/2 hour 
before sunrise to 1/2 hour after sunset) (Appendix 4.2); and 
3) for those data recorded while the otter could be seen 
(Appendix 4.3). 


The various age/sex classes of otters spent about the 
same percentages of time resting and feeding during daylight 
hours (Table 4.6) as they did over the entire 24-hour period. 
When only the data collected visually were considered, 
juvenile females still fed more than adult males and adult 
females (Table 4.6); juvenile females and adult males fed a 
greater, and adult females a smaller percentage of the time 
than indicated in the other two data sets. Insufficient 
visual data were collected on juvenile males and adult females 
with pups to allow a comparison of these groups. Comparisons 
of coefficients of variation (for the means of individuals 
for each sex/age category for the given method of data 
collection) indicated that the data collected visually were 
the most variable and those collected over the complete 24-- 
hour period the least variable (Table 4.6). 


DISCUSSION 


Activity in relation to time of day 


Sea otter activity patterns can be affected by a variety 
of factors including geographical location, weather, season, 
available prey, and age/sex class (Garshelis, 1983, Estes et 
al., 1986). Otters in California tend to be crepuscular, 
resting mainly in the middle of the day (Ribic, 1982; 
Loughlin, 1980; our data); although Estes, et al., found that 


75 


TABLE 4.3 - A comparison among the methods of calculating time 
budgets. The unknown data are classified by four different 
methods: including them in the "other" category, separating 
them as unknown, excluding them, and including them in 
"feeding". OP = observation periods, AM = adult male. Other 
abbreviations as in Table 4.1. 


TREATMENT OF AGE/SEX PERCENTAGE OF TIME SAMPLE SIZES 
UNKNOWN DATA CLASS Rest Feed Other Unknown Otters Hours OP 
AS OTHER AF 48 37 15 i 8 830 28 
AS UNKNOWN 48 36 8 7 

EXCLUDED 52 37 9 —— 

AS FEEDING 48 43 8 -- 

AS OTHER AFP 45 39 16 -- 6 264 8 
AS UNKNOWN 44 39 11 4 

EXCLUDED 46 41 12 —=— 

AS FEEDING 44 43 11 i 

AS OTHER AM 50 36 14 i 4 216 7 
AS UNKNOWN 50 37 9 4 

EXCLUDED 52 38 9 i 

AS FEEDING 50 41 9 —= 

AS OTHER JF 40 48 12 -- 9 417 12 
AS UNKNOWN 40 49 10 1 

EXCLUDED 40 49 10 -- 

AS FEEDING 40 50 10 -- 

AS OTHER JM 34 37 29 == 5 218 8 
AS UNKNOWN 34 37 24 4 

EXCLUDED 35 39 26 == 

AS FEEDING 34 41 24 Oe 


TABLE 4.4 - Analysis of variance testing for differences in 
percent of time spent feeding between the various sex/age 
classes. 


Source d.f. Sos. M.S. F 
Among sex/age 
classes 4 956 426.8 10.7 

Error 24 1707 39.8 
Test of means (% of time feeding) 

JF AFP — AM JM AF 

48 39 

36 37 37 


Means underlined by the same line are not significantly 
different (p<.05; Scheffe's test). 


TABLE 4.5 - A comparison of time budgets between females that 
were accompanied by small pups and those accompanied by large 


pups. 


ACTIVITY MEAN VARIANCE NUMBER OF 
PERCENTAGE OBSERVATION 
OF TIME PERIODS 
FEMALES WITH SMALL PUPS 
Rest 40.60 257.30 4 
Feed 42.71 194.80 4 
Other 16.53 12.29 4 
FEMALES WITH LARGE PUPS 
Rest 48.45 17.97 3 
Feed 36.02 104.70 3 
Other 15.53 39.38 3 


Table 4.6 - A comparison of the percent of time spent in each 
of the three activity categories, resting, feeding and other, 
for activity data collected over the entire 24-hour period, 
for daylight data only, and for visual data, for the age/sex 
categories of adult females, adult females with pups, adult 
males, juvenile females, and juvenile males. 


ENTIRE 24-HR PERIOD DAYLIGHT ONLY VISUAL DATA 
MEAN NO.OBS.* MEAN _NO.OBS.* MEAN _NO.OBS. * 
ADULT FEMALES 
Rest 48.2 28 47.72 28 54.4 24 
Feed 36.8 28 38.10 28 28.0 24 
Other 15.0 28 14.18 28 17.7 24 
ADULT FEMALES WITH PUPS 
Rest 44.52 8 46.99 8 
Feed 39.39 8 39.29 8 
Other 16.09 8 13.71 8 
ADULT MALES 
Rest 50.43 7 47.54 7 43.6 5 
Feed 35.8 7 38.06 7 41.2 5 
Other 13.77 7 14.40 7 14.6 5 
JUVENILE FEMALES 
Rest 39.76 12 34.83 12 38.5 11 
Feed 47.81 12 51.61 12 57.8 11 
Other 12.43 12 13.56 12 3.7 11 
JUVENILE MALES 
Rest 33.96 8 28.44 8 
Feed 36.84 8 37.15 8 
Other 29.20 8 34.42 8 


*NUMBER OF OBSERVATION PERIODS 


77 


the otters in one of the areas they sampled had no apparent 
24-hour pattern. Garshelis (1983) found deviations from the 
crepuscular pattern when locally preferred prey tended to be 
active at night. Also, he suggested that, in Alaska, short 
day lengths and poor weather conditions, combined with poor 
resources, may have made it impossible for otters to maintain 
body temperature during long rest periods. It might seem less 
likely that this would occur in the milder California climate; 
however, one of our juvenile females was observed shivering 
on several occasions before she disappeared and presumably 
died. Estes, et al., (1986) suggested that environmental 
factors such as wind and waves may disrupt the 24-hour pattern 
in California. Although we found that otters of all age/sex 
classes in California did some feeding at night, none fed 
primarily at night as did male otters at one location in 
Alaska where the preferred prey was dungeness crabs, which are 
thought to be more active, and hence more vulnerable, at night 
(Garshelis, et al., 1986). Because we did not collect 
activity data in extremely bad weather, we were unable to 
determine whether or not otters in California rest for shorter 
periods at midday under such conditions. 


In contrast to the other age/sex classes, juvenile 
females tended to rest for the greatest proportion of the time 
from midnight to eight in the morning rather than feed. Thus, 
for our data the juvenile females departed most from the more 
usual crepuscular pattern. Sea otters are known to steal food 
from other individuals (Fisher, 1939; pers. obs), and we 
observed that juvenile females were the group from which food 
was often taken. Perhaps they tend to feed at different times 
than the majority of their conspecifics to reduce the risk of 
losing prey in this manner. 


Time budgets in relation to other studies 


As Ribic (1982) did not separate feeding from other kinds 
of activity, Loughlin's (1980) telemetry data on six otters 
in Monterey Bay are the only California data comparable to 
ours. Our data on adult males and females are very similar 
to Loughlin's data in spite of his small sample size, the 
different study areas and the number of years between the two 
studies (Table 4.7). Garshelis, et al., (1986) present time 
budget data based on telemetry for sea otters from two 
localities in Prince William Sound, Alaska: Green Island and 
Nelson Bay. Sea otters had been present in the Green Island 
area since the 1950's or earlier. Females remained in this 
area throughout the year; males visited for various periods 
of time during the breeding season. In contrast, otters had 
moved into Nelson Bay fairly recently. Most of the otters 
there were males, some of which moved seasonally to breeding 
areas such as Green Island. Both adult males and females fed 
Significantly less time at Nelson Bay, where large, high 


78 


quality prey such as clams and crabs were easily available, 
than at Green Island, where such items were rare. 


TABLE 4.7 - A comparison of the activity budgets for sea 
otters calculated in this study and those in the literature. 


AGE/SEX LOCATION METHOD PERCENTAGE OF TIME REFERENCE 


CLASS Rest Feed Other 

AM CALIF. TELEMETRY 57 33 10 Loughlin 1979 
AF CALIF. TELEMETRY 50 36 14 Loughlin 1979 
AM CALIF. TELEMETRY 50 36 14 This study 
AF CALIF. TELEMETRY 48 37 15 This study 
AFP CALIF. TELEMETRY 45 39 16 This study 
JM CALIF. TELEMETRY 34 37 29 This study 
JF CALIF. TELEMETRY 40 48 13 This study 


UNKNOWN CALIF. VISUAL 53-63 21-28 9-22 Estes, 
et al., 1986 


M AK** TELEMETRY 50 47 3 Garshelis*** 
AF AK** TELEMETRY 50 47 3 Garshelis*** 
AFP AK** TELEMETRY 43 53 3 Garshelis*** 
PUPS* AK** TELEMETRY 45 51 4 Garshelis*** 
M AK*¥*** TELEMETRY 49 37 aS Garshelis*** 
AF AK**** TELEMETRY 51 37 - 12 Garshelis*** 


*Independent pups 

**GREEN ISLAND (Alaska) 

***Garshelis, et al., 1986 (reference) 
***k*NELSON BAY (Alaska) 


Our data were collected in the central portion of the sea 
otter range in California, in areas where sea otters have been 
present for many years. The general patterns in our data were 
the same as those at Green Island, Alaska, where otters have 
also been established for a long time. At Green Island, adult 
males and females fed for about the same percentage of time 
and females with pups fed slightly more than females without 
pups. At Green Island, recently weaned, independent pups fed 
more than adults. Juvenile males in California fed about the 
same amount of time as adults, but juvenile females fed more 
than adults, except for females with pups. 


Estes, et al., (1986) estimated the proportion of time 
spent foraging by scan sampling (Altmann, 1974) the otters 
visible along the California coastline from dawn to dusk at 
1/2 hour intervals and recording the activity of each otter 
observed. Estimates of the proportion of time spent feeding 
based on this technique are lower than those based on 


79 


radiotelemetry (Table 4.7). If otters in California spend a 
higher proportion of their feeding during the night than 
during the day, then scan samples during the day would tend 
to underestimate the proportion of time spent feeding. 
Garshelis, et al., (1986) found that, in some areas of Alaska, 
otters do feed mainly during the night. However, we found no 
difference in the proportion of time spent feeding during 
daylight hours and that over the entire 24-hour day. The most 
probable explanation for the difference between the scan 
sample and telemetry data is that feeding otters have a lower 
probability of being seen than resting otters (Estes and 
Jameson, 1988), and are thus more often missed during scan 
samples. 


Time budgets as indicators of population status 


Estes, et al., (1986) concluded that further growth, in 
recent years, of the California population was not because of 
food limitation, primarily because their estimates of the 
proportion of time the population spent foraging were similar 
to those of populations in Alaska known to be below 
equilibrium density. However, when all the available data are 
considered (Table 4.7) it seems likely that their scan 
sampling data underestimated the time spent feeding. 
Certainly, other factors than food availability, such as 
accidental capture in gill-nets, have contributed to the 
reduced growth of the California .sea otter population 
(Wendell, et al., 1986). Further, time budgets may be 
affected by factors other than the prey availability, such as 
weather conditions, prey type, and study methods that obtain 
data in different ways. It seems likely that the best 
comparisons of activity data are those between studies based 
on telemetry methods. We did not collect data during bad 
weather. Garshelis, et al., (1986) were able to use automatic 
recording to obtain data during bad weather in Alaska, and 
found that otters fed more during such periods. Thus, we 
probably underestimated the proportion of time California 
otters spend feeding. Nevertheless, we found juvenile females 
fed at least 48 percent of the time; which represents a 
substantial effort when compared to any existing activity 
data. 


They also tended to have longer feeding bouts than otters 
in the other age/sex classes, although the intervals between 
these bouts were about the same as in the other classes 
(Chapter 5). Two of our juvenile female otters (numbers 44 
and 46) had their prey stolen repeatedly. The prey stealing 
was selective: only large desirable items were stolen. Otters 
whose prey is stolen may temporarily stop foraging, move to 
another location, or capture apparently less desirable species 
of prey. Otter 44 subsequently died: her stomach was empty 
although there were shells in her intestine. The only otter 


80 


in our study that regularly hauled out was another juvenile 
female (otter 42). Hauling-out is a behavior that may help 
to conserve energy (Garshelis, 1983). 


Differences in the ability of members of different 
age/sex classes to compete for food resources are common in 
vertebrates (Sutherland and Parker, 1985; Clutton-Brock and 
Albon, 1985; Dunbar, 1985). Because of these differences in 
competitive ability, the effects of food shortage are usually 
concentrated on particular individuals. If these individuals 
cannot obtain sufficient food, they must emigrate or starve. 


After constructing models of vertebrate populations 
composed of individuals with varying degrees of competitive 
ability, Sutherland and Parker (1985) argued that "the average 
individual in the population can be doing very well despite 
the population being at carrying capacity". They concluded 
that the proportion of time spent feeding at the population 
level is not a good measure of whether the population is 
limited by food supply and that it will probably be necessary 
to concentrate on the factors affecting the poorest feeders 
in the population to understand the carrying capacity and 
population dynamics of many vertebrates. In sea otters in 
California, the poorest feeders appear to be the juvenile 
females; also, many pups die before weaning (Chapter 2). In 
Alaska, some females apparently abandoned pups prior to the 
age of weaning (Garshelis and Garshelis, 1987). These authors 
hypothesized that the abandonment was related to the poor 
health of the female due to nutritional stress. 


The exact mechanisms that are operating to slow the 
growth of the California sea otter population remain unclear. 
However, we believe that additional research on mortality 
factors in pups and independent juveniles, particularly 
juvenile females, would be likely to provide a better 
understanding of the dynamics of the California sea otter 
population. 


LITERATURE CITED 


Altmann, J. 1974. Observational study of behavior: sampling 
methods. Behaviour 49: 227-267 


Ames, J. A. and G. V. Morejohn. 1980. Evidence of white 
shark, Carcharodon carcharias, attacks on sea otters, 
Enhydra lutris. Calif. Fish and Game 66: 96-209. 


Clutton-Brock, T. H. and S. D. Albon. 1985. Competition and 
population regulation in social mammals. Pages 557-575 


in R. M. Sibly and R. H. Smith, eds. Behavioural 
Ecology. Ecological Consequences of Adaptive Behaviour. 


Blackwell Scientific Publications, Oxford, U.K. 


81 


Dunbar, R. I. M. 1985. Population consequences of social 
structure. Pages 507-519 in R. M. Sibly and R. H. Smith, 
eds. Behavioural Ecology. Ecological Consequences of 
Adaptive Behaviour. Blackwell Scientific Publications, 
Oxford, U.K. 


Eberhardt, L. L. 1977. "Optimal" management policies for 
marine mammals. Wildl. Soc. Bull. 5: 162-169. 


Estes, J. A., K. E. Underwood and M. J. Karman. 1986. 
Activity-time budgets of sea otters in California. J. 
Wildl. Manage. 50:626-636. 


Estes, J. A. and R. J. Jameson. 1983. Summary of available 
population information on California sea otters. POCS 
Tech. Pap. 83-11 for Interagency Agreement 114-12-001. 
U. S. Fish and Wildlife Service and U. S. Minerals 
Management Service. 29 pp. 


Estes, J. A. and R. J. Jameson. 1988. A double-survey 
estimate for sighting probability of sea otters in 
California. J. Wildl. Manage. 52:70-76. 


Estes, J. A., R. J. Jameson, and A. M. Johnson. 1981. Food 
selection and some foraging tactics of sea otters. Pages 
606-641 in J. A. Chapman and D. Pursley, eds., Worldwide 


Furbearer Conference Proceedings. August 3-11, 1980. 
Frostburg, Maryland. 


Estes, J. A., R. J. Jameson, and E. B. Rhode. 1982. Activity 
and prey selection in the sea otter: influence of 
population status on community structure. Am. Nat. 242- 
258. 


Estes, J. A., K. E. Underwood, and M. J. Karmann. 1986. 
Activity-time budgets of sea otters in California. J. 
Wildl. Manage. 50:626-636. 


Fisher, E. M. 1939. Habits of the southern sea otter. J. 
Mammal. 20:21-36. 


Garshelis, D. L. 1983. Ecology of sea otters in Prince William 
Sound, Alaska. Ph. D. Thesis, University of Minnesota, 
Minneapolis, Minnesota. 321 pp. 


Garshelis, D. L. and J. A. Garshelis. 1987. Atypical pup 


rearing strategies by sea otters. Marine Mammal Science 
33:263-270. 


82 


Garshelis, D. L., J. A. Garshelis, and A. T. Kimker. 1986. 
Sea otter time budgets and prey relationships in Alaska. 
J. Wildl. Manage. 50: 637-647. 


Jameson, R. J. and J. A. Estes. 1988. Status of the California 
sea otter population. Abstracts, American Society of 
Mammalogists Annual Meeting, Clemson, South Carolina, 
June, 1988. 


Loughlin, T. R. 1980. Radio telemetric determination of the 
24-hour feeding activities of sea otters, Enhydra lutris. 
Pages 717-724 inc. J. Amlaner, Jr., and D. W. MacDonald, 


eds. A Handbook on Biotelemetry and Radio Tracking. 
Pergamon Press, Oxford, U. K. 


Miller, D. J. 1980. The sea otter in California. Calif. 
Coop. Oceanic Fish. Invest. Rep. 21: 79-81. 


Ralls, K., D. B. Siniff, C. Monnett, T. Eagle, and L. Ferm. 
1985. Summary of information pertaining to California 
permit to capture sea otters for scientific research. 
Report to California Fish and Game Commission, 104 pp. 


Ralls, K, J. Ballou, and R. L. Brownell, Jr. 1983. Genetic 


diversity in California sea otters: theoretical 
considerations and management implications. Biol. Consv. 
25:3:209-232. 

Ribic Ce JAG 1982. Autumn activity of sea otters in 


California. J. Mammal. 56:701-703. 


Siniff, D. B. and K. Ralls. 1986. Summary of information 
obtained on sea otter for MMS study on population status 
of California sea otters. Report to California Fish and 
Game Commission and Office of Sea Otter Coordination 
(USFWS). 86 pp. 


Sutherland, W. J. and G. A. Parker. 1985. Distribution of 
unequal competitors. Pages 255-273 in R. M. Sibly and 
R. H. Smith, eds. Behavioural Ecology. Ecological 


Consequences of Adaptive Behaviour. Blackwell Scientific 
Publications, Oxford, U.K. 


Wendell, F. E., R. A. Hardy and J. Ames. 1986. An assessment 
of the accidental take of sea otters, Enhydra lutris, in 
gill and trammel nets. Calif. Dept. Fish Game, Mar. Res. 
Tech. Rep. No. 54, 31 pp. 


83 


CHAPTER 5 


FEEDING PATTERNS OF CALIFORNIA SEA OTTERS 


K. RALLS, B. HATFIELD AND D. B. SINIFF 


NOVEMBER 30, 1988 


84 


INTRODUCTION 


Because sea otters can often be easily observed from 
shore, their diet and foraging patterns have been studied in 
various parts of their range (Estes, Jameson, and Johnson, 
1981; Ostfeld, 1982; Garshelis, 1983; Lyons, 1987). Most 
existing data have been obtained by visual observations of 
foraging otters. Relatively little is known about night-time 
foraging patterns, although studies using radio-telemetry have 
shown that sea otters in California do forage by night as well 
as by day (Loughlin, 1977; Ribic, 1982). 


Early studies primarily yielded information about diet 
and foraging patterns at the population level. They showed 
that the California population feeds almost entirely on 
macroinvertebrates, although some populations in Alaska and 
the U.S.S.R. also feed on epibenthic fish, and that there 
appeared to be a great deal of variation in the diet and 
foraging patterns of individual otters (Estes, et al., 1981). 
Recently, studies on otters that could be individually 
identified by flipper-tags have confirmed that individuals 
tend to specialize on one to three or more of the many 
available types of prey and shown that these patterns of 
specialization may be maintained for three or more years 
(Lyons 1987). 


Although we collected some data on uninstrumented otters, 
we focused on the foraging patterns of individual instrumented 
otters as indicated by radio-telemetry. Foraging sea otters 
alternatively dive to search for prey and return to the 
surface to breathe and consume their catch. Because radio 
Signals are not transmitted through sea water, we could 
measure the length of dives and surface intervals whether or 
not we could see the instrumented otters. We present data on 
dive and surface intervals, the length of feeding bouts and 
the intervals between them, diurnal and nocturnal foraging 
patterns, and variation in foraging patterns within and 
between age/sex classes. We also compare data collected by 
listening to the radio signal with those collected visually 
and discuss the way in which variation in foraging patterns 
appears to be related to competition among individuals. 


METHODS 


The majority of data were collected during watches 
intended primarily to obtain information on time budgets and 
activity patterns (Chapter 4). Most of these watches were 24 
to 48 hours in length. Shorter watches, designed specifically 
for the collection of feeding data, were conducted during 
morning and evening feeding periods on the otters in Morro Bay 
(otters 1-4). Data on the feeding activities of unidentified, 


85 


uninstrumented otters were also collected during one two-week 
period to obtain additional information on dive and surface 
interval lengths in relation to the size and type of captured 
prey. 


The length of dives and surface intervals was measured 
from the radio signal from the instrumented animals and by 
visual observation of the uninstrumented otters. As visual 
observations indicated that the occasional dives shorter than 
10 seconds were rarely feeding dives and that surface 
intervals shorter than five seconds were almost always the 
result of interruption by another otter, these were excluded 
from analyses. When possible, foraging individuals were 
observed from the shore through a high resolution telescope 
(Questar, 50x or 80x magnification), and the number, size, and 
species of captured prey were recorded. 


Lengths of feeding bouts and the intervals between bouts 
were measured to the nearest minute. If more than 30 minutes 
elapsed between two feeding dives, these two dives were taken 
as the end of one feeding bout and the beginning of a second 
feeding bout, respectively. Records with a large amount of 
activity recorded as "unknown" and those where there was any 
ambiguity as to the end of a feeding bout were not used for 
the determination of bout and interval length. Only complete 
feeding bouts were used for the calculation of feeding bout 
length but intervals between bouts were measured as long as 
the ending of the first bout and the beginning of the next 
were known. We defined the day as the period from 1/2 hour 
before sunrise to 1/2 hour after sunset. 


Otters were assigned to sex and age classes based on 
their weight at capture; estimated age, often based on the 
examination of cementum layers in a vestigial premolar 
extracted for this purpose; and, in the case of females, 
reproductive performance. All juveniles were judged to be no 
more than two years of age (Chapter 2). 


Statistical comparisons among age and sex classes were 
performed using analysis of variance, controlling for 
variation among individuals within classes. We performed a log 
(base 2) transform on the data to reduce heterogeneity of 
variances. All statements that differences are statistically 
significant are based on the 0.05 probability level. 


RESULTS 
Observations from the shore 

Because our instrumented otters often foraged at times 
when or in areas where they could not be easily observed, many 


of our visual observations were made on uninstrumented otters. 


86 


The data presented here consist of all visual observations on 
both instrumented and uninstrumented otters. The mean length 
of observed dives was 52.14 seconds (n = 712). Dive length 
varied with prey type to some extent, being least for mussels 
and greatest for octopus (Table 5.1) but was not related to 
prey size (Table 5.2). Surface times were clearly related to 
both prey type (Table 5.3) and prey size (Table 5.4). They 
were longest for large prey such as crabs, abalone, and 
octopus that often took an otter several minutes to eat. 


Success rate varied with prey type. Otters foraging on 
mussels and small, hard-bodied prey that they pounded with a 
rock had the highest success rates while those foraging on 
large, calorically rich prey such as clams, abalone, and crabs 
of the genus Cancer had the lowest success rates. These 
relationships were evident even within age/sex classes for a 
sample of the instrumented animals (Table 5.5). 


Data collected from the radio signal 


The data presented here consist of dive lengths, surface 
intervals, feeding bout lengths, and the lengths of intervals 
between feeding bouts, all of which were collected from the 
instrumented otters regardless of whether or not the animals 
were seen. 


Dive length -- The unweighted mean dive length for all 
instrumented otters was 73.56 seconds (n = 8254). Mean dive 
lengths for the individual instrumented otters ranged from 41 
to 149 seconds (Table 5.6). 


Analysis of variance on the log-transformed data 
indicated that there were significant differences among the 
lengths of the dives made by individuals within age/sex 
classes (Appendix 5.1). There were significant differences in 
dive length between age/sex classes (Table 5.7). Scheffe's 
test showed that adult males made shorter dives than adult 
females, who in turn made shorter dives than juvenile females 
and adult females with pups. Juvenile males made the longest 
dives, with a mean length of 116 seconds. The short dive 
times for adult males reflect the large proportion of dives 
from adult males in the relatively shallow waters of Morro Bay 
in our sample for this age/sex class. 


The distributions of dive lengths for the different 
age/sex classes showed that adult males and adult females with 
pups had more individual variation in dive lengths than the 
members of the other age/sex groups (Fig. 5.1). The 
distribution of dive lengths for juvenile males indicates 
considerable internal consistency, even though the average 
dive length for this group was by far the longest (Fig. 5.1). 


87 


TABLE 5.1 - A comparison among the average dive lengths (sec) 
for sea otters prior to the capture of various types of prey 
in California. 


PREY DIVE LENGTH 

TYPE MEAN (SEC) N VARIANCE 
MUSSEL 33.81 296 DATO 
CRABS (All) 56.07 261 831.76 
CRABS (Pugettia) 56.27 60 1186.86 
CLAMS 58.09 220 517.94 
CRABS (Cancer) 59.10 79 790.88 
SEA STAR 64.65 23 1361.44 
ABALONE 71.69 39 557.34 
POUNDED WITH ROCK 78.66 196 1576.49 
TUNICATE 79.18 28 230.29 
OCTOPUS 101.71 28 2493.63 


TABLE 5.2 - The average length of the feeding dives (sec) made 
by sea otters in California prior to the capture of prey of 
different sizes. 


PREY DIVE LENGTH 


SIZE MEAN N VARIANCE 
(SEC) 

NONE 63.34 1149 741.06 

SMALL 57.91 636 1245.61 

MEDIUM 60.39 214 716.18 

LARGE 64.90 229 882.29 

EXTRA LARGE 61.37 147 674.06 

Surface interval -- The unweighted mean surface interval for 
all otters was 64.50 seconds (n = 7944). Mean surface 
intervals for individual otters ranged from 25.5 to 155.3 
seconds (Table 5.8). Analysis of variance on the log- 


transformed data indicated that there were significant 
differences among the lengths of surface intervals for the 
individuals in all age/sex classes except the juvenile males 
(Appendix 5.2). There were also significant differences among 


88 


TABLE 5.3 - A comparison among the lengths of the surface 
interval (sec) that were required to consume the various prey 
items taken by sea otters in California. 


LENGTH OF SURFACE INTERVAL 


PREY MEAN 

TYPE (SEC) N VARIANCE 
TUNICATE 33.25 28 98.76 
SEA STAR 54.00 23 675.65 
MUSSEL 58.18 291 1290.32 
CRABS (Pugettia) 94.22 55 6498.75 
CLAMS 95.14 219 4398.47 
POUNDED WITH ROCK 97.60 195 3860.78 
CRABS (ALL) 120.67 267 19198.31 
OCTOPUS 132.89 28 21610.38 
ABALONE 150.92 39 25252.22 
CRABS (Cancer) 213.39 85 34860.80 


TABLE 5.4 - A comparison among the average lengths of the 
surface intervals (sec) that were required to consume various 
sizes of prey taken by sea otters in California. 


PREY LENGTH OF SURFACE INTERVAL 
SIZE MEAN N VARIANCE 
(SEC) 
NONE 30.36 1095 1050.30 
SMALL 55.56 635 2839.58 
MEDIUM 62.34 216 2436.76 
LARGE 122.59 226 14734.20 
EXTRA LARGE 177.66 148 22886.74 


the age/sex classes (Table 5.9). Adult females with pups 
tended to have the longest surface times, followed by juvenile 
males, and then a group consisting of the juvenile females, 
adult males, and adult females (Table 5.9). These trends were 
also apparent in the distributions of surface intervals for 
the age/sex classes (Fig. 5.2). 


89 


TABLE 5.5 - A comparison among the instrumented individuals 
in the various sex/age categories for the percentage of 
successful dives during feeding bouts. AF = adult female 
without pup; AFP = adult female with pup; AM = adult male; JF 
= juvenile female. No data were obtained for the juvenile 
males. 


OTTER AGE/SEX NUMBER NUMBER SUCCESS PRINCIPAL 
NUMBER CLASS SUCCESSFUL UNSUCCESSFUL RATE PREY 
DIVES DIVES PERCENT TYPE 
15 AF 102 115 47 ABALONE/OTHER® 
6 AF 173 59 75 CRABS 
22 AF 41 2 95 SHB** 
16 AFP 34 32 52 CRABS 
25 AFP 109 5 96 MUSSELS 
1 AM 244 616 28 CLAMS 
3 AM 33 63 34 CLAMS 
7 AM 47 12 80 CRABS 
Vi AM 82 15 85 CRABS | 
/SHB** 
4 AM 162 18 90 MUSSELS 
44 JF 31 40 47 CRABS#* 
37 JF 25 24 51 CRABS 
40 JF 44 27 62 CRABS 
45 JF 31 15 67 CRABS 
46 JF 142 8 95° SHB** @ 
* -- Otters with two prey types listed caught approximately 
equal numbers of each type. 
SHB** -- unidentified small, hard-bodied prey that were 
pounded with a rock. 
# -- Four of the prey items captured were stolen by other 
otters. 
@ -- Three of the prey items captured were stolen by other 
otters. 
-- Pugettia 
@ -= Cancer 


90 


TABLE 5.6 - A comparison of the average dive lengths (sec) 
recorded for individual instrumented sea otters during feeding 
bouts. AF = adult female without pup; AFP = adult female with 


pup; JF = juvenile female; AM = adult male; JM = juvenile 
male. 
OTTER AGE/ DIVE LENGTH 
NUMBER SEX MEAN N VARIANCE 
(SEC) 
46 JF 41.48 427 485.47 
4 AM 41.50 161 255.80 
9 AF 45.66 772 323.87 
15 AF 52.28 774 653.49 
1 AM 57.05 733 647.24 
37 JF 57.77 70 258.72 
3 AM 58.03 88 650.03 
34 AM 61.46 69 3694.60 
16 AFP 61.73 194 487.53 
16 AF 62.69 179 1224.25 
25 AFP 66.91 487 1297.79 
45 JF 71.52 406 689.53 
39 JF 71.85 239 369.36 
19 AF 73.94 48 330.10 
2 AM 74.91 22 626.45 
6 AF 76.89 1006 818.20 
7 AM 77.48 101 1298.72 
17 AM 80.13 210 3411.17 
36 AFP 83.20 119 976.18 
44 JF 83.52 326 540.65 
42 JF 92.33 271 1062.93 
40 JF 92.86 308 1129.69 
13 JM 95.91 107 1861.67 
43 JM 100.50 22 2245.70 
27 AFP 100.56 179 1077.39 
47 JF 104.93 160 2580.98 
30 JM 115.64 119 1183.44 
41 JM 132.43 127 231.01 
29 JF 132.46 118 1409.91 
35 JM 135.82 123 1717.04 
22 AF 140.18 153 1019.31 
14 AFP 149.40 136 1106.62 
Feeding bouts -- Mean feeding bout lengths for individual 


otters ranged from 77 to 373 minutes (Table 5.10). The 
shortest mean bout length was for an adult male feeding 
primarily on clams in Morro Bay and the longest was for a 
juvenile female feeding on small, hard-bodied prey items in 
the Piedras Blancas area. Tallying the number of individuals 
within the age/sex classes according to the mean length of 
feeding bout suggested that juvenile females, and to a lesser 


91 


FIGURE 5.1 -- The distribution of dive times during feeding 
bouts for adult males, adult females, adult females with 
pups, juvenile females and juvenile males in California. 


40 


Adult Males 


30 


20 


Adult Females 


Adult Females With Pups 


Juvenile Females 


Percent of Dive Times 


Juvenile Males 


60 150 240 >315 


Seconds 
92 


FIGURE 5.2 -- The distribution of the length of time of the 
surface intervals during feeding bouts for adult males, adult 
females, adult females with pups, juvenile females and 
juvenile males in California. 


40 


Adult Males 


30 


20 


Adult Females 


Adult Females With Pups 


Juvenile Females 


Percentage of Surface Times 


20 


10 


Juvenile Males 


20 


10 


1 
60 150 240 >315 


Seconds 
93 


TABLE 5.7 -- Analysis of variance testing for differences in 
the length of the dives made by otters belonging to the 
various age/sex classes. Abbreviations as in Table 5.6. 


Source df Mean square F p 
Total 7937 
Among age/sex classes 4 128.37 242.2 <0.001 
Error 7933 6.53 
Test of means: 

AM AF Bye Nn JM 
Mean dive length (sec) 64 75 83 92 116 


The means for the age/sex classes underlined by the same line 
are not significantly different from each other (P<0.05, 
Scheffe's test). 


extent, adult females, tended to have long feeding bouts 
(Table 5.11A). The two individuals with mean feeding bout 
lengths over 250 minutes were both juvenile females (otters 
45 and 46). When we tallied the number of bouts of various 
lengths for each age/sex class (Table 5.11B), we noted that 
juvenile females tended to be the most different from the 
others. Chi-square analysis testing for shifts in the 
distribution of lengths of feeding bouts among sex/age classes 
(Table 5.11B) showed a highly significant difference (p<0.01). 
When we eliminated the juvenile females from the analysis, the 
Chi-square was no longer significant, even at the 0.05 
level,indicating no difference in bout length among the 
remaining age/sex classes. 


Interval between feeding bouts -= The mean interval between 
feeding bouts was 187.7 minutes (n = 228). This is close to 
the approximately 180 minutes required for food to pass 
through the digestive system (Stulken and Kirkpatrick, 1955; 
Costa, 1982). Values for individuals ranged from 80.9 to 300.9 
minutes (Table 5.12). Juvenile females and females with pups 
appeared to have shorter intervals between feeding bouts than 
the other age/sex classes. However, Chi-square analysis of the 
data in Table 5.13B indicated that there was not a significant 
difference among age/sex classes in the pattern of the 
distribution of lengths of the time between feeding bouts. 


Comparison of day and night foraging patterns 


We were able to compare day and night foraging patterns 
in four age/sex groups: adult females, adult females with 
pups, juveniles males and juvenile females. Unfortunately, we 
had insufficient data on the nocturnal foraging patterns of 
adult males to include them in this analysis. 


94 


TABLE 5.8 - The average surface times (sec) for individual 
instrumented sea otters that were recorded during feeding bouts. 
Abbreviations as in Table 5.6. 


ES 


OTTER AGE/ SURFACE TIME 
NUMBER SEX MEAN N VARIANCE 
(sec) 

37 JF 25.52 69 1137.81 
46 JF 38.72 424 1835.33 
19 AF 39.86 44 696.03 
9 AF 43.43 744 1972.60 
39 JF 44.19 242 1016.50 
3 AM 49.23 86 4547.46 
15 AF 49.43 732 4359.00 
1 AM 50.17 701 3140.20 
45 JF 50.56 396 7165.18 
6 AF 52.14 985 3300.86 
7 AM 52.52 99 3119.66 
2 AM 53.40 20 1745.54 
4 AM 59.99 158 2359.77 
25 AFP 65.98 453 2467.77 
44 JF 67.27 324 11250.17 
42 JF 69.38 276 6846.35 
40 JF 77.71 301 8128.68 
16 AFP 78.22 188 6867.25 
41 JM 84.26 125 293.02 
35 JM 84.47 113 5525.51 
17 AM 88.36 190 9636.31 
34 AM 91.17 63 3750.24 
47 JF 92.47 158 9386.40 
27 AFP 93.54 163 10118.73 
30 JM 94.43 115 6304.85 
16 AF 96.74 170 16644.03 
43 JM 102.29 21 7415.98 
13 JM 106.17 103 9990.12 
29 JF 106.93 109 6643.02 
14 AFP 131.90 119 1697.86 
36 AFP 139.50 105 10092.75 
22 AF 155.33 148 7428.96 


TABLE 5.9 - Analysis of variance testing for differences 
between the mean length of the surface intervals made by 
otters of the various age/sex classes. Abbreviations as in 
Table 5.6. 


Source af Mean square F Pp 
Total 7645 168.98 132.01 <0.001 
Among age/sex classes 4 
Error 7641 
Test of means: 

JF AM AF JM AFP 
Mean surface 
interval length (sec): 63 64 WS 94 102 


Dive length -- Analysis of variance of the log-transformed 
data indicated that there were significant differences between 
the length of day and night dives for the adult females with 
pups and the juvenile females but not for the adult females 
without pups and the juvenile males (Appendix 5.3). However, 
there were significant interactions between individuals and 
the length of day and night dives for all four age/sex groups. 


The mean lengths of the day and night dives of the 
individual otters are compared in Table 5.14. Some individuals 
made longer dives at night, some during the day, and others 
made dives of about the same length during both periods. Only 
two of the four adult females with pups tended to make longer 
dives at night, even though the analysis of variance indicated 
a significant difference between day and night dive length for 
this group. The juvenile females tended to make longer dives 
during the day but again there was no consistency within the 
group (as indicated by the significant interaction terms in 
the ANOVA), with only four of the seven individuals making 
significantly longer dives during the day. 


Surface intervals -- Analysis of variance of the log- 
transformed data indicated that there were significant 
differences between the length of the day and night surface 
intervals for the juvenile males and females but not for the 
adult females or the adult females with pups (Appendix 5.4). 

There were significant interactions between individuals and 
the length of day and night surface intervals for all age/sex 
classes except the adult females with pups. The mean lengths 
of the day and night surface intervals of the individual 
otters are compared in Table 5.15. The general pattern was 
similar to that for the dive length data in that some otters 


96 


TABLE 5.10 - The average lengths of the feeding bouts (min) 
that were recorded for individual instrumented sea otters. 
Abbreviations as in Table 5.6. 


OTTER AGE/SEX BOUT LENGTH 


NUMBER CLASS MEAN N 
(MIN) 

1 AM 77.33 15 
13 JM 88.00 5 
36 AFP 90.00 9 
“41 JM 97.50 24 
30 JM 99.50 8 
27 AFP 103.90 10 

9 AFP 106.60 5 
11 AF 109.50 6 
34 AM 111.57 7 
29 JF 118.67 6 
22 AF 119.23 13 
16 AF 126.27 15 
43 JF 130.40 10 
14 AFP 133.60 5 
35 JM 133.64 17 
17 AM 134.63 11 
16 AFP 138.29 7 

7 AM 139.71 7 

6 AF 142.66 18 
43 JM 145.67 3 
25 AFP 145.77 9 
39 JF 146.50 9 
19 AF 166.56 16 
15 AF 171.06 16 
42 JF 176.43 7 
37 JF 186.60 5 
40 JF 195.70 10 

9 AF 201.25 4 
36 AF 237.33 3 
47 JF 244.42 7 
45 JF 368.50 4 
46 JF 373.00 3 


97 


TABLE 5.11 - The lengths of feeding bouts (min), grouped by 
time intervals, classified by the five age/sex categories of 
adult males, adult females, adult females with pups, juvenile 
females, and juvenile males. Part A is a tabulation of the 
average feeding bout lengths of each instrumented otter. Part 
B is a tabulation of the lengths of the individual feeding 
bouts. 


A. MEAN FEEDING BOUT LENGTHS OF INDIVIDUAL OTTERS WITHIN 
CLASSES 


BOUT ADULT JUVENILE FEMALES ADULT JUVENILE | 
LENGTH MALES MALES WITH PUPS FEMALES FEMALES 


(MIN) ———————— EEE 


<150 4 5 6 3 3 
150-250 0) (0) 0 4 4 
>250 0 0) 0 0 2 


B. LENGTH OF INDIVIDUAL FEEDING BOUTS WITHIN CLASSES 


BOUT ADULT JUVENILE FEMALES ADULT JUVENILE 
LENGTH MALES MALES WITH PUPS FEMALES FEMALES 
(MIN) 

<150 19 41 30 40 29 
150-250 6 14 13 23 14 
>250 2 2 2 10 17 


had longer surface intervals at night, some during the day, 
and others had surface intervals of about the same length 
during both periods. However, fewer otters (nine) had 
significant differences between the length of day and night 
surface intervals than between day and night dive lengths (15) 
and values of the test statistic, K, tended to be smaller for 
surface intervals for dive lengths. 


Feeding bouts -- The mean length of entirely nocturnal feeding 
bouts (119.16 minutes, n = 43) was similar to that of those 
that occurred entirely during daylight hours (120.67 minutes, 
n = 62) but bouts that spanned the transition period from day 
to night or night to day tended to be considerably longer 
(209.80 minutes, n = 25). 


DISCUSSION 


Comparison of data with existing data sets 


The results of our visual observations agree with those 
of the principal previous study in California (Estes, et al., 
1981). Our overall mean dive time was 52 seconds; Estes, et 


98 


TABLE 5.12 - The average lengths of the intervals between 
feeding bouts (min) that were recorded for the individual 
instrumented sea otters. Abbreviations as in Table 5.6. 


OTTER AGE/SEX LENGTH OF INTERVAL BETWEEN BOUTS 
NUMBER CLASS MEAN N 
(MIN) 
27 AFP 80.89 9 
7 AM 96.67 3 
14 AFP 103.60 5 
37 JF 113.50 4 
29 JF 121.60 25 
41 JM 121.91 22 
44 JF 122.89 9 
25 AFP 139.77 9 
39 JF 143.57 7 
47 JF 146.57 7 
9 AFP 160.20 5 
36 AF 167.33 3 
42 JF 170.00 5 
35 JM 170.29 14 
40 JF 175.38 8 
45 JF 176.00 4 
17 AM 177.78 9 
16 AFP 179.17 6 
36 AFP 184.00 6 
6. AF 222.35 14 
16 AF 233.54 11 
13 JM 236.25 4 
11 AF 242.80 5 
9 AF 244.33 3 
22 AF 254.50 12 
30 JM 254.86 7 
34 AM 272.60 5 
19 AF 287.46 


PR 
Nu 


15 AF 300.91 


99 


TABLE 5.13 - The lengths of the intervals between feeding 
bouts (min), grouped by time intervals, classified by the five 
sex/age categories of adult males, adult females, adult 
females with pups, juvenile females, and juvenile males. Part 
A is a tabulation of the average interval between feeding 
bouts for each instrumented otter. Part B is a tabulation of 
the individual intervals between feeding bouts. 


A. AVERAGE INTERVAL BETWEEN BOUTS FOR INDIVIDUAL OTTERS 


INTERVAL ADULT ADULT ADULT JUVENILE JUVENILE 


LENGTH FEMALES FEMALES MALES MALES . FEMALES 
(MIN) WITH PUP 

<150 0) 3 1 1 6 
150-250 4 3 2 2 3 
250-350 3 io) 1 2 0 
350-450 0) 0) 0 ce) (0) 
450-550 (0) 0) 0 0) fe) 
>550 0) te) 0) 0) 0) 


B. INDIVIDUAL BOUTS 


INTERVAL ADULT ADULT ADULT JUVENILE JUVENILE 


LENGTH FEMALES FEMALES MALES MALES FEMALES 
(MIN) WITH PUP 

<150 19 25 7 30 35 
150-250 16 10 6 7 7 
250-350 8 3 9 4 1 
350-450 7 0) (0) 5 4 
450-550 8 0) 1 2 3 
>550 3 3 (0) 1 0) 


al. (1981) reported mean dive times of 50 to 60 seconds. Like 
Estes, et al., (1981), we also found no relationship between 
dive length and prey size and that dive length was not greatly 
affected by prey type, with the exception of dives for a few 
prey types such as mussels, which tended to be short, or 
octopus, which tended to be long. We found that surface 
times were longer for large, calorically rich prey items and 
that success rates were higher for small prey items such as 
mussels; again this agrees with previous studies in both 
California (Estes, et al., 1981) and Alaska (Garshelis, 1983). 


Although Loughlin's (1977) mean dive length of 57 seconds 
based on telemetry data was similar to that derived from 
visual observations (Estes, et al., 1981), the relatively few 
individuals he studied appear to have foraged close to shore 


100 


TABLE 5.14 - A comparison of the mean lengths of dives made 
during the day and night by the individual instrumented otters 
in the various age/sex classes. The Kruskal-Wallis test was 
used to test for differences between day and night means, at 
the 0.05 probability level. 


OTTER DAY NIGHT LONGER 
NUMBER MEAN N MEAN N DIVES 
Adult females 
15 62.8 433 47.8 285 day 
9 45.1 736 56.4 36 night 
6 71.3 538 83.3 468 night 
22 141.2 62 139.5 91 ns 
36 66.9 52 95.9 67 night 
Adult females with pups 
25 53.9 334 95.3 153 night 
16 63.1 286 59.1 87 ns 
27 94.9 75 104.6 104 night 
14 147.7 80 151.8 56 ns 
Juvenile males 
13 105.7 72 75.8 35 day 
30 114.7 68 116.9 51 ns 
43 40.6 7 128.7 15 night 
41 131.7 105 137.9 22 ns 
35 129.3 91 154.3 32 night 
Juvenile females 
45 72.9 272 68.7 134 day 
39 71.7 97 72.8 141 ns 
40 102.3 206 73.8 102 day 
46 49.9 217 38.9 210 day 
44 84.6 266 78.8 60 day 
42 90.2 124 94.2 147 ns 
47 94.0 116 133.9 44 night 


Adult males 
17 62.5 161 138.2 49 night 


101 


TABLE 5.15 - A comparison of the mean lengths of surface 
intervals made during the day and night by the individual 
instrumented otters in the various age/sex classes. The 
Kruskal-Wallis test was used to test for differences between 
day and night means, at the 0.05 probability level. 


OTTER DAY NIGHT LONGER 
NUMBER MEAN N MEAN N SURFACE 
INTERVAL 
Adult females 
15 48.6 406 Bra) B/S) night 
9 43.3 709 46.6 35 ns 
6 54.0 522 50.1 463 ns 
22 160.9 62 151.3 86 ns 
36 119.4 46 155.2 59 ns 
Adult females with pups 
25 66.6 312 64.7 141 ns 
16 86.2 273 89.6 85 ns 
27 78.6 67 104.0 96 night 
14 130.6 73 134.0 46 ns 
Juvenile males 
13 132.0 70 51.4 33 day 
30 81.4 66 111.9 49 night 
43 112.8 6 98.1 15 ns 
41 87.2 102 Yakoil 23 ns 
35 93.9 83 53.8 30 day 
Juvenile females 
45 49.2 266 Bs} oS} als}(o) ns 
39 49.4 96 40.7 146 day 
40 78.5 206 75.9 95 ns 
46 47.4 213 30.0 211 day 
44 69.6 271 55.4 53 ns 
42 77.8 127 62.2 149 day 
47 80.6 114 123.3 44 ns 
Adult males 
17 Gio aay 145.7 43 night 
in the Monterey area. Our more extensive telemetry data 
indicated that visual observations of otter foraging in 
California tend to underestimate mean dive lengths. As 


Garshelis (1983) found that dive length was correlated with 
water depth in Alaska, this is probably because feeding otters 
can only be observed easily when they are foraging close to 
shore in relatively shallow water. The mean dive length for 
our instrumented animals was about 13 seconds longer than our 
estimate based on visual observations of both instrumented and 


102 ° 


uninstrumented otters. Although Estes, et al., (1981) 
reported, based on visual observations, that "dives longer 
than 125 seconds almost never occurred" in California, we 
found that five of 31 otters had mean dive lengths that 
exceeded 125 seconds and that twelve otters had maximum dive 
lengths of over 200 seconds. The longest dive we timed was 
246 seconds; previous reports of maximum dive lengths are 200 
and 275 seconds for California (Estes, et al., 1981 and 
Loughlin, 1979, respectively), and 205 seconds in Alaska 
(Garshelis, 1983). 


Visual observations have given the impression that 
adults forage in deeper water than juveniles (Estes, et al., 
1981). Although our otter with the shortest mean dive length 
was a juvenile female, another juvenile female had a mean dive 
length of 132 seconds. Furthermore, juvenile males spent much 
of their time far offshore beyond the kelp beds and tended to 
forage farther offshore, and hence probably in deeper water, 
than otters of other age/sex classes (see Chapter 3). Because 
our juvenile males tended to forage so far from shore, we were 
unable to observe them feeding, but the radio signal indicated 
that they tended to make longer dives than otters in the other 
age/sex classes, with a mean length of 116 seconds. The 
behavior of such juveniles is clearly not reflected in 
previous data sets on sea otter diet and feeding patterns. 


Diurnal and nocturnal foraging patterns 


Many individual otters displayed differences in diurnal 
and nocturnal dive length patterns that may reflect a tendency 
to specialize on different prey species, that may tend to 
occur at different mean depths, by day and night. However, 
there was no general tendency for longer dive lengths during 
either time period. Some of the many prey items available to 
the California population may be more vulnerable at night. 
For example, crabs belonging to the genus Cancer and octopuses 
are generally thought to be nocturnal (Estes, et al., 1985; 
Ricketts, et al., 1986; Barr and Barr, 1983). Individual 
otters also vary in the extent to which they tend to feed at 
night. Since the mean length of diurnal and nocturnal feeding 
bouts was similar, differences in the distributions of diurnal 
and nocturnal surface intervals were less frequent than those 
for dive lengths, and the length of surface intervals is 
related to the size of prey consumed, most individuals may 
have similar diurnal and nocturnal rates of caloric intake. 


Our data on the length of surface intervals were 
generally similar to those in other studies, in that the 
length of surface intervals increased with the size of the 
captured prey. However, the time required to consume captured 
prey is not the only factor affecting the length of surface 
intervals. Adult females with pups and juvenile males had the 


103 


longest surface intervals. Visual observations showed that 
these were, in part, the result of social interactions of the 
adult females with pups and of the juvenile males with other 
juvenile males. 


Individual variation in foraging patterns 


The degree of individual difference in foraging patterns 
among California sea otters is striking. Data presented in 
this chapter indicate that individuals vary with respect to 
the size and species of prey consumed, dive length, surface 
interval length, feeding bout length, and the degree of 
difference between diurnal and nocturnal feeding patterns. 
Data presented in Chapter 4 indicated that individuals also 
vary in the total amount of time spent feeding per 24-hr day 
and the proportion of time they forage at night. 


Taken as a whole, these individual differences suggest 
that prey items are not equally available to all individuals 
in the California population. Juvenile females appear to be 
at a disadvantage compared to adults. They tended to feed for 
long periods and for a higher proportion of time than the 
other age/sex classes (Chapter 4). Much of their prey 
consisted of items that were too small to be identified and 
when they were successful in capturing a large, desirable prey 
item such as a crab belonging to the genus Cancer, this was 
often stolen by another otter. They often fed during the day 
when most otters were resting (Chapter 4), which probably 
helped them avoid competition with older animals. 


Juvenile males often fed farther off-shore than the 
other age/sex groups (Chapter 3), on unknown prey species. 
As a group, juvenile males had longer surface intervals than 
juvenile females and their feeding bout lengths were similar 
to those of adults. In general, our results on foraging, time 
budgets and activity, movements, and survival strongly suggest 
that juvenile females tend to be at a disadvantage in the 
portion of the range where we studied. 


LITERATURE CITED 


Barr, L. and N. Barr. 1983. Under Alaskan Seas: the 
Shallow-water Marine Invertebrates of Alaska. Northwest 
Publishing Company, Anchorage, AK. 208 pp. 


Costa, D. P. 1982. Energy, nitrogen, and electrolyte flux 
and sea-water drinking in the sea otter Enhydra lutris. 
Physiol. Zool. 55: 35-44. 


Estes, J. A., R. J. Jameson, and A. M. Johnson. 1981. Food 
selection and some foraging tactics of sea otters. Pp. 
606-641 in J. A. Chapman and D. Pursley (eds.). Worldwide 


104 


Furbearer Conference Proceedings. August 3-11, 1980. 
Frostberg, Maryland. 


Estes, J. A. and G. VanBlaricom. 1985. Sea-otters and 
shellfisheries. Pp. 187-235 in J. R. Beddington, R. J. 
Beverton and D. M. Lavine (eds.). Marine Mammals and 


Fisheries. George Allen and Unwin, London. 


Garshelis, D. L. 1983. Ecology of sea otters in Prince 
William Sound, Alaska. Ph. D. Thesis, University of 
Minneapolis, Minnesota. 321 pp. 


Loughlin, T. R. 1977. Activity patterns, habitat 
partitioning, and grooming behavior of the sea otter, 
Enhydra lutris, in California. Ph. D. Thesis, University 
of California, Los Angeles, 110 pp. 


Lyons, K. 1987. Abstract. Individual variation in diet and 
foraging strategy in the female California sea otter, 
Enhydra lutris. Animal Behavior Society, 21-26 June 1987, 
Williamstown, Mass. 


Ostfeld, R. S. 1982. Foraging strategies and prey switching 
in the California sea otter. Oecologia 53: 170-178. 


Ribic,» C..5 A. 1982. Autumn activity of sea otters in 
California. J. Mamm. 63:702-706. 


Ricketts, E. F. and J. Calvin. 1986. Between Pacific Tides. 
4th edition, revised by J. W. Hedgpeth. Stanford 
University Press, Stanford, CA. 614 pp. 


Stulken, D. E. and C. M. Kirkpatrick. 1955. Physiological 
investigation of captive mortality in the sea otter 
(Enhydra lutris). Trans. 20th N. Amer. Wildl. Conf.: 
476-494. 


105 


CHAPTER 6 


AGE DETERMINATION OF CALIFORNIA SEA OTTERS FROM TEETH 


P. PIETZ, K. RALLS, AND L. FERM 


NOVEMBER 30, 1988 


106 


INTRODUCTION 


Determining the age of individuals by counting incremental 
lines in tooth cementum has proved to be a useful tool for a 
wide variety of mammal species (for a review see Grue and 
Jensen 1979). It has been used to estimate ages of both 
salvaged and living sea otters in Alaska (Schneider 1973 and 
Garshelis 1984, respectively). 


Accurate age estimates of living animals offer potential 
insights into many aspects of sea otter biology, such as the 
age of first reproduction in females, the ages of territorial 
males, and age-related differences in movement patterns. 
Accurate age estimates of salvaged animals may also be useful 
as a means of estimating population age structure and, thus, 
for constructing models that can detect and predict changes 
in population parameters. 


METHODS 


In an effort to gain insight into the age structure of the 
California population, we have studied a sample of premolars 
from more than 580 sea otters. We extracted 30 premolars 
(PM,, as recommended by Schneider 1973) from animals that were 
captured and radio-tagged; we collected the rest from skulls 
of dead otters salvaged by the California Department of Fish 
and Game and the U.S. Fish and Wildlife Service. Skulls of 
many salvaged animals had been deposited at numerous 
institutions and agencies. The majority of our teeth were 
taken from skulls in collections at the following 
institutions: Santa Barbara Museum of Natural History, 
California Polytechnic State University (San Luis Obispo), 
California Department of Fish and Game (Monterey, Morro Bay), 
U.S. Fish and Wildlife Service (Piedras Blancas), University 
of Puget Sound, and San Jose State University. 


The teeth were decalcified, sectioned, and stained by Gary 
Matson, P.O. Box 308, Milltown, Montana. From teeth mounted 
in paraffin, he cut longitudinal sections 14 m thick and 
stained them with Wolbach's Giemsa. Basic procedures (Luna 
1968) were modified in conjunction with advice from Aleta Hahn 
of the National Marine Fisheries Service (S.W. Fisheries 
Center, P.O. Box 271, La Jolla, California). To determine 
ages, we counted bands in the cementum of the sectioned teeth 
using criteria outlined by Schneider (1973) for Alaskan 
otters. 


RESULTS AND DISCUSSION 
Despite the extensive analyses of sea otter teeth 
conducted by Schneider (1973) and Garshelis (1984), there are 


presently only a few teeth available from known-age otters. 


107 


This makes it difficult to evaluate the relationship between 
cementum lines and annual time intervals with certainty. 
However, we have been able to examine teeth from ten 
California otters of known minimum age, and the age estimates 
from these teeth compare quite favorably with the age 
estimates of the otters made by field biologists (Table 6.1). 
Comparisons of ages determined from tooth cementum with known 
ages for animals of a variety of other species are shown in 
Table 6.2. Although there is exact agreement in most cases, 
there is a difference of more than one year in a few 


S———SESE>E>E>S>S>SEeS=S=E=ES|"=a=Lh)™)™“|["Sl=S]=e@a|*[*i[is| _is{s(7(“os)sS>S> Si i i Sass ———————, 


WAIBINESs) Gra Tooth age estimates for sea otters of minimum known age. 
Univ. FUS* CDF&G Sex Date Est. age Date Est. age Tooth 
Minn. no. no. tagged when found at death age 
no. tagged” dead 
42 116 1142 F 10-82 {2 2 1-83 1° 2 1+ 
65 074 1012 F 10-80 0.5 - 1 9-81 Weds) 2 2 
68 115 1182 F 10-82 0.5 - 2 5-83 1-3 265 2 
617 097 1106 M 10-81 0.5 - 1 8-82 ino © 2 2 
41 = 1577 Mientec38)5 OS onl 4-87 A 2 B55 3 
47 090 1170 M 10°81 O.5 - 2 4-83 B95) Sob & 
616 : 1573 M 6-82 < 0.5 3-87 5 5 
55 051 1269 F 7-79 4+ 1-84 8.5+ 6 
440 : 715 F 12-72 i+ 8-79 7.5% 8 
615 = 14964 F 10-79 3+ 5-86 9.5+ 11 


A 


e All animals with FWS numbers were initially tagged by FWS. The 


others were tagged by CDF&G. 


2 Ranges include estimates by R. Jameson and J. Bodkin of FWS, to 
the nearest half-year. 


LLLL—h—w—haha Eh —L— LL 


cases. Judging by the data in Tables 6.1 and 6.2, the 
accuracy of the technique for sea otters may be similar to 
that for other species. 


Our optimism regarding the value of the cementum technique 
for age determination for sea otters arises from three 
sources: (1) the correlation between cementum lines and 
yearly intervals in known-age animals of numerous other 
species (Grue and Jensen 1973, Grue and King 1984); (2) the 
correlation between cementum lines and age estimates made by 
field biologists in our ten sea otter teeth of "known"-age 
(Table 6.1); and (3) the correlation between age class 
assignments determined from sea otter skulls and from teeth 
Cmnigis Bese 


108 


Age estimates based on skull features (e.g. suture 
closure, ridge development, tooth wear) were available for 
over 200 of the animals for which we have sectioned teeth. 
These estimates were provided by Jack Ames of the California 
Department of Fish and Game. He assigned skulls to five age 
categories: pups (< 6 months old); immatures (probably about 
1/2 to 11/2 years); subadults (probably 1-4 years); adults 
(probably at least 4 years), and old adults (probably at least 


10 years). Comparing tooth age estimates from our most 
experienced reader to these categories, we found that 96-100% 
of our readings for animals in the "pup", "immature", and 


"Subadult" categories fell within the appropriate age ranges 
(Fig. 6.1). As expected for the broader and less well defined 
categories of "adult" and "old adult", the range of tooth ages 
within each category was greater and the overlap between 
categories was greater. Nevertheless, 72% of teeth from 
animals in the "adult" category were assigned ages from 4-10 
years and 71% of those in the "old adult" category, ages from 
7-16 years. 


Ages of teeth from older animals are the most difficult 
to assess. Garshelis (1984) noted that it was difficult to 
determine exact ages when more than about 10 cementum annuli 
were present because annuli were spaced so closely together 
in older animals. Teeth obtained from skulls stored in 
museums may offer an additional problem; according to 
Schneider (1973), teeth allowed to air-dry developed dark 
edges which made it difficult to differentiate outer annuli. 


Despite these potential difficulties, our age assignments 
for teeth from "old adult" skulls were not unreasonably low. 
Male and female sea otters in Alaska have been estimated to 
live 10-15 and 15-20 years, respectively (Calkins and 
Schneider 1984). In California, the oldest tagged otters 
under observation are a 13-14 year-old female and two > 12 
year-old males (M. Riedman and J. Estes, pers. comm.). In our 
sample of 580 animals, the oldest age estimates based on tooth 
cementum were 16 years for two females and 15 years for two 
males. 


The technique of determining age by counting cementum 
lines can best be evaluated with an extensive reference 
collection of teeth from known-age sea otters. We are now in 
the process of developing a reference collection, but this 
requires long-term effort and inter-agency cooperation. In 
the meantime, we have attempted to evaluate other aspects of 
the technique: (1) variation within and among readers, (2) 
variation among premolars of the same individual, and (3) 
variation in readability due to different methods of preparing 
skulls and teeth. 


109 


Percent in Each Age Class 


FIGURE 6.1 -- Comparison of age estimates based on teeth to 
age-class assignments for the same sample of otters made by 
the California Department of Fish and Game using skull 
features. 


Subadults 


Adults 


234567891011 12 
Old Adults 


4° 8-456 7 BUononane ise liserc 


Estimated Ages (Years) 


110 


A total of 614 teeth were sectioned, stained, and "read" 
by at least one reader. A sub-sample of nearly 200 teeth was 
examined once by two readers and twice by one reader. About 
100 of these teeth were examined once by four readers. 
Statistical analysis of the sub-sample of 100 showed no 
significant differences among the age distributions determined 
by three of the four readers or between readings made by the 
same reader. In two trials by the same reader, exact 
agreement between age assignments occurred in over 77% of 177 
teeth. This reader showed a 0-1 year difference between age 
assignments for 98% of the teeth. None of this reader's age 
assignments were more than two years apart for a given tooth. 
Age assignments made by two different readers showed exact 
agreement for 52% of 179 teeth, a 0-1 year difference for 87%, 
and < 2 year differences for 96% of the teeth. Age 
assignments among three readers showed a 0-1 year difference 
for 73% of 85 teeth, and < 2 year differences for over 93% of 
these teeth. Table 6.3 gives more detailed results of 
comparisons between and within readers. 


Multiple teeth from about 40 otters were sectioned, 
stained, and read by four readers in order to help us evaluate 
variability among teeth of individual otters. All four first 
premolars were available for 14 otters (boiled and broken 
teeth were excluded). Comparing ages assigned by individual 
readers with the modal age (across readers and teeth) for an 
individual otter, we found that over 67% of the assigned ages 
for 406 teeth were exactly the same as the mode, over 90% were 
0-1 year away from the mode, and 96% were < 2 years away from 
the mode (Table 6.3). These results compare well with age 
estimates of duplicate teeth from other species (Table 6.2). 


Boiling skulls to facilitate cleaning has been a standard 
practice in many museums and laboratories. Schneider (1973), 
however, noted that extensive boiling of teeth made cementum 
lines less distinct. We compared the definition or 
distinctness of lines noted for boiled and unboiled teeth in 
our sample. Among the 516 teeth for which treatment during 
preparation was known, there were 60 teeth for which the 
reader noted "indistinct lines" or "poor definition." Half 
of these 60 teeth had been boiled, whereas only 34% of the 
other 456 teeth had been boiled. This difference was 
statistically significant (Chi? Saya, Che = il, 1) <— Ms Ons) - 
thus agreeing with Schneider's findings. We also examined 
pairs of teeth from individual otters that were purposely 
processed in different ways; i.e., of two teeth taken from the 


same animal, one was boiled and the other was not. Tooth 
sections were rated within pairs on quality of cementum line 
definition. For the six pairs available, the boiled tooth 


was always rated as poorer quality than the unboiled tooth 
(sign test, P = 0.03). We concur with Schneider in recom- 


111 


TABLE 6.2 -- A comparison of animal ages determined from 
tooth cementum by Gary Matson: (1) with animals of known 
age; (2) with duplicate determinations based on a different 
tooth from the same animal; and (3) with ages estimated from 
the degree of tooth wear (reprinted from Matson's Tooth 
Cementum Age Analysis, Progress Report No. 9, Spring 1987, 
Table 1). 


Period Kind of Species Numberlin Exact 1 Year More than § . Average Age 


Comparison Sample Agreement a/ Difference Year Difference 
1978-82 Known Age Various 28 22 4 2 5.0 
1983-85 Known Age Various 14 8 5° q 5.4CA b/ 
54 KA 
1986 Known Age 8T Deer i] 1 0 0 4.5CA 
4.5 KA 
1986 Known Age WT Deer 25 22 1 2 33CA 
3.5 KA 
1986 Known Age Caribou 14 12 1 0 29CA 
29KA 
1986 Known Age Kit Fox 9 7 2 (0) 25CA 
; 21KA 
1986 Known Age 8.H. Sheep 2 0 2 0 85CA 
95 KA 
1986 Known Age R. Oster 7 7 (3) 0 A5CA 
Q5KA 
1986 Known Age Raccoon 19 14 5 (0) ZICA 
1.9KA 
1984-85 Duplicate c/ Elk, Coyote, 172 8&9 271 2 29 od/ 
Bobcat 27 
1986 Ouplicsate Bobcat 100 93 7 0 25 
: 26 
1986 Ouplicate Black Bear 138 9 34 13 52 
5.2 
71986 Ouplicate Pronghorn 10 & 2 (¢) &2 
5.0 
1983-86  Duplicate®/ — Elk 1,915 1,730 152 33 5.8 
5.8 
1986 Duplicate Ek 88 45 36 7 4.9 
; 5.0 
1986 Comparison M. Deer 192 123 58 10 23CA 
with wear 3.5 WAI 


& Matson’s age compared with age from 8 second source. 

6. CA = Cementum sage; KA = Known age. 

c A “duplicate” age determination was the analysis of 2 teeth from the same individual mammal without Matson’s knowlecge 
of the identity of the duplicates. 

Oo. The first number given is the average age according to the first analysis, and the second number is according to the 
second analysis. 

€. Two primary incisors were aged together to obtain greatest accuracy for each elk, Matson’s knew of the duplication 
while aging, and the table shows the number and size of changes meade when the ages of paired incisors differed, 

f. Upper canine teeth. Sample collected and cementum aged over 8 period of several years and aged again in 1986. 

9 WA = Aged by tooth wear and replacement, : 


mending that boiling be avoided in future specimen 
preparation. 


TABLE 6.3 -- Comparisons of sea otter age assignments based 
on counts of incremental lines in tooth cementum. 


Age assignments No. of Age assignments 
compared teeth in : 
between/among: sample * no 1-yr 2-yr >3-yr 


Cbised, we ehisaig wen Colbie ohbsay 


Trials A and B 
by reader 2 177 137 36 4 0 


Readers 1 
and 2A 179 93 62 16 8 


Readers 1, 
2A, and 3 85 37 25 17 6 


Readers 1, 2A, I 
3, and 4 85 16 29 14 26 


Teeth from the 
same otter 406 274 93 23 6 oaks) 


* No boiled teeth were included in these comparisons. 


Finally, we evaluated the effects of air drying teeth on 
our ability to read outer annuli. For a group of otters from 
the same skull age category ("old adult"), we compared the 
average ages assigned to teeth collected "fresh" and stored 
frozen until slide preparation and teeth collected from museum 
specimens that were stored dry at room temperature. We found 
no statistically significant difference between the age 
distributions of these two groups (Chi* = 3.04, df = 3, P > 
0.25). Our sample size was small (47 teeth), however, and 
thus should not be considered an endorsement for dry shelf 
storage of teeth. 


We believe that counting cementum lines in teeth is a 
useful technique for determining ages of sea otters. We used 
tooth ages assigned by our most experienced reader for the age 
estimates of radio-tagged otters in California given in 
Chapter 2 and to obtain the age distributions of males and 
females, based on 425 teeth from dead animals of known sex in 
Fig 6.2. 


113 


Percent in Each Age Class 


FIGURE 6.2 -- Distribution of age estimates for 425 dead sea 
otters, based on incremental lines in tooth cementum. 


[] Males 


Mi Females 


Estimated Age (Years) 


114 


Literature Cited 


Calkins, Donald G., and Karl B. Schneider. 1984. Species 
account: the sea otter (Enhydra lutris). Alaska Dept. 
Fish and Game. 14 pp. 


Garshelis, David L. 1984. Age estimation of living sea 
otters. J. Wildl. Manage. 48(2) :456-463. 


Grue, Helen, and Birger Jensen. 1973. Annular structures in 
canine tooth cementum in red foxes (Vulpes vulpes L.) of 
known age. Danish Rev. Game Biol. 8(7):1-12. 


Grue, Helen, and Birger Jensen. 1979. Review of the 
formation of incremental lines in tooth cementum of 
terrestrial mammals. Danish Rev. Game Biol. 11(3):1-48. 


Grue, Helen E., and Carolyn M. King. 1984. Evaluation of age 
criteria in New Zealand stoats (Mustela erminea) of known 
age. New Zealand J. Zool. 11:437-443. 


Luna, Lee G., ed. 1968. Manual of histologic staining 
methods of the Armed Forces Institute of Pathology. 
McGraw-Hill, New York. 258 pp. 

Morejohn, G. Victor, Jack A. Ames, and David B. Lewis. 1975. 
Post mortem studies of sea otters, Enhydra lutris L., in 
California. Calif. Dept. Fish and Game, Marine Resources 
Technical Report No. 30. 82 pp. 


Schneider, Karl B. 1973. Age determination of sea otter. 
Alaska Dept. Fish and Game, Fed. Aid in Wildlife 
Restoration, Final Report, Proj. W-17-4 and W-17-5, Job 
8.10R. 23 pp. 


115 


CHAPTER 7 


ANALYSIS OF THE PRECISION AND ACCURACY OF RADIOTELEMETRY 


EQUIPMENT AND METHODS USED IN CALIFORNIA 


A. MERCURE 


NOVEMBER 30, 1988 


116 


INTRODUCTION 


The positions of the sea otters, instrumented with radio 
transmitters, off the California coast in this study were 
estimated, when possible, by triangulating on the radio 
signals. The accuracy and precision of this method were 
studied by placing a radio transmitter on a buoy, establishing 
the location of the buoy by visual methods, and taking a 
series of compass bearings on the buoy's radio signal. In 
this chapter, I consider the field techniques associated with 
the estimation of the precision and accuracy of the 
triangulation measurements on these buoys, and provide 
quantitative estimates of error for otter-location data 
collected by field personnel. 


METHODS 


General procedures 


Radio transmitters were placed inside a styrofoam buoy 
to simulate a resting sea otter and the buoy was anchored off 
the California coast in two locations (south and north 
locations) near Piedras Blancas. 


One of the first necessities of this work was to 
determine the map location of the two buoy locations. Both 
the south and the north buoy locations were determined by 
sighting through a Questar telescope from three positions 
along the shore. The coordinates of the three shore positions 
were obtained from our UTM grid on topographic maps of the 
study area. Compass bearings to the buoy were obtained by 
holding a Silva ranger compass against the Questar and reading 
the magnetic bearing from the compass. These bearings, along 
with the map coordinates of the three positions from which 
they were taken, were then used to plot the map location of 
the buoy. This method is illustrated in Fig. 7.1. 


The accuracy of this method was initially tested by 
sighting on a prominent landmark (Piedras Blancas Lighthouse) 
and recording 12 compass bearings from the Questar to the 
landmark. Subsequently, the UTM topographic map coordinates 
of the Questar position and the landmark location (in UTM 
coordinates) were obtained and the true bearing between the 
two locations was determined from these coordinates. This 
bearing was then compared to the mean of the 12 Questar 
bearings. 


Our normal field procedure when locating otters was to 
hold the compass by hand when taking-bearings, rather than 
holding it against the Questar scope. To determine any 
additional variance that may have been contributed by holding 


117 


FIGURE 7.1 -- Illustration of the method used to determine 
the location of the buoy by taking compass bearings, from 
three locations along the California coast, to the signal 
from the radio transmitter on the buoy. 


———— 


1 kilometer 


118 


the compass by hand, an additional 25 bearings to the Piedras 
Blancas Lighthouse were taken using this method. 


To check for possible effects of tides on the location 
of the buoys, we collected data on the location of the south 
buoy over several tidal cycles. One hundred twenty-one compass 
bearings using the Questar scope were taken from one position 
over a period of 300 hours. 


Once the locations of the buoys had been determined in 
UTM coordinates, our next task was to use these known 
locations of the buoys to measure our ability to record 
accurate locations of sea otters. To do this, we used the 
same techniques and equipment we normally used to record 
bearings on the instrumented otters to record bearings on 
these buoys. The telemetry equipment consisted of two Ford 
vans with directional four-element yagi antennas and Cedar 
Creek Bioelectronics Laboratory radio receivers. The antenna 
was pointed towards the direction of the strongest radio 
signal, usually taken as mid-way between the two points at 
which the signal disappeared ("nulls"), and a compass bearing 
of this direction was determined by sighting along the antenna 
with the hand-held compass. Four different observers helped 
collect bearings toward the stationary buoys. Observers took 
bearings in blocks of 50. In order to disrupt the tendency 
to mechanically repeat prior readings, the mobile receiving 
unit was moved about one meter after every two readings and 
was completely turned around after every 10 readings. 
Bearings were taken from three different positions along the 
shore to the south buoy location and from two positions to the 
north buoy location. 


The shore positions used for the south buoy location were 
Similar to those used for the collection of actual sea otter 
location data, in that the angles of the intersections of the 
bearings from these positions were generally 90 degrees or 
less. However, the two shore positions used for the north 
buoy location were chosen to test for effects of signal bounce 
and estimate the extent to which error was related to the 
distance of the shore position from the transmitter. Here, 
bearing measurements were taken from positions four and six 
kilometers from the buoy and the angle between the bearings 
taken from these two positions was almost 180 degrees. 


Establishing measures of precision and accuracy 


Several methods have been developed to estimate the 
accuracy and precision of radio signal locations determined 
from bearing data. Lenth (1981) described three methods for 
estimating the location of a radio signal from _ the 
intersection of three bearings: the maximum likelihood 
estimator, the Huber estimator and the Andrews estimator. All 


119 


three are based upon probability distributions. The Huber and 
Andrews estimators are designed to be relatively insensitive 
to outlying points. Precision is defined by Lenth (1981) on 
the basis of an iterative algorithm and a variance co-variance 
matrix. Computer program TRIANG (Garrott, et al., 1986) 
calculates these estimators and defines the area of a 95% 
confidence ellipse around them, using the methods described 
by Lenth. Detailed mathematical descriptions are presented 
in Lenth (1981). Garrott, et al., (1986) also did an 
empirical test of the three estimators and found the Andrews 
estimator to be superior to the other two estimators. 
Confidence ellipses generated from the Andrews estimator were 
more likely to include the actual transmitter location and 
were found to be more accurate, i.e., the plotted point was 
usually closer to the actual transmitter location. 


Our work with the buoys gave me a sample of bearings on 
a known location from several positions along the coast. 
These bearings did not correspond exactly to those used to 
locate otters, since they were repeated bearing measures on 
a known location. To more closely simulate field data, I took 
random samples of these bearings and used them to represent 
bearings that would have been recorded during field 
operations, if the buoy had been a resting sea otter. 


One hundred bearings from each of the three shore 
positions for the south buoy were randomly chosen and one 
bearing from each position was randomly chosen and combined 
into a set of three. These 100 sets of three bearings were 
then used to triangulate the location of the transmitter and 
estimate precision and accuracy for the south buoy location. 


The angle generated from bearings taken from the two 
shore positions for the north buoy approached 180 degrees, 
thus these data could not be used, as recorded, to estimate 
precision and accuracy. However, I was able to use the 
distribution of errors (degrees difference from the actual 
bearing) of the bearings taken on the north buoy. These error 
measures were placed in a random order and 50 sets of three 
were randomly drawn. Three positions along the California 
coast near the north buoy location were then selected so that 
the distance between them represented the maximum distance 
between the positions normally used when locating otters. The 
actual bearings from these positions to the north buoy were 
calculated from our topographic maps and the sets of errors 
were then added to the actual bearings to simulate field data. 


Precision was defined as the 95% confidence ellipse, and 
accuracy was estimated by calculating the distance from the 
plotted point to the actual transmitter location. 
Calculations were made using the Andrews estimator and 
computer program TRIANG (Garrott, et al., 1986). 


120 


Comparisons of methods for estimating location from bearings 


After the bearing data to the two buoy locations were 
generated, it was then necessary to compare the location 
estimates produced by our field method of plotting data and 
the Andrews estimator. Our field method consisted of plotting 
at least two, but preferably three, bearings to an otter on 
the topographic maps and then assigning a position in UTM 
coordinates to the point at the intersection of two bearings 
or in the center of the triangle formed by three bearings. To 
compare this method with the Andrews estimator, the 100 sets 
of bearings toward the south buoy and the 50 sets of simulated 
bearings toward the north buoy were plotted by the field 
method. The resulting location estimates and the Andrews 
estimators for the same sets of bearings were then compared 
to the actual buoy locations. 


A direct comparison of the individual location estimates 
produced by the field method and the Andrews estimator was 
made using 50 otter locations, randomly chosen from our otter 
position data plotted in the field. These locations were 
recalculated by entering the bearings obtained in the field 
into a computer and again using program TRIANG to obtain the 
Andrews estimator. 


RESULTS 


Accuracy of determining buoy locations visually 


The difference between the mean of the bearings to the 
lighthouse taken with the compass held against the Questar and 
the calculated true bearing was 0.6 degrees, with a standard 
deviation of 0.4 degrees. The difference between the mean of 
the hand-held compass bearings toward the lighthouse and the 
calculated true bearings was 1.6 degrees, with a standard 
deviation of 1.2 degrees. These data are summarized in 
Figures 7.2 and 7.3. It was not possible to read the compass 
to an accuracy greater than one degree. 


The bearings taken to evaluate the possible effects of 
buoy movement due to changing tides are plotted against time 
aligy ARSC PG eI TAG CNG No movements of the buoy over time were 
apparent. 


Accuracy of bearings towards the radio signal from the buoys 


The difference between the mean of the 897 hand-held 
compass bearings to the radio signals on the buoys and the 
calculated true bearing was 0.5 degrees (Table 7.1). The 
bearing error did not differ significantly with buoy location, 
shore position, van, or observer (Table 7.1). The 


121 


distribution of the differences between these bearings taken 
by field personnel and the calculated true bearings is shown 
nlp SEKe fa PY/6 Sn 


Precision and accuracy of triangulations on buoys 


Table 7.2 summarizes calculations of the precision and 
accuracy of our simulated locations of the north and south 
buoys through radiotelemetry. Precision, defined as the 95% 
confidence ellipse, was estimated as between 0.03 and 0.08 
hectares. The mean accuracies of our estimations of the 
locations of the buoy were 51 meters for the south location 
and 110 meters for the north location. 


Comparison of field-method estimates with Andrews estimator 


The results obtained by hand-plotting the 100 sets of 
bearings to the south buoy and the 50 sets of bearings to the 
north buoy and the Andrews estimator for these same data sets 
are compared in Table 7.3. Although the two methods gave 
similar results, in that, on the average, the estimated 
locations were about equally close to the true location of 
the radio signal, the actual data points calculated by the two 
methods were different. The mean difference between the 50 
field data points and the Andrews estimator for the same 
bearings was 162 meters with a standard deviation of 104 
meters. The reason for this difference is that the two 
procedures use different methods of weighting the bearings. 
The Andrews estimator is designed to be robust to outliers and 
thus assigns different weightings to individual bearings based 
upon a probability distribution. However, field personnel 
usually plotted the location of the otter in the middle of the 
triangle formed by the three bearings and thus gave equal 
weight to each of three bearings. 


DISCUSSION 


Errors in determining the true locations of the buoy by 
visual sightings with the Questar were minimal and appeared 
to be the result of our inability to read the compass to an 
accuracy of less than one degree and/or the limitations of our 
ability to locate points precisely on the topographic maps. 
The positions of the buoys did not change with the changing 
tides. 


The accuracy of our bearings compared quite favorably 


with that of those taken in other radiotelemetry studies. 
Several studies have reported mean errors and standard 


122 


Number of Bearings 


PLCUREOW (20 ou Disticlome tony ote | 2) compass) sbeatingse toma 
prominent landmark sighted through a telescope (Questar). 
The actual bearing of the landmark (zero degrees in the 
figure) was calculated from the positions of the telescope 
and the landmark on a topographic map. 


-20 -16 =| =3) -4 0 4 3 12 16 20 


Visual Questar Bearings Deviation From Actual 


In Degrees 


12% 


FIGURE 7.3 -- Distribution of 25 hand-held compass bearings 
to a prominent landmark. The actual bearing to the landmark 
was calculated as in Figure 7.1. 


Number of Bearings 


-20 -16 S122 -8 4 0 4 8 12 16 20 


Hand Held Compass Bearings Deviation From Actual 


In Degrees 


124 


Degrees 


226 


225 


224 


_ 225 


FIGURE 7.4 -- Compass bearings to a radio transmitter on a 
buoy off the California coast, taken over a three-hour period 
when the tide was changing. 


100 200 300 400 


Time in Hours 


Number of Bearings 


FIGURE 7.5 -- Distribution of 1125 hand-held compass bearings 


to the signals from radio transmitters on buoys off the 
California coast. 


120 


100 


50 


60 


40 


207 


=20%m =16 12) @Oe-2) 1 ea 0 4 er Hebi2 Ke O 
All Telemetry Bearings Deviation From Actual 


In Degrees 


TABLE 7.1 - Summary statistics for the 897 hand-held compass 
bearings to the signals from the radio transmitters on buoys 


off the California coast. 


N Mean difference Standard 
from true bearing deviation 
(degrees) (degrees) 

All bearings 897 0.5 4.2 
By buoy location: 

South 447 -0.7 4.4 

North 450 1.6 4.0 
By van: 

1 597 0.5 4.3 

2 300 0.5 4.1 
By observer: 

1 50 -3.6 1.7 

2 250 (0) 6 al 3.8 

3 297 1.6 4.4 

4 300 0.6 4.2 
By shore position: 

1 (south buoy) 247 -1.4 4.1 

2 (south buoy) 100 0.3 4.4 

3 (south buoy) 100 0.3 Siew 

4 (north buoy) 300 2.0 4.1 

5 (north buoy) 150 0.8 365 


127 


TABLE 7.2 - Summary of the calculated precision and accuracy 
of the methods used to judge the locations of radio 
transmitters on buoys, based on hand-held compass bearings to 
the direction of the radio signal and details of the locations 
of the transmitters on the buoys and the mobile receivers near 
which the bearings were taken. 


A. Precision 


South buoy location 


N 100 
mean size of the 95% confidence ellipse -03 hectares 
standard deviation -O1 hectares 


Simulated North buoy location 


N 50 

mean size of the 95% confidence ellipse -08 hectares 

standard deviation -02 hectares 
Accuracy 


South buoy location 


N 100 
mean deviation from actual transmitter 

location 51 meters 
standard deviation 32 meters 


Simulated North buoy location 


N 50 
mean deviation from actual transmitter 

location 110 meters 
standard deviation 66 meters 


B. Distance to transmitter 


South buoy location 


receiver location 1 
receiver location 2 
receiver location 3 


Simulated North buoy location 
receiver location 1 
receiver location 2 
receiver location 3 

Degrees between locations 


South buoy location 


receiver location 1-2 
receiver location 2-3 


Simulated North buoy location 


receiver location 1-2 
receiver location 2-3 


700 meters 
600 meters 
613 meters 


1642 meters 
780 meters 
1262 meters 


34 degrees 
78 degrees 


74 degrees 
52 degrees 


—aaaaaaaBnBnBnBnBanBnBnDBanBn9n9nBnBRaRe 


TABLE 7.3 - Comparisons between the field method of plotting 
data and the Andrews estimator for the same set of data. 


Field 
method 
South buoy location 
N 100 
mean deviation from actual 
transmitter location 51 meters 
standard deviation 37 meters 
North buoy location 
N 50 
mean deviation from actual 
transmitter location 119 meters 
standard deviation 57 meters 


129 


100 


51 


32 


50 


110 
66 


Andrews 
estimator 


meters 
meters 


meters 
meters 


deviations greater than we observed. For example, in flat 
terrain, Hupp and Rati (1983) recorded mean errors between 0.4 
and 3.2 degrees, with standard deviations of 1.3-5.0 degrees. 
In areas with mountains and trees, their estimates of mean 
error were greater: 4.5 to 28.2 degrees, with standard 
deviations of 52.7-83.8 degrees. Lee, et al., (1985), taking 
bearings from fixed towers, reported mean errors between 1.76 
and 5.27 degrees (after removing all bearing errors greater 
than 10 degrees from the sample); Brewer (1983) found that 
25%-40% of his bearings were unusable due to the inability to 
distinguish between direct and reflected signals, and Garrott, 
et al., (1986) noted that 52% of transmitter locations that 
were not along the line-of-sight to the position from which 
bearings were taken, resulted in bearings with large mean 
errors and/or large standard deviations because of signal 
reflection. 


As shown by the results of these other studies, signal 
bounce and interference from rugged terrain and obstructions 
are often major problems during field studies using telemetry. 
The preferred position to take bearings toward a signal is 
from a topographically elevated site with a direct and 
unobstructed line-of-sight to the transmitter. In our study, 
we often had nearly ideal conditions for telemetry, as many 
of the positions along the coast from which we took bearings 
were located at the top of shoreline cliffs, above the 
transmitters in the otters on the surface of the ocean, and 
there were no obstructions between the receiver and the 
transmitter. These conditions minimized the possibility for 
signal bounce. The bearings which were taken toward the north 
buoy were taken from positions chosen to be most likely to 
produce signal bounce. Specifically, the positions of the 
tracking vans with the receivers were 4 and 6 kilometers from 
the buoy, so that the bearing direction to the buoy was almost 
parallel to the general direction of the coast and provided 
the maximum possibility for interference from intervening land 
forms. As the error of bearings taken under these conditions 
was not significantly different from those taken towards the 
south buoy location, where the bearing direction was towards 
the open ocean and there were no obstructions, we believe that 
signal bounce was not a significant source of error in the 
collection of telemetry location data on the California sea 
otter. 


The accuracy and precision that we calculated are 
applicable to otters located within approximately 800 meters 
from shore. At least 75 percent of the plotted locations of 
adult male and female and juvenile female otters fell into 
this category (Chapter 3, Fig. 3.8). However, since juvenile 
males were often located more than 800 meters from shore, the 
accuracy and precision of our locations for this class of 


130 


otters are probably considerably worse than our calculated 
values. 


It should also be realized that our calculations of 
accuracy and precision represent the error present under ideal 
conditions, when the radio signals are clear, continuous, and 
strong and observers are attempting to take optimum readings. 
Under field conditions, signals are frequently interrupted or 
weak -- this is particularly likely to be true for those from 
juvenile males far offshore -- and observers are sometimes 
tired or hurried. 


The locations estimated by the field method of hand- 
plotting, using bearings from the buoy data sets, were as 
close to the actual buoy location as the Andrews estimator 
calculated by program TRIANG (Table 7.3). However, the 
estimate of the average distance between the hand-plotted 
locations and the Andrews estimator, using bearings on actual 
otters, was 162 meters. However, this mean difference cannot 
be used to quantify the error of the otter locations we 
estimated in the field, as there are a variety of possible 
spatial relationships between the two estimated locations. For 
example, if the actual location of the sea otter is in between 
the location estimated by the field method and the one 
indicated by the Andrews estimator, then the difference 
between the location plotted in the field and the one 
resulting from Andrews estimator will exceed the distance from 
either of these estimated points to the actual otter location. 
Conversely, if both estimated points lie on the same side of 
the otter's actual location, then the distance between the 
Andrews estimator and the field estimate could be less than 
the distance of either to the otter. The distributions of 
both the points produced by the Andrews estimator and those 
produced by the method we used in the field in relation to 
actual otter locations are unknown. The distribution of the 
Andrew estimator would be dependant upon the differential 
weighting to bearings given by program TRIANG. We believe, 
therefore, that it would be inappropriate to add the 162-meter 
mean difference between these two estimates to our estimate 
of accuracy. 


LITERATURE CITED 
Andrews, D.F., P.J. Bickel, F.R. Hampel, P.J. Huber, W.H. 


Rogers and J.W. Tukey. 1972. Robust Estimates of 
Location: Survey and Advances, Princeton University Press. 


Brewer, L.W. 1983. Radio tracking the spotted owl in 
Washington state. A discussion of equipment and 
technique. In Proceedings 4th International Wildlife 


Biotelemetry Conference, Ed. D.G. Piniock, Applied 
Microelectronics Institute, Halifax, Nova Scotia. 


131 


Garrott, R.A., G.C. White, R.M. Bartmann and D.L. Weybright. 
1986. Reflected signal bias in biotelemetry triangulation 
systems. J. Wildl. Manage. 50:747-752. 


Lee, J.E., G.C. White, R.A. Garrott, R.M. Bartmann and A.W. 
Alldredge. 1985. Accessing the accuracy of a 
radiotelemetry system for estimating animal locations. J. 
Wildl. Manage. 49:658-674. 


Lenth, R.V. 1981. On finding the source of a signal. 
Technometrics 23:149-154. 


Springer, J.T. 1979. Some Sources of Bias and Sampling Error 
in Radio Triangulation, J. Wildlife Mgmt. 43:4 pp 926- 
935. 


132 


CHAPTER 8 


MOVEMENT PATTERNS OF ADULT FEMALE AND WEANLING 


SEA OTTERS IN PRINCE WILLIAM SOUND, ALASKA 


C. MONNETT AND L. ROTTERMAN 


NOVEMBER 30, 1988 


133 


INTRODUCTION 


The tendency for sea otters (Enhydra lutris) to exhibit 
spatial segregation of the sexes is a well established feature 
of their social system (Lensink 1962; Kenyon 1969; Peterson 
& Odemar 1969; Schneider 1978; Garshelis & Garshelis 1984; 
Garshelis, Johnson & Garshelis 1984). Males and females 
segregate into geographically discrete portions of habitat 
that are generally referred to as "male areas" and "female 
areas" (Kenyon 1969, p. 208) or "breeding areas" (Garshelis, 
Johnson & Garshelis 1984, p. 2648). Male areas are occupied 
almost exclusively by males of all ages (e.g. Kenyon 1969; 
Garshelis, Johnson & Garshelis 1984) whereas, female areas 
tend to contain a mixture of mature males and females of all 
ages. 


The movement patterns of mature males and/or males in the 
male areas are relatively well understood as a result of 
tagging and short-term radio-telemetry studies. Young males 
are born in female areas. In Alaska they apparently leave 
their natal female areas shortly after weaning. They move to 
and reside within a single male area or travel among several 
male areas until maturity (e.g. Kenyon 1969; Garshelis, 
Johnson & Garshelis 1984), which occurs at about 5-6 years 
of age (Green 1978; Schneider 1978; Garshelis 1983). As 
adults, males may re-enter the female areas, wherein they may 
employ one of two non-mutually exclusive reproductive 
strategies. The most conspicuous of these two strategies is 
that of males that defend territories. Territorial males may 
copulate in serial fashion with females that enter their 
territories (Vandevere 1970; Calkins & Lent 1975; Loughlin 
1977; Garshelis & Garshelis 1984; Garshelis, Johnson & 
Garshelis 1984; but see Kenyon 1969). Other males, or 
possibly the same males on _ other occasions, may 
opportunistically search for and attempt to pair and/or 
copulate with females (Kenyon 1969). Reproductive activities 
normally are concentrated during the fall but some males may 
remain on their territories year-around. Others return to the 
male areas (Garshelis, Johnson & Garshelis 1984) where they 
rejoin the male aggregations and remain until the following 
breeding season. 


Less is known about the movement patterns or distribution 
of females that reside within the female areas. Short-term 
studies of individuals, using radio-telemetry, suggest that 
females are somewhat less mobile and less gregarious than 
their male counterparts (e.g. Garshelis, Johnson & Garshelis 
1984). However, available data have been inadequate to 
indicate how the female areas are used by individuals or to 
permit evaluation of variation in females' movement patterns 
associated with seasonal or functional (e.g. breeding, 
pupping, wintering, etc.) needs. Also, little information has 


134 


been available on the movements of weanlings and on the manner 
in which they become established within their respective male 
and female areas. 


This chapter describes the movement and habitat use 
patterns of mature female and immature male and female sea 
otters in Prince William Sound, Alaska. The movements of 
adult females are examined at different stages of the 
reproductive cycle. The movement patterns of weanling males 
and weanling females are contrasted and discussed in the 
context of the evolution of dispersal patterns. The 
relationships between the observed movement patterns and the 
sea otter's social system are considered. 


STUDY AREA AND METHODS 


The study was located in the eastern portion of Prince 
William Sound, in south-central Alaska (Figure 8.1). During 
the past decade, sea otters have recolonized the deep bays, 
mud flats and channels that are located to the west of the 
fishing community of Cordova. The local population of sea 
otters, its history and its habitat have been described by 
various authors (Gabkinsycs bent, 19757 sSinith, (Williams! 
Johnson & Garshelis 1982; Garshelis & Siniff 1983; Garshelis 
1983; Garshelis & Garshelis 1984; Garshelis, Johnson & 
Garshelis 1984; Garshelis, Garshelis & Kimker 1986). 


Data were collected during 18 months between June, 1984 
and October, 1986. The subjects included 8 adult females and 
35 pups (and indirectly their mothers) from 2 cohorts; 14 
during 1984 and 21 during 1985. All otters were captured in 
Sheep Bay or Simpson Bay. Adults and a few dependent pups 
were captured in floating tangle nets (91 m long by 5 m deep 
with a 22 cm stretch mesh) during June, August or September 
1984. Most dependent pups were captured in dip nets during 
August or September, 1984 or 1985. Pups ranged in size from 
8 - 20 kg and all still accompanied their mothers. Upon 
capture, otters were brought aboard a 5.5 m skiff and 
immobilized with a combination of fentanyl (0.05 mg/kg) and 
azaperone (0.20 mg/kg) (Williams, Williams & Siniff 1981). 
Each otter was weighed and its sex was recorded. One or more 
nylon tags were inserted through the interdigital webbing of 
one, or both, hind flippers for identification. A 
radio-transmitter was surgically implanted in each animal's 
peritoneal cavity by a veterinarian, as described in Chapter 
abe 

Radio-instrumented sea otters were monitored during 
daylight in August-October and December 1984; April-December 
1985; February, May-June and October 1986. Visual 
observations were made from a skiff or from the shore with 
binoculars or 50-80X telescopes (Questar Corp., New Hope, PA, 
USA 18938). Instrumented otters were normally monitored from 


135 


1984-1987. 


ALASKA 
S 


PRINCE WILLIAM 
SOUND 
? fi stupy area 
I } 
| 7 Dafa, VALDEZ 


‘fe & 


+ ae 
’ rs fii . is WE 


——— 4 
| = © 
—— 5 yi 

ETrra, SP 
%s, 
Ee) 


a skiff that was traveling at 20-30 knots. fThe skiff was 
equipped with 2 yagi antennas mounted on 4-m aluminum masts. 
Antennas were attached at 60 and 300 degree angles from the 
plane of the boat. Periodically, (approximately 250 total 
flight hours), instrumented sea otters were monitored from 
small aircraft that were equipped with 4-element yagi antennas 
mounted under each wing (Gilmer et al., 1981). Aircraft 
speed was set at about 100 knots and preferred altitude was 
600-750 mn. 


Radio-transmitter frequencies were scanned on a 2000 
channel programmable scanning receiver (Cedar Creek Lab). 
Radio fixes were determined by triangulation or by moving the 
boat in the direction of the radio signal until the individual 
was observed. Otter locations (fixes) were recorded either 
as coordinates of the Universal Transverse Mercator Grid 
System, or marked directly on large scale maps or tracings of 
the various bays and channels. The latter were used 
predominantly during aerial surveys. Distances were measured 
on U.S.G.S. scale 1:250,000 or 1:63,360 contour maps. 


As Garshelis & Garshelis (1984) pointed out, the annual 
home range of Prince William Sound sea otters is composed of 
numerous centers of activity connected by long travel 
corridors. The area of any portion of the annual home range, 
or rather, any cluster of fixes, can be estimated by measuring 
the area of the minimum convex polygon enclosing the fixes 
(Odum & Kuenzler 1955; Garshelis & Garshelis 1984). In Prince 
William Sound, sea otter travel corridors often cross, and 
enclose, deep, broad, and presumably, inhospitable expanses 
of water. As a consequence, the same procedure, when applied 
to estimation of annual home ranges, drastically over 
estimates areas. The large number of fixes required for 
characterization of such habitat utilization patterns, at 
least 40 per activity center (Garshelis & Garshelis 1984), 
makes an accurate measurement of annual, or longer-term, home 
range impractical. Garshelis & Garshelis (1984) suggested an 


index of home range: "distance between extreme locations" 
(DBEL). Herein, it is used to describe the magnitude of the 
movements of individuals. The distance between extreme 


locations is the minimum distance an otter would have to swim 
to go between its two most widely spaced fixes during some 
time interval. It is approximately equivalent to the maximum 
dimension of the home range (Garshelis & Garshelis 1984). 


RESULTS 


Adult female home ranges 


Eight adult females were implanted and monitored for 
periods ranging from 15-20 months. All eight females survived 
the duration of the study. Four gave birth to pups. All 


137 


study females traveled extensively throughout the eastern 
Prince William Sound. The median distance between extreme 
locations of the eight females was 41 km (range 27 - 85 km) 
(Figure 8.2). 


Some females made long, circuitous trips which crossed 
major bodies of water. During the summer of 1985, one female 
(84001) traveled beyond the limits of the area that was 
routinely monitored (Figure 8.3). Contact was lost in 
mid-May and reestablished on October 31. On that date she was 
near Green Island. She had returned to Sheep Bay by November 
7. The short time interval between sightings, at two distant 
locations, suggests that Hinchenbrook Entrance was traversed. 
Hinchenbrook Entrance is a channel that spans 11.5 km at its 
narrowest and is over 300 m deep at its shallowest crossing. 
It has rapid tidal currents and intemperate conditions. The 
only alternative to crossing that channel, or other comparably 
deep, broad channels would have been for female 84001 to have 
circumnavigated Prince William Sound, a minimum trip in excess 
of 200 kn. All study females traveled between major bays. 
Most traveled across large expanses of deep water. However, 
if other study females journeyed beyond the limits of the 
regularly monitored portions of the study area, they must have 
done so for only brief periods, since all were located 
regularly with no comparable periods of lost contact. 


The study area was divided into 12 habitat zones in order 
to illustrate the movements of the eight study females 
(Figure 8.4). All of the females traveled in four, or more, 
zones (range 4 - 9) during the time they were monitored 
(Figure 8.5). 


Females tended to occupy the western portions of the 
study area during the late spring and summer but to travel to 
the easternmost area where they spent the late fall and winter 
(Figure 8.6). The eight radio-implanted females used 
superzone A heavily during May, June and July. At that time 
most females were pupping or tending small pups (Chapter 9). 
Females were aggregated into rafts (often containing over 100 
individuals) in shallow, protected coves and over shoals. 
During the fall, many females moved into the bays on the north 
side of the study area (superzone B). They formed less dense 
aggregations, weaned their pups and presumably, in some cases, 
mated with resident males. As winter approached, females 
became rare in superzone A. This may have been in response 
to winter storms which often batter those coastlines from the 
east or northeast. During the late fall and winter, females 
became abundant in zones 5, 8, 10 and the western portion of 
zone 9. 


138 


ADULT FEMALES 


84001 V7 31/604 
soos 93/546 

84005 50/486 
soos, 301848 

4007 | re 

saoiol WJ“, 80542 
sis} 35481 

asoi7 J 4788 


i 20. Ao so eo TOI co 
DISTANCE BETWEEN 
EXTREME LOCATIONS (km) 


FIGURE 8.3 -- Movements of an adult female sea otter in 
Prince William Sound, Alaska, during a 20 month interval, 
June 1984 - February 1986. Summers were spent in the western 
portion of the study area and winters in the eastern portion, 
near the Cordova male area. 


31 LOCATIONS ya 
v7 
7 
7 
7 
? 
? 
Y 
? ae 
ye ee) 
7 ge 
a a 
7 ys 
Vie 
7 oO 
12 

GREEN We 
ISLAND 
S cb 

3 

7) se oo 10 20 
_———————S ee | 


Kilometers 


140 


FIGURE 8.4 -- Division of study area in Prince William Sound, 
Alaska, into numerically designated habitat zones and 
superzones. Zones correspond to major bays or passages. 


~~ SUPER ZONES 
eS 


141 


FIGURE 8.5 -- Use of habitat zones in Prince William Sound, 
Alaska, by eight radio-instrumented adult female sea otters. 


1 Be oh SH ts Selva 


cen | 7/7 
4004 JV yy 

84005 |e FF 

84006 end 

e007 | L777, LZ 
84010) V77/77/7|— 

aos | (7, V7 


e417 | U7 | | 
12345678 9101112 
HABITAT ZONES 


FIGURE 8.6 -- Seasonal changes in the use of portions of 
eastern Prince William Sound by eight radio-instrumented 
adult female sea otters. Superzones are delineated on Figure 
8.4. 


100% 
DEC 84 
a fe) 
100% 
APRIL 85 
— O 
100% 
MAY 85 pee 
0 
100% ,, 
z 
100% i 
JULY 85 as ee 5 
0 2 
100% & 
AUG 85 i Zz 
= 
0 se 
‘ 100% © 
0 O 
100% 32 
OCT 85 Pe wy. 
9) 
100% & 
100% 
DEC 85 
| 0 
100% 
FEB 86 
io... te) 


AneaB ihe Cinti2) 
SUPER ZONES 


143 


FIGURE 8.7 -- Distance between extreme locations of 26 female 
sea otters in Prince William Sound, Alaska, that were 
accompanied by dependent pups. Most observations are based 
on females accompanying radio-instrumented pups. 


FEMALE PUP PAIRS 
p> 


"o 10 20 30 40 50 60 
DISTANCE BETWEEN 
EXTREME LOCATIONS (km) 


144 


Females with pups 


Females traveled extensively while they were accompanied 


by a pup (Figure 8.7). The median distance between extreme 
locations for females with pups born during 1984 (m = 17.25 
km, range = 6.5 - 38.5, n = 10) was shorter than that of 


females with pups born during 1985 (m = 33 km, range = 15 - 
6275),-ene =" ay) This apparent difference probably resulted 
from the fact that there are only limited data available on 
the movements of several of the individuals from 1984. 
Monitoring was discontinued between late September - mid 
December, 1984. Consequently, fewer telemetry fixes were 
available for assessing individual movements during that 
period than in 1985. 


Females apparently only occupied a portion of their 
annual home range while they were accompanied by pups. Based 
on data collected on radio- implanted females during 1985, the 
DBEL of females during the time interval when they were 
accompanied by pups were smaller than the annual DBEL of the 
radio-implanted adult females (m = 33 km vs. m = 41 kn, 
Mann-Whitney U test, p < .02). This was probably because the 
formers' home ranges did not include the wintering areas in 
the eastern portion of the Sound. Accompanied females 
confined their movements to trips between the western nursery 
areas and the north-central bays where most weaning took 
place. Weaning occurred before movement into the wintering 
areas. 


Movements during the last month before weaning.--During 
the last month before a pup was weaned, the female and pup 
usually occupied a relatively small, shallow cove or channel. 
It can be inferred that the pair's home range was smaller, 
since the distance between extreme locations was shorter than 
it was during the earlier portion of the dependency period 
(Figure 8.8). Data for both sexes of pups are pooled for 
analysis, since data for male and females are similar (last 
30 days: t = -0.17, N.S.; earlier interval t = -0.91, N.S.). 


-Potentially, there are two ways that the observed 
differences in home range size could be an artifact of 
sampling design. The first relates to the relative sample 
sizes, the second to differences in timing between the 
respective samples. 


First, it has previously been shown that home range area 
is correlated with the number of fixes analyzed when sample 
sizes are small (under 40 fixes) (Garshelis & Garshelis 1984). 
Those authors also found estimated home range area to be 
correlated with monitoring interval. Thus, if the number of 
fixes and monitoring intervals were not about the same in the 
"last 30 days" and "earlier" samples, any differences could 


145 


be an artifact. Two arguments can be made against such a 
problem distorting patterns in the data from this study. One 
is that the sampling interval and the number of fixes were 
similar in the two treatments described; sampling interval: 
last 30 days, mean = 25.6 days (SD = 4.5), earlier, mean = 
28.6 days (SD = 11.6); number of fixes: last 30 days, mean 
= 8.9 (SD = 3.0), earlier, mean = 10.2 (SD = 4.0). The other 
is that no correlation existed between extreme locations 
within the pooled samples (r = -0.03, N.S.; r = .21, N.S., 
respectively). 


Second, since most weaning took place in the late fall 
(Chapter 9), one reasonable argument could be that shrinking 
home ranges result from the tendency for adult sea otter 
females to move into protected areas and to restrict their 
movements at the onset of winter weather patterns. MThis does 
not appear to be the case. Independent females continued to 
travel, and thus, had large home ranges during the late fall 
and winter. The distance between extreme locations was 
longer for independent females, at that time, than it was 
during the last 30 days before weaning for female-pup pairs 
(m = 19.5 km, n = 8 vs. 7.5 km, n = 19, respectively; 
Mann-Whitney U-test p < .02). 


Sexual differences in adult home range size 


A direct comparison can be made between the extent of 
movements of adult males and adult females, within 
northeastern Prince William Sound, by combining the results 
of this study with those of the earlier studies of Garshelis 
and Garshelis (1984, p. 674, Fig. 7). Those authors argued 
that male home ranges were larger than female home ranges in 
Prince William Sound. However, they pointed out that at 
least part of the difference could have been due to females' 
movements being constrained by geographic boundaries. That 
is, female areas at Green Island, in central Prince William 
Sound, were smaller than male areas in Nelson Bay, in 
northeastern Prince William Sound. Thus, a comparison of 
males and females in the same general area (i.e. within 
northeastern Prince William Sound) should be a better test for 
sex differences in movements. 


As mentioned above, the relative number of fixes during 
each study and the durations of the studies could affect the 
results. Thus, an attempt was made to ensure that the data 
from the two studies were directly comparable. Only data from 
the July - Sept. interval, 1-3 months of monitoring, are 
considered. However, data from the latter study includes that 
on both independent females and females with pups. 


To test for differences in movements of adults, the 
proportions of males and females in 2 distance categories 


146 


FIGURE 8.8 -- Changes in the home ranges of sea otter female- 
pup pairs in Prince William Sound, Alaska, that occur as the 
pups approach weaning age. The distances between extreme 
locations of pairs are compared for the last 30 days before 
weaning and for the earlier period when the pup was younger. 


ener" yyy 
“... Wy 
en ...... AA 
Se LLL 
eS Yj: 
a GUM 
MH LAST 30. DAYS 
aE al LE 
ZZ, t-6.05 , p<.001 
A WY 
S BMEIZIZZZZZ 
ee = ede 


OM oes Ores Seere0 seo 30). G0. 40 40 


DISTANCE BETWEEN 
EXTREME LOCATIONS (km) 


147 


were compared: individuals with DBEL < 15 km and those with 
DBEL > 15 km. The DBEL of adult females were larger than 
those of adult males (respective ranges 15 -60 km vs. 4.8 - 
37 km; Chi square = 14.31, 1 DF, p < .001). 


The sea otter population of Prince William Sound is still 
increasing after near extirpation by fur traders during the 
end of the 18th century (Lensink 1962). The history of the 
local remnant population is fairly well documented. Simpson 
Bay, in zone 4, and superzone C went through a transition from 
being a male area to being a female area in the early 1980's 
(Garshelis & Garshelis 1984; Garshelis, Garshelis & Kimker 
1986; this study). By 1986 the male area was entirely 
contained within superzone D. The eastern Prince William 
Sound contained but a single, well consolidated male area 
which was surrounded by female areas. At its eastern edge, 
in zone 10, the population was dominated by females. That 
area was heavily used by females, some of which were 
accompanied by dependent pups, during the winter and spring 
1984-1987. 


Movements of weanlings 


Most pups were born during May and weaned during the 
subsequent fall (Chapter 9). Two characteristic movement 
patterns were exhibited by sea otters during their first year 
following weaning. Some weanlings stayed within, or very 
close to, the home range they had occupied during the month 
preceding weaning. However, most weanlings immediately made 
a relatively large movement, then occupied a small home range 
until spring. During spring, they expanded their home range. 
Their movements within the extended home range took them still 
further from their site of weaning. 


By the end of monitoring (maximum 21 mo.) many of the 
weanlings had traveled far from the site at which they 
separated from their mothers (Figure 8.9). Males had moved 
longer distances than their female counterparts. Mortality 
was high (Chapter 9). Many of the weanlings' trips had 
culminated with their death. The next few sections deal with 
sea otter behavior during their first year of independence. 


If weanlings moved significantly from their site of 
weaning, they almost always began their travels abruptly, 
within 2 weeks after weaning. For example, 14 weanlings in 
the 1985 cohort traveled 20 km or more. Within their first 
two weeks of independence, 13 of the 14 had moved at least 20 
km from their weaning location. The 14th weanling also moved 
abruptly but did not do so until about two months had passed. 
Most weanlings departed almost immediately following 
separation. Since departures were abrupt and movements tended 
to be fairly long, contact was usually lost until a search 


148 


FIGURE 8.9 -- Distances traveled from the site of weaning in 
Prince William Sound, Alaska, -by male and female weanling sea 
otters. Monitoring interval varied from a few days to 
approximately 18 months. Short monitoring intervals resulted 
when pups died during their travels. 


BB Mates 


AW FEMALES 
MANN-WHITNEY 


P<.002 


WEANLINGS 


0 10 20 30 40 50 60 70 ,80 90 100 110 120 
DISTANCE (km) FROM WEANING LOCATION 


could be made from an aircraft. By the time that the first 
post-weaning radio fix was taken, the weanling's travels were 
usually completed. 


Most traveled at least 20 km from where they were weaned 
to their post-weaning home range. During their first trip 
males tended to travel further than females (Figure 8.10). 
Weanlings usually completed their first trip quickly. Data 
on three weanlings that were weaned during the first week of 
November, 1985, illustrate this point. Two of these weanlings 
were males that were weaned on, or at most, a few days before 
11/4 and 11/6, respectively. The third was a female that was 
weaned during the night of 11/6. All departed from their 
pre-weaning home ranges in Sheep Bay or Simpson Bay on the 
night rote 1 7/6). Those bays were searched thoroughly from a 
skiff on 11/7. None of the three were found within the 10 km 
search radius. An aerial survey was flown on 11/9. The 
female was found 38 km to the southwest on the far side of 
Orca Bay. The two males were not located on 11/9. At that 
time they were not within 50 km of their weaning location. 
The search area was expanded during a second aerial survey on 
11/16. One male was found near Valdez, at a distance of 123 
km. The other male had traveled about 80 km along the same 
coastline. On the next aerial survey of that area, a few 
weeks later, he was also near Valdez, 109 km from the place 
where he was weaned. 


The extent of weanlings' movements that defined the 
post-weaning home ranges appeared to vary substantially 
between individuals. The distance between extreme locations 
was greater for males than females, however, not significantly 
(Figure 8.11). 


The process of segregation of sea otters into male and 
female areas figures prominently in early sea otter behavior. 
When weanling males travel to their first post-weaning home 
ranges they usually leave the female areas in which they were 
reared (Figure 8.12). Conversely, young females usually do 
not leave. In this study, three young females had home ranges 
outside the female area during their first winter. Two of 
these females survived until spring, at which time both 
returned to the female area. Only two of twelve males known 
to have survived their first winter were not within the male 
area. 


DISCUSSION 


The data given in this paper indicate that the extent of 
movement by sea otters in eastern Prince William Sound varies 
with age, sex and reproductive status. Relatedly, otters use 
specific portions of their habitat for different purposes. 
Thus, densities within a given area can change dramatically 


150 


FIGURE 8.10 -- Distance between weaning of sea otters 
location in Prince William Sound, Alaska, and their first 
post-weaning home range. The distance was traveled in a 
single relatively rapid trip. 


BB mates 
FEMALES 


MANN-WHITNEY 


P<.009 


WEANLINGS 
(oe) 


7 N 


0 10 20 30 40 50 60 70 80 90 100 110 120 
DISTANCE (km) FROM WEANING LOCATION 


151 


FIGURE 8.11 -- Relative size of weanling male and female sea 
otter home ranges in Prince William Sound, Alaska, during the 
first winter following weaning. Only weanlings with well 
defined home ranges are included. 


MALES 


MANN-WHITNEY 
P > 0.05 
N.S. 


FEMALES 


A 10 20 30 40 50 60 
DISTANCE BETWEEN 
EXTREME LOCATIONS (km) 


152 


FIGURE 8.12 -- Tendency for weanling sea otters in Prince 
William Sound, Alaska, to leave natal female area after being 
weaned. Female weanlings usually do not leave natal female 
area, whereas males usually do. Female area consists of 
zones 1, 2, 3, 4, 6, 7, 8, and 11 on Figure 8.4. 


xX? = 12.7, P<0.01 


sy 
q 12 INSIDE 
=) 10) FEMALE 
OQ g AREA 
= 6 
(a) 
=> 

0 
le OUTSIDE 
<q 12 FEMALE 
=2)) 3100 AREA 
QO ¢s 
= 6 
O 4 
=i 

0 


MALES FEMALES 


IL5)3) 


over the course of a year. These data also show that sexual 
segregation occurs very early in life as a result of 
differences in the behavior of male and female weanlings. 
Weanlings of both sexes were competent and capable of making 
considerable movements as soon as they became independent. 
These observations are discussed below. 


Adult movements 


The observed movement patterns of the adult female sea 
otters in Prince William Sound were remarkable for several 
reasons. Overall, adult females were more mobile than had 
been anticipated. Individual females used a considerable 
portion of the total female area. However, both the extent 
of their movements and their destinations changed seasonally 
and with the age of their dependent offspring. Adult females 
became quite sedentary when accompanied by large dependent 
pups that were close to the age of weaning. They were also 
more gregarious and exhibited more pronounced seasonal habitat 
use patterns than suggested by previous investigations. 


The median distances between extreme locations (DBEL) of 
independent females (41 km) and females with dependent pups 
(33 km) were longer than those measured in previous studies. 
Kenyon (1969) suggested that the home range of females usually 
included less than 17 km of coastline. Garshelis and 
Garshelis (1984) studied independent females at Green Island, 
which is located about 80 km southwest of the current study 
area. They measured DBEL that ranged from 2.6 - 15 km. The 
early studies in California reported female home ranges that 
were comparatively smaller than those observed by Garshelis 
and Garshelis (Loughlin 1977; Ribic 1982). None of those 
studies reported female movements in excess of about 17 kn. 
However, the results of ongoing studies at other locations in 
Alaska and California (Monnett, 1987) (Chapter 3) suggest that 
DBEL as long as those observed in this study may be common. 


Garshelis and Garshelis (1984) recognized that females 
were leaving their study area and that their data did not 
necessarily reflect annual or lifetime home ranges. Ie ahs} 
possible that some of the studies that have reported 
relatively sedentary behavior by females have obtained that 
result as a function of the study design. If individuals are 
monitored for periods of a few months or less, large distance 
movements between functional sub-habitats might not be 
identified (e.g. breeding vs. wintering area). Seasonal 
movements have also been suggested for the population that 
inhabits the Bering Sea along the Alaska Peninsula (Lensink 
1962; Cimberg & Costa 1975; but see Monnett 1987b). 


Based on a limited number of observations at Green 
Island, Garshelis & Garshelis (1984) concluded that females 


154 


occupied larger home ranges when accompanied by young pups 
than they did when their pups were nearly ready for weaning. 
They posited that the change in behavior, as the pups became 
older, was necessitated by the need for pups to gain 
experience in self feeding. Such experience could only be 
gained if movements were restricted to a limited portion of 
the habitat where water depths were shallow and suitable prey 
were available. Their hypothesis is consistent with the 
findings of this study. However, there are no data from this 
study on pup feeding behavior in weaning, or other, areas. 


It is likely that real variation in the movement patterns 
of adult females exists between the populations of sea otters 
at separate study areas due to differences in population 
status, geography or genotypes. 


In the eastern portion of Prince William Sound, range 
reoccupation is still occurring, with subsequent changes in 
distribution. The Green Island area, where Garshelis and 
Garshelis (1984) conducted most of their work on female 
movements, has been occupied much longer and, hence, 
distributions per se are probably not changing because of 
recolonization. 


In areas where coastlines are complex, the extent of 
movements could also be constrained by local geography 
(Garshelis & Garshelis 1984). Relatedly, individuals that 
inhabit calm waters such as Prince William Sound might find 
travel less physically demanding and/or less risky than 
individuals that inhabit more exposed waters such as the Gulf 
of Alaska or the Pacific Ocean off California. For example, 
females with pups might find travel difficult if they inhabit 
waters along unprotected coastlines and, thus, restrict their 
movements to only the most protected areas (D.B. Siniff 
personal communication). 


Authors have not agreed on the relative extent of 
movements of males and females. Kenyon (1969) and Garshelis 
and Garshelis (1984) have argued that males have larger home 
ranges, whereas Loughlin (1977) and Ribic (1982) reported 
larger home ranges for females. The significance of these 
reported differences should be viewed conservatively. It is 
likely that some comparisons reflect differences in movements 
that are contained within single male areas or female areas 
and exclude movements between major habitat units that are 
of functional significance (e.g. the criticism by Garshelis 
and Garshelis (1984) of Ribic's (1982) result). It is well 
documented that individuals of both sexes can make very long 
movements (> 80 km) between breeding areas and wintering areas 
(reviewed in Garshelis & Garshelis 1984; this study). 
Conversely, individuals of both sexes can be relatively 
sedentary (e.g. territorial males (Garshelis & Garshelis 1984) 


155 


vs. females with dependent pups just before weaning (this 
study)). Mobile and sedentary periods may not coincide in 
males and females. As a consequence, measures of home ranges 
are likely to be biased if studies are short term and do not 
take into account sex-specific seasonal differences in habitat 
use patterns. General conclusions about sex differences are 
premature until longer-term data on both sexes are available. 


From observations at Green Island, Garshelis and 
Garshelis (1984) concluded that females formed considerably 
smaller groups than did males. Indeed, only relatively small 
groups of females (< 50) were encountered in the vicinity of 
Green Island during field work conducted during 1985 and 1986 
(personal observation). However, recent observations in other 
parts of Prince William Sound and elsewhere suggest that this 
conclusion may not be general. Females frequently form large 
aggregations, well in excess of 100 individuals. Such 
aggregations have been seen in Prince William Sound (this 
study), in the Bering Sea along the Alaska Peninsula (Monnett 
1987) and in the vicinity of Kodiak Island, Alaska (A. DeGange 
personal communication). As Garshelis and Garshelis (1984) 
point out, group size is likely to be related to the type of 
sea otter activity and local density. 


Weanling Movements 


Observations during this study indicated that weanlings 
established home ranges in their respective male and female 
areas shortly after separation from their mothers. Most 
weanlings moved a long distance from their site of weaning 
before doing so. Since pups were weaned in female areas, 
female weanlings were not required to travel as far as males. 
Garshelis, et al. (1984) observed a similar rapid departure 
of male weanlings from natal female areas. However, they also 
noted that during May-August the age structure of males in 
male areas had an excess of 2 and 3 year old males (by two to 
three fold) compared to yearling males. Because of this age 
structure they suggested that some weanling males may delay 
entering male areas until after their first year. 
Observations on radio-instrumented pups in this study do not 
support that contention. Another way to interpret the age 
structure observation is that they are observing variation 
in cohort recruitment. Two observations support this latter 
alternative. First, there was significant variation in the 
sex ratio of dependent pups caught in the same area between 
1984-1986 (Chapter 9). Yearly differences in male birth 
rates, or survival rates, could result in corresponding cohort 
differences in the male rafts. Moreover, the delayed entry 
hypothesis does not explain the relative lack of 4-5 year old 
males in the male area observed by Garshelis, et al. (1984). 
Four and five year old males are generally assumed to be 
immature and should have been present in the male area in 


156 


their actual proportions. The relative scarcity of 4 and 5 
year old males is consistent with the notion of varying cohort 
sizes. 


Evidence from other investigations also suggests that 
weanlings in Alaska segregate into respective male and female 
areas relatively soon after weaning. Kenyon (1969) reported 
more juvenile females than juvenile males (53:31) in female 
areas near Amchitka, in the western Aleutian Islands. 
Lensink (1962) examined dead juveniles at male hauling grounds 
and found the sex ratios to favor males (30:6). This ratio 
seems too extreme to be due to the easier identification of 
male carcasses in poor condition (see Chapter 6). 


The recently independent otters in this’ study, 
particularly males, apparently were capable of traveling long 
distances in a fairly short period of time. Young male sea 
otters have been observed to make similar long distance 
movements to male areas in previous studies. Garshelis, et 
al., (1984) observed a male weanling that left a female area 
and traveled over 100 km to a male area in Prince William 
Sound. 


Garshelis, et al. (1984) suggested that young male sea 
otters may move into the aggregations found in male areas 
because of certain benefits that may be derived from 
gregariousness. Those authors suggested that benefits may be 
derived from social facilitation, opportunities for assessment 
of conspecific competitors, safety from predation and 
metabolic advantages gained from hauling out on sandbars that 
exist in those areas. Data on survival rates of weanlings 
from this study supports the hypothesis that movement from 
the natal area to a male area may be generally beneficial. 
Male weanlings, that spent their first winter following 
weaning in the male area, were more likely to survive than 
those that remained in the female areas (Monnett 1987). 


The relationship between movement data and dispersal 


For this discussion we follow the terminology of 
Greenwood (1980:1141) who distinguished between several common 
usages of the term dispersal. He defined natal dispersal as 


that dispersal "...from birth site to first breeding or 
potential breeding site...". He contrasted natal dispersal 
with breeding dispersal: "...movement of individuals, which 
have reproduced, between successive breeding sites..." and 


effective dispersal: natal or breeding dispersal followed by 
successful breeding. These definitions are not universally 
accepted. Rather, the terminology of dispersal has been used 
inconsistently between authors (e.g. Greenwood 1980 cf. 
Lidicker 1975; Gaines and McClenaghan 1980; Dobson 1982). 


157 


The movement data given herein should not be taken as 
accurate measurements of natal, effective or breeding 
dispersal as defined by Greenwood (1980). That is, measures 
are incomplete because data are not given on the distribution 
of mature individuals or their breeding sites. These data are 
of post-weaning movement patterns. 


From observations made during this study, it does appear 
that young females are more conservative in their dispersal 


movements than males. Several observations support this 
contention. First, males traveled further than females during 
the weeks immediately following weaning. The very long 


movements were all made by males. Moreover, the sex ratio of 
the individuals with which contact was lost strongly favored 
males. That suggests that some males may have left the 
monitored region. Second, males usually left the female area 
of birth when weaned, whereas females rarely did. Third, 
juvenile males remained outside the female area as long as 
they were monitored. Conversely, those females that left the 
female area following weaning, and that survived the winter, 
returned to the female area in the spring. Fourth, as 
juveniles, most males continued to travel further from their 
weaning location. Females restricted their travels to sites 
within the female areas. 


The observations made during this study do suggest that 
future interpretations of the relationship between breeding 
sites and natal sites in sea otter populations are likely to 
be problematic. In order to evaluate an individual's 
dispersal, a starting point has to be determined, as well as 
an ending point. Howard (1960) referred to this starting point 
as "...its point of origin..." which Greenwood (1980) took to 
mean its place, or group, of birth. There is little 
confusion when the young remain at a single location 
throughout dependency (e.g. nests, dens or burrows) or when 
they remain within a cohesive social group (e.g. prides, packs 


or troops). However, as shown in this study, sea otter 
females travel extensively while accompanied by dependent 
pups. Consequently, the sites of conception, birth and 
weaning may not coincide. Any, or all, of those locations 


could be regarded as the point of origin, depending upon the 
question under consideration (e.g. inbreeding avoidance vs. 
site experience). 


Available evidence seems to indicate that females stay 
within their natal female areas, near the areas they inhabited 
when they accompanied their mothers. However, many leave the 
immediate location where they were weaned. Conversely, 
juvenile male departure from the same areas appears to be 
almost obligate. Only two males out of the 12 that survived 
their first winter had not left the female area by January. 
In order to exhibit natal philopatry equal to that exhibited 


158 


by female weanlings, males would have to re-enter natal female 
areas after becoming sexually mature and breed in areas they 
inhabited with their mothers. However, from the perspective 
of the juvenile male, any attraction to the natal female area 
that might result as a consequence of benefits derived from 
site familiarity or individual recognition (Greenwood 1980) 
are likely to have been diminished by the length of tenure 
in the male area. Male sea otters mature and apparently begin 
to seek breeding opportunities after five or more years of 
residence in the male area. It would seem questionable 
whether what was learned about the natal female area during 
dependency could be retained until the age of potential 
re-entry. Moreover, even if retained, some of that knowledge 
would be likely to be obsolete because of changes in habitat 
conditions during that period (e.g. food abundance, 
conspecific distributions). 


LITERATURE CITED 


Cimberg, R.L. and Costa, D.P. (1985). North Aleutian Shelf 
sea otters and their vulnerability to oil. Proceedings: 
Oil spill conference, Los Angeles, CA. 


Calkins, D.G., and Lent., P.G. (1975). Territoriality and 
mating behavior in Prince William Sound sea otters. 


Journal of Mammalogy, 26, 528-529. 


Dobson, F.S. (1982). Competition for mates and predominant 
juvenile male dispersal in mammals. Animal Behaviour, 
30, 1183-1192. 


Garshelis, D.L. (1983). Ecology of sea otters in Prince 
William Sound, Alaska. Unpublished Ph.D. thesis, 
University of MInnesota, Minneapolis, MN. 330 pp. 


Garshelis, D.L., and Siniff, D.B. (1983). Evaluation of 
radio-transmitter attachments for sea otters. Wildlife 
Society Bulletin, 11,378-383. 


Garshelis, D.L., and Garshelis, J.A. (1984). Movements and 
management of sea otters in Alaska. Journal of Wildlife 
Management, 48,665-678. 


Garshelis, D.L., Johnson, A.M., and Garshelis, J.A. (1984). 
Social organization of sea otters in Prince William 


Sound, Alaska. Canadian Journal of Zoology, 
62,2648-2658. 


Garshelis, D.L., Garshelis, J.A., and Kimker, A.T. (1986). 


Sea otter time budgets and prey relationships in Alaska. 
Journal of Wildlife Management, 50,637-647. 


159 


Gilmer, D.S., Cowardin, L.M., Duval, R.L., Mechlin, L.M., 
Schaiffer, C.W., and Kuechle, V.B. (1981). Procedures 
for the use of aircraft in wildlife biotelemetry studies. 
United States Department of the Interior, Fish and 
Wildlife Service Research Publication 140. Washington, 
D.C. 19 pp. 


Green, B.D. (1978). Sexual maturity and senescence of the 
male California sea otter (Enhydra lutris). Unpublished 
M.A. Thesis, San Jose State University, San Jose, CA. 


Greenwood, P.J. (1980). Mating systems, philopatry and 
dispersal in birds and mammals. Animal Behaviour, 
28,1140-1162. 


Greenwood, P.J., and Harvey, P.H. (1982). The natal and 
breeding dispersal of birds. Annual Review of Ecology 
and Systematics, 13,1-21. 


Howard, W.E. (1960). Innate and environmental dispersal of 
individual vertebrates. American Midland Naturalist, 
63,152-163. 


Kenyon, K.W. (1969). The sea otter in the eastern Pacific 
Ocean. United States Fish and Wildlife Service. North 
American Fauna, No. 68, 352 pp. 


Lensink, C.J. (1962). The history and status of sea otters 
in Alaska. Unpublished Ph.D. Thesis, Purdue University, 
Lafayette, IN 186 pp. 


Lidicker, W.Z. Jr. (1975). The role of dispersal in the 
demography of small mammals. In: Small Mammals: 
Productivity and Dynamics of Populations. (Ed. by K. 
Petrusewicz, E.B. Golley, and L. Ryszkowski), pp. 
103-128. Cambridge University Press, London. 


Loughlin, T.R. (1977). Activity patterns, habitat 
partitioning, and grooming behavior of the sea _ otter, 
(Enhydra _lutris), in California. Unpublished Ph.D. 
thesis, University of California, Los Angeles, CA. 110 
Pp. 


Monnett, Charles W. 1987. Movement, development and 
mortality patterns of sea otters in Alaska. Ph.D. 
thesis, University of Minnesota, Minneapolis, MN. 141 
PPp- 


Odum, E.P., and Kuenzler, E.J. (1965). Measurements of 
territory and home range size in birds. Auk, 72,128-137. 


160 


Peterson, R.S., and Odemar, M.W. (1969). Population growth 
of the sea otter in California: results of aerial census 
and behavioral studies. Proceedings Annual Conference 
on Sonar Diving Mammals, 6,69-72. 


Ribic, C.A. (1982). Autumn movement and home range of sea 
otters in California. Journal of Wildlife Management, 
45,795-801. 

Riedman, M.L. (1986). Draft environmental impact statement 


of proposed translocation of southern sea otters. Volume 
II: Technical support documents, United States Fish and 
Wildlife Service and University of California, Santa 
Cruz, CA. 


Schneider, K.B. (1978). Sex and age segregation of sea 
otters. Alaska Department of Fish and Game, Final 
Report, Federal Aid Wildlife Restoration Projects W-17-4 
to W-17-8. 


Siniff, D.B., Williams, T.D., Johnson, A.M., and Garshelis, 
D.L. (1982). Experiments on the response of sea otters 
Enhydra lutris to oil contamination. Biological 
Conservation, 2,261-272. 


Vandevere, J.E. (1970). Reproduction in the southern sea 


otter. Proceedings of the Annual Conference on the 
Biology of Sonar Diving Mammals, 7,221-227. 


Williams, T.D., Williams, A.L., and Siniff, D.B. (1981). 
Fentanyl and azaperone produced neuroleptananalgesia in 
the sea otter (Enhydra _ lutris). Journal of Wildlife 


Diseases, 17,337-342. 


161 


CHAPTER 9 


SEX-RELATED PATTERNS IN THE POST-NATAL DEVELOPMENT AND 


SURVIVAL OF SEA OTTERS IN PRINCE WILLIAM SOUND, ALASKA 


C. MONNETT AND L. ROTTERMAN 


NOVEMBER 30, 1988 


INTRODUCTION 


Pronounced sex differences in morphology, development 
(Glucksman, 1974, 1978), and behavior (Clutton-Brock and 
Albon, 1982) exist between the males and females of many large 
mammals. While the general theoretical explanations for these 
differences are fairly well developed, there is still little 
known about the proximate causes or consequences of these 
differences for most species (Clutton-Brock and Albon, 1982). 


According to current evolutionary theory, parental 
investment (PI) (Trivers, 1972) should be distributed among 
progeny so as to maximize parental inclusive fitness (sensu 
Hamilton, 1964). Disproportionate allocation of PI to male 
or female offspring could occur in situations where the 
production of one sex has a potentially greater effect on 
parental inclusive fitness than the production of the other 
sex. In polygynous mammals, if an inequality is to occur, 
males should be the favored sex for two reasons: intrasexual 
competition and male-biased dispersal. 


Among polygynous mammals, variation in reproductive 
success is likely to be greater among males than among females 
(Trivers, 1972; Clutton-Brock, et al., 1977; Clutton-Brock and 
Albon, 1982). The allocation of parental resources should 
favorably influence an offspring's body condition as an adult 
and thus, its relative ability in intrasexual competition with 
conspecifics. As a result, if their reproductive success is 
affected by their competitive ability, individual offspring 
that receive the most PI during dependency should achieve the 
highest reproductive success as adults (Trivers and Willard, 
1973). Since male mammals generally compete, either directly 
or indirectly, for access to females, high quality sons are 
likely to leave more offspring than high quality daughters 
(Trivers and Willard, 1973). Consequently, females should 
tend to invest more resources in individual sons than in 
individual daughters (Clutton-Brock and Albon, 1982). 


The pattern of PI in individual male and female offspring 
could also be influenced by the tendency for one sex to 
disperse more than the other sex (e.g. Greenwood, 1980). If 
survival during dispersal is influenced by maternal PI before 
weaning, females should contribute more PI to progeny of the 
dispersing sex (Clutton-Brock and Albon, 1982). In most 
mammals those progeny would be males (reviewed in Greenwood 
1980). 


Given the typical mating system and dispersal patterns 
seen in mammals, a unit of PI apparently has a greater 
potential to effect the survival and the reproductive success 
of male progeny than female progeny. Thus, more PI should 
be allocated to dependent males relative to females. This 


163 


could be accomplished in 2 general ways: First, parents could 
manipulate offspring sex at conception (Trivers and Willard, 
1973). Second, during the period of dependency, parents could 
preferentially allocate more PI to male offspring than to 
female offspring (Reiter et al., 1978). For example, males 
could be given high quality and/or greater quantities of food 
or permitted to have longer dependency periods. 


Sea otters (Enhydra lutris) are a sexually dimorphic, 
polygynous mustelid that is highly specialized for the marine 
environment (Kenyon, 1969). They have a resource-defense 
mating system (Greenwood, 1980) with males occupying and 
defending breeding territories (e.g., Calkins and Lent, 1975; 
Loughlin, 1977; Garshelis and Garshelis, 1984). In Alaska, 
males tend to leave the area in which they were reared at the 
end of parental care, whereas females do not (Chapter 8). A 
radio-telemetry study was conducted from 1984-1987 in order 
to investigate differences in juvenile development and 
behavior, with specific focus on differences between the 
sexes. Comparisons are made of time of birth, growth rates, 
dependency periods and mortality patterns. Under the 
theoretical arguments reviewed above, male sea otters should 
be born earlier in the spring, grow more rapidly and have 
longer dependency periods. Differences in post-weaning 
behavior between sexes, especially in movement patterns, might 
subject juvenile males and females to different risks. Since 
male sea otters travel farther from their weaning site, they 
should be subject to greater risk and exhibit correspondingly 
higher mortality during the first few months following 
weaning. 


STUDY AREA AND METHODS 


Studies of sea otter pups and weanlings were carried out 
in two general areas within Prince William Sound (PWS) in 
south-central Alaska (Figure 9.1). Observations were made in 
the northeastern portion of the Sound during 1984-1986. 
Observations were made at Green island, in the south-central 
sound, during the summer of 1985. Research activities were 
coordinated from cabins in Sheep Bay and on Green Island, 
within the Sound, and from a United States Fish and Wildlife 
Service warehouse and University of Alaska marine advisory 
office in Cordova. The local population of sea otters, its 
history, habitat and ecology, have been described by a number 
of authors (Calkins and Lent, 1975; Siniff, et al., 1982; 
Garshelis and Siniff, 1983; Garshelis and Garshelis, 1984; 
Garshelis, et al., 1984; Garshelis, et al., 1986). 


164 


William Sound, 1984-1987. 


\ 
PRINCE WILLIAM 
SOUND 


[ill] stuby AREA 


: ‘i 


is i 


Pup capture 


Pups (N=157) were captured for routine tagging and data 
collection with dip nets and tangle nets. Several types of 
nets, including commercially available salmon dip nets, were 
tested. The most satisfactory results, and the most pups, 
were obtained by using a custom fabricated, aluminum dip net 
(Alaska Power Services, Cordova, Ak 99574). This net was 
characterized by a long (4m) handle and semicircular "basket" 
that attached to the handle at a 90 degree angle. Mother-pup 
pairs were pursued in a Boston Whaler skiff (5.5 M) until the 
pair surfaced near the bow. The basket of the net was then 
dropped in front of the moving animal(s) and drawn, with the 
pup, up and back over the bow of the boat, the pup captured 
in the net. Large pups were usually not carried by their 
mothers and, since they were generally incompetent divers, 
were easily netted. Smaller pups were carried by their 
mothers as they made repeated dives. Successful captures 
usually occurred in 3 circumstances: 1. The mother was caught 
as she carried the pup, separated from the pup and released. 
2. The pup was separated from the mother by drawing the net 
between the pair, over the pup, as they surfaced for air. 3. 
The mother released the pup after a few dives and it was 
scooped as it floundered on the surface. Exceptions to the 
above scenarios included a few mothers (less than 10%) that 
abandoned their pups on the surface immediately on the 
approach of the boat and a few cases when mothers were on 
foraging dives when their pups were captured. 


Pup handling 


Upon capture, each pup was tagged with numbered, nylon 
Temple or button tags (Ames, et al., 1983) in the interdigital 
webbing of one, or both, hind flipper(s). The weight, length 
and sex were recorded. Weights were taken in pounds because 
equipment was available only with those _ scales. The 
unevenness of some of the values assumed for calculations 
reflects the conversion of those measures to metric scale. 
Pups were weighed to the nearest 0.5 pound, with the exception 
of the newborns that were weighed to the nearest 0.25 pound. 
Pups were held for 5-15 minutes depending upon whether a 
blood sample was taken. 


Pup release 


Pups were released into the water and generally observed 
until they reunited with their mothers. If the mother was 
not near the boat when a pup was to be released, cassette 
playbacks of pup vocalizations (loud cries) were used as an 
attractant. Mothers appeared not to discriminate between the 
sounds of their own pup and those of others. The recorded 
cries of a single pup were used effectively on many different 


166 


mother-pup pairs. Females were attracted to the boat with 
these playbacks from distances of over 1/2 kn. Pups also 
responded to the playbacks by swimming toward the boat and, 
on some occasions, crying back. They were particularly 
indiscriminate and often could be stimulated into lengthy 
conversations by crude, human imitations of their own cry. 


Recaptures 


Forty-one dependent pups were recaptured on at least one 
occasion. Previously caught pups were selected for capture 
and identified by their flipper tags. They were captured and 
handled as described above. Capture activities were usually 
spaced to insure that intervals of 30-70 days had passed 
between successive captures so that growth rates could be 
determined. 


Telemetry 


Radio-transmitters were surgically implanted in the 
peritoneal cavity of 37 dependent pups, during August or 
September, 1984 or 1985, by veterinarians, as described in 
Chapter 1. Most pups weighed 9-14 kg when implanted (range 
7-20 kg). 


Upon capture, pups were brought aboard a 5.5 m skiff, 
weighed and immobilized with a combination of fentanyl 
(generally 0.05 mg/kg) and azaperone (0.20 mg/kg) (Williams, 
et al., 1981). Naloxone (0.01 mg/kg), an antagonist to 
fentanyl, was injected in all subjects following surgery, but 
before release. Pups were released into the water near their 
capture site and generally were observed until they reunited 
with their mothers. Normally less than 60 minutes elapsed 
between an animal's capture and release. Playbacks of pup 
cries, as described above, were used to keep mothers attentive 
and near the boat during pups' surgeries. 


Radio-implanted otters were monitored from small aircraft 
or small boats equipped with yagi antennas and 2000 channel, 
programmable scanning receivers (Cedar Creek Bioelectronics 
Lab). Radios had ranges of 1-5 km and 6-10 km when monitored 
from boats and aircraft, respectively. An attempt was made 
to observe most pups 1-2 times per week during the fall, 
before they became independent. 


The transmitters had a maximum life expectancy of about 
700 days. One-hundred twenty-five transmitters of the same 
design were implanted in sea otters in Alaska and California, 
between March, 1984, and September, 1986. To date, only two 
are believed to have malfunctioned and many have operated 500 
days or longer. The durability of the units is evidenced by 
the recovery of 14 operating transmitters from intertidal 


167 


marine areas at various times following the deaths of the 
subjects. Some were found buried under boulders on beaches 
and had been subjected to heavy surf. One radio was still 
operating after at least 13 months on a gravel beach, in the 
intertidal, near Valdez. 


Birth date estimates 


Sea otter births are seldom, if ever, observed in 
natural situations. Consequently, birth dates for pups must 
be estimated. Such estimates can be based upon pup weight at 
capture, if information is available about normal birth 
weights and about pup growth rates (Wendell, et al., 1984). 
If relevant data are not available, assumptions must be made 
about birth weights and growth rates, based on population 
averages. 


For estimates given here, birth weight was assumed to be 
approximately 4.5 lb (2.04 kg). This value was chosen based 
on several types of information available from sea otters in 
Alaska. First, 2 kg approximates the average birth weight 
observed in this study. Second, based on his observations of 
fetuses and newborns in a population near Amchitka, Kenyon 
(1969) argued that "normal" birth weight is between 1.87 and 
Doss Ikeja | Uslolalagols Schneider (1978b), used the same types of 
data to estimate that the mean birth weight in the central 
and western Aleutian Islands was 1.8 - 1.9 kg. Te S's 
important to note that individual variation in birth weight 
has little effect on the accuracy of birth date estimates 
since growth rates are fast. An error of several hundred 
grams would only change the estimated birth date by a few 
days. 


Based on data from this study, growth rates are assumed 
to be approximately 95 g/day (0.21 1b) for males and 86 g/day 
(0.19 lb) for females. No other growth rate data are 
available for sea otters in Alaska. 


Dependency period estimates 


Dependency periods for the radio-implanted pups were 
calculated from estimated birth dates and separation dates. 
In order to do this, several assumptions were made. As noted 
above, birth weight was assumed to be 2.04 kg. Actual growth 
rates were used for those pups that had been recaptured and 
hence, had been weighed on two or more occasions. If multiple 
weights were not available, it was assumed that females gained 
86 g/day and that males gained 95 g/day. In these cases, 
estimated birth dates were bracketed by estimates made by 
assuming growth rates of plus and minus 1 standard deviation 
(SD). 


168 


Survival rates 


Survival estimates, based on telemetry data, were 
calculated using the method developed by Trent and Rongstad 
(1974). 


Separate survival estimates were calculated for males 
and females during pre-weaning and post-weaning intervals. 
Since the exact day of death was rarely known, two 
survival-related calculations were made. inp tehe) | rales te 
calculation, it was assumed that the animal died the last day 
it was seen alive. In the second calculation, it was assumed 
that it died the first day it was known to be dead. In some 
instances individuals became missing but it was not certain 
whether the cause was death, dispersal or radio failure. 
Thus, the suggestion of Heisley and Fuller (1985) was followed 
and two survival rates were calculated. In the first, it is 
assumed that missing individuals were dead; in the second, 
that they were alive. 


One pup died shortly following surgery as a result of a 
veterinary error. This pup is not included in estimates about 
survival since it is unlikely that the error will be repeated, 
and thus the case is not relevant to understanding normal 
survival probabilities or factors influencing survival 
schedules. 


RESULTS 


Birth weight 


It was assumed that pups were newborn if pink umbilical 
fragments were still attached when they were captured (Kenyon 
1969). Three such newborn pups had weights and total body 
lengths of 1.7 kg and 48 cm, 1.8 kg and 50 cm and 2.4 kg and 
55 cm, respectively. Three other small pups were captured 
that had no trace of umbilical fragments. These measured 
1.9 kg and 48 cm, 2.5 kg and 57 cm and 2.7 kg and 55 cm. 


Growth rates 


Twenty-nine pups were recaptured after intervals of 34 


days or longer (mean interval = 65.7 days, SD = 11.9, range 
(34-98). Male growth rates (mean = 95 g/day, SD = 15 g, 
range = 67 - 123, N = 18) were faster than female growth rates 
(mean = 83 g/day, SD = 10 g, range = 63 - 88, N= 11) (Fig. 


9.2; Mann-Whitney U-test: U = 149,49; N = 18,11; P .03). 


Fourteen pups were recaptured after intervals of 11-28 


days. As would be expected, rates were much more variable 
than those observed for the longer intervals: males mean = 
79 g/day, SD = 33 g, range = 41 - 132 g, N = 6; females mean 


169 


= 72 g/day, SD = 34 g, range = 39 - 132 g, N= 8. A _ female 
and a male pup both achieved the maximum growth rate observed 
in this study by gaining 3.2 kg in 24 days. 


Small pups appeared to grow at approximately the same 
rates as large pups (Table 9.1). The five smallest pups 
gained on average 92 g/day, whereas, the 6 largest pups 
averaged 93 g/day. 


TABLE 9.1 - Comparison of growth rates for large vs. small 
pups. 
SEX WT. (kg) WT. (kg) INTERVAL GROWTH 
ist CAPTURE 2nd _ CAPTURE DAYS RATE d 
M 1.8 7.7 62 95 
SMALL F 2.4 8.4 68 88 
PUPS F 3.2 8.4 67 78 
M 3.6 Atal dt 65 115 
1E 3.6 9.8 ii, 86 
MEANS 2.9 9.1 67 92 
M BG) ta 58) 63 86 
M 5.9 10.9 51 98 
LARGE F 5.9 11.8 68 87 
PUPS M 6.8 12.0 51 102 
M 6.8 14.5 74 104 
E 6.8 14.1 88 83 
MEANS 6.4 12.4 66 93 


Timing of parturition 


The modal estimated birth date for Prince William Sound 
pups was between May 20 and May 29 (Fig. 9.3). Male and 
female pups were born in nearly constant proportions 


throughout the spring. The rapid increase in births after 
April 30, and decrease after June 28, is a conspicuous 
characteristic of the distribution. It is widely accepted 


that the timing of the seasonal peak in birth rates varies 
seasonally throughout the sea otter's range, but that pups 
are born in all seasons. The spring abundance of young sea 
otters is readily observable in the Prince William Sound, and 
elsewhere. However, since pups were caught only between early 
June and late September, it could be argued that the apparent 
rapid changes in birth rate were an artifact of the sampling 
scheme. A perceived late April increase could result if large 
pups, those born in April or earlier, were present in June, 
but were too large to be captured in a dip net. In) fact; 
capture success does tend to be lower for pups that are larger 
than about 10 kg. However, relatively large pups were 


170 


captured. Seventy-three pups weighing at least 9.1 kg (20 lb) 
and 10 pups of at least 13.6 kg (30 1b) were dip-netted. If 
pups were commonly born in early April, they should have been 
obvious and easily captured since they would have weighed only 
8-10 kg by the end of June. However, at that time of year 
pups of that size were seldom observed. The pups dip-netted 
in June (N = 44) weighed on average 4.9 kg (10.9 1b) with the 
largest being only 8.6 kg (19 lb). 


Few pups were born during the mid and late summer. Given 
the intensity of capture effort after August 28 (N = 50), 
quite a few small pups should have been captured if they were 
in the study area in the late summer and early fall. Data on 
22 pups, dip-netted between 21 and 30 September, 1986, support 
the contention that pups were rarely born after June in the 
Prince William Sound. The average pup captured during this 
interval weighed 10.4 kg (22.8 1b). The smallest pup captured 
weighed 5.2 kg, a value heavier than the average weight of 
June pups. 


Incorrect assumptions about growth rates could cause 
errors in calculations. Moreover, such errors would be 
greatest for individuals that were not caught until they were 
fairly large. In order to illustrate the potential magnitude 
of such errors, we calculated the difference between the 
estimated birth dates of individuals under two assumptions 
about growth rates. That is, we made two separate 
calculations: the first using a growth rate of + 1 standard 
deviation (SD) and the second using - 1 SD (Table 9.2). 
Since it was noted that pups caught before August 1 tended to 
be smaller than those caught afterward, data were displayed 
accordingly. For example, male pups born before August 1 
weighed, on average, 5.7 kg. If a growth rate of 82 kg/day 
(-1 SD) was assumed, and the real growth rate was 95 g/day, 
the estimated age of one sixth of the male pups, when 
captured, would have been at least 6.6 days greater than it 
really was. Likewise, if the growth rate was assumed to be 
109 g/day (+1 SD) the estimated age of one sixth of the males 
would have been at least 4.8 days less than their real age. 
Calculations on pups caught after August 1 tended to have a 
greater error potential, since the pups were larger. 
Conversely, calculations on females had less error potential 
since females were smaller and had slower growth rates. 


Progeny sex ratio 


Sex was determined for 156 dependent pups (Table 9.3). 
Total sex ratio favored males, but not significantly so. 
Substantial differences in sample sizes and proportions 
existed between years so a yearly average was also calculated. 
This average favored females, slightly. Additional data on 


171 


TABLE 9.2 - Error in estimation of birth dates from growth 
rate assumptions. 


ASSUMED DEVIATION FROM 

CAPT. AVE WT GROWTH RATE EST. AGE 

DATE (kg) -SD AVE _+SD -SD +SD 

EARLY 5.7 +6.6 -4.8 
MALES 82 95 109 

LATE 9.6 +13.2 -9.8 

EARLY 5.6 +4.8 -4.5 
FEMALES 77 86 95 

LATE 9.0 +9.5 =7 od 


22 dependent pups near False Pass, on the Alaska Peninsula, 
included 14 males and 10 females. Thus, the observed total 
sex ratio, for dependent pups from 2 populations was 95:85. 


Dependency period 


We assumed that pups were weaned at the time they become 
separated from the mothers. This follows from _ the 
observations of Schneider (1978b) who found that "...females 
with even the largest pup were found to be lactating." and 
Payne and Jameson (1984). The peak time of maternal 
separation for 21 pups that were instrumented during 1985 was 
October 16 - November 15 (Fig. 9.4). The study otters were 
not monitored between December 15, 1985 and February 8, 1986. 
The fact that three pups were weaned during this interval is 
reflected on Fig. 9.4. 


The chronology of the dependency periods of radio- 
implanted pups is given (Fig. 9.5). Twenty-seven pups are 
represented; 6 from 1984 and 21 from 1985. Other pups were 
monitored during 1984 but monitoring was inadequate during the 
fall and winter to determine relatively accurate weaning 
dates. Estimated birth dates for individuals with known 
growth rates are displayed as unbracketed open circles. 


TABLE 9.3 - Sex of dependent Prince William Sound sea otter 
pups. 


1984 1985 1986 
MALE FEMALE MALE _ FEMALE MALE FEMALE 
EASTERN PWS 34 36 19 10 08 14 
GREEN ISLAND BO oS 20 15 = =F 
TOTAL 34 36 39 25 08 14 
GRAND TOTAL 81 75 
TOTAL PROPORTIONS -519 ~481 


AVERAGE PROPORTIONS 2486 -514 


172 


FIGURE 9.2 -- 
otter pups 


Growth rate of depende rarer and female sea 


ased on two we ee ngs at le 30 days apart. 


Males grew faster than female 


NUMBER OF INDIVIDUALS 
= nm ow p> wo ro) 


\Y FEMALES 


a MALES 


MANN-WHITNEY 


P<.03 


GON NGS) 7 O75) Ol msSi 9025) 100) 105) 1110) 11115) 1120 
GROWTH RATE (GRAMS/DAY) 


FIGURE 9.3 -- Estimated birth dates and capture dates of sea 
otter pups in Prince William Sound, Alaska. No tendency was 
found for one sex to be born- earlier than the other. 


EM Mace BintH OATES 

Y 
KW FEMALE BIRTH DATES 
CAPTURE DATES 


n°) _W 


NUMBER OF PUPS 


avid 40 400 5/10 §20 590 69 69 629 79 79 7129 ae sie ame 97 917 9/27 
FIRST DAY OF 10 DAY INTERVAL 


174 


FIGURE 9.4 -- Weaning dates of instrumented sea otter pups in 
Prince William Sound, Alaska, 1985-1986. Pups were 
considered weaned when they separated from their mothers. No 


tendency existed for either sex to be weaned earlier than the 
other. 


FEMALES 


NUMBER OF PUPS 


ae a a be 
So lo) ie) SR) Om 

rr - - wu 
Me the Sr Om SO) 


WEANING DATES 


L75) 


RADIO-IMPLANTED PUPS 


FIGURE 9.5 -- Chronology of dependency periods of 27 sea 
otter pups in Prince William Sound, Alaska, 1984-1986. 
Horizontal bars represent individual pups. The main events 
during dependency are signified by symbols. 


DS eee ee ee 
(ot ma ©) 
eee 
0 SS 5 0S 


To) LEGEND 


a @aoms BIRTHDATE INTERVAL 
sr Fe MIN OE 
WEANING INTERVAL 
——— = . ——D WEANING DATE 
——————— ei ——* WEANED BY INTERFERENCE 
—=— CAPTURE DATES 


Fiche An ROM iit didi iaiiii Aino S. ah OnseiNae | Dined nb iae Mirae Maidan C3 
MONTHS OF THE YEAR 


176 


When weaning dates were known within plus or minus 4 days they 
are also displayed as open circles. When dates were less 
certain, they are displayed as intervals; the intervals being 
the time between the last sighting before separation and the 
first sighting afterwards. Four pups were weaned as a direct 
result of research activities. They are not included in the 
summary statistics but are considered separately. 


The average duration of dependency was 169 days (5.6 
months). The range was 76 - 333 days (N = 23). Males did 
not differ from females (12 males: 170.1 days, SD = 62, 11 
females: 167.6 days, SD = 45.6). 


Size at weaning was estimated, assuming that growth rates 
remained constant until weaning. Two individuals that were 
weaned after 276 and 333 days, respectively, were excluded 
from this analysis because growth rates could not remain 
constant for that duration. Males were weaned at a slightly 
larger size than females (16.2 kg vs. 15.2 kg.). These 
estimated weaning weights seem reasonable and are consistent 
with field observations, since large dependent pups are not 
uncommon. Ten dependent pups were caught in September, or 
earlier, that weighed between 13.6 kg and 20.0 kg. 


The histories of several individuals are remarkable and 
illustrate the ability of young sea otters to survive even if 
weaned at very young ages. These are presented as case 
histories: 


Case 1. A male pup was weaned, apparently under normal 
circumstances, at about 76 days of age and weighing 
approximately 12 kg. It survived until it drowned in a gill 
net 9 months later. Both this pup and his mother were 
instrumented. Thus, the individuals' histories were fairly 
well known: The mother weaned her previous pup between May 
3rd and June 3rd, 1985. She was alone when observed on June 
3rd. Although her location was determined several times by 
telemetry, she was not observed again until August 18, when 
she was observed accompanied by this male pup. The pup was 
caught the same day and weighed 5.7 kg. On September 11, 24 
days later, it weighed 8.8 kg and was implanted with a 
radio-transmitter. It was weaned between October 1 and 
October 10. Based on a measured growth rate of 132 g/day, 
its birth date was estimated to be 76 days prior to weaning. 
The maximum possible dependency period, assuming that the pup 
was born the last day the female was seen unaccompanied and 
that the pup was weaned the first day it was seen alone, would 
be 129 days. However, assuming a birth weight of 2.04 kg, 
this would require a growth rate of 48 g/day. The slowest 
growth rate observed for a male pup was 67 g/day. If the pup 
was actually weaned on the first day it was seen alone, the 
pup would have to had traveled about 50 km on the day it was 


177 


weaned, which was possible but unlikely. Thus, this male pup 
was probably weaned at considerably less than 129 days and 
possibly at 76 days of age. 


Case 2. A female pup was weaned on August 14 when its 
mother left the vicinity with a male while the pup was 
undergoing surgery to implant a radio-transmitter. When 
weaned, it was estimated to be 76 days old. It weighed 9.5 
kg. This pup had been previously captured on June 29, when 
it weighed 5 kg. Thus, based on a measured growth rate of 95 
g/day, it was assumed to have been born on May 30. When it 
was released following the surgery, the pup swam several km 
across a bay and took up residence in a small, protected cove. 
It was known to have remained within a km of that site until 
its radio expired, approximately 20 months later. 


Cases 3 and 4. Two pups were weaned during the early 
fall at under 100 days of age as a result of separation from 
their mothers during surgery. They weighed 10.4 kg and 12.0 
kg when weaned. One, the 10.4 kg animal, died over 4 months 
later of unknown cause. The other was apparently killed as 
an incidental take in the local salmon fishery the following 
June. 


Mortality 


Of the 36 pups that were released with radio-implants, 
15 are known to have died and contact was lost with 8 during 
the life of the study (Fig. 9.6). One of the deaths occurred 
during dependency and 14 occurred following weaning. 


Deaths were most common during the January-April 
interval, but did occur during all seasons of the year. It 
was difficult to ascertain the cause of most of the deaths, 
since beach-cast carcasses were rapidly consumed by bald 
eagles and other scavengers. However, starvation and 
predation were both observed and may have been significant 
during the fall and winter months. 


It was apparent that one female starved to death. Her 
carcass was recovered very shortly after death. At that time 
she weighed 26% less than she had when implanted, 37 days 
earlier. 


A recently weaned female was killed by coyotes (Canis 
latrans) in Olsen Bay, at the back of a long tidal flat. It 
is possible that she became trapped by the falling tide. Her 
radio was found in newly fallen snow with fresh blood and 
coyote urine markings. The blood was bright red indicating 
that it was well oxygenated and thus, that she was killed, 
rather than scavenged. Three other weanlings were probably 
killed by coyotes in a nearby tidal basin, locally referred 


178 


FIGURE 9.6 -- Sea otter pups instrumented in Prince William 
Sound, Alaska, that died or with which radio contact was 
lost. Contact was more frequently lost with males than 
females. Females tended to die during the January-April 
interval whereas male deaths were distributed year around. 


MISSING PUPS 
Le) 


; \\ 


SEPT-DEC JAN-APR MAY-AUG 


MALES 


FEMALES 


DEAD PUPS 


SEPT-DEC JAN-APR MAY-AUG 


179 


to as Hell's Hole. All died within a few hundred meters of 
each other. The first of these weanlings to die had been seen 
hauling-out on a bank that was heavily used by coyotes, 
judging from the numerous fresh tracks in the snow cover. 
This weanling was dead within a week, leaving no traces except 
the radio on the beach within 100 m of the haul out site. The 
second weanling's radio was recovered on a nearby beach 2 
months later. On the same day, the third weanling was 
observed alive a few hundred meters away. It had become 
stranded on an exposed mussel bed, by the out going tide, 
several hundred meters from deep water. During the next 
survey, 3 months later, its radio was recovered at the same 
location as the second weanling's. The radios of several 
other weanlings were found on large tidal flats. It is 
conceivable that those individuals became stranded and were 
preyed upon. Other radios were recovered from rocky beaches 
in areas subject to wave action. Otters were not observed to 
haul out at these locations. Thus, it is probable that they 
died elsewhere, remained intact, and floated to those sites. 


Entanglement in commercial fishing gear was a significant 
problem and a cause of deaths to the sea otters in this study. 
Commercial fishing is permitted in the general area from May 
until October. All four of the males that were dead or 
missing during the May-June interval may have been killed in 
the fishery. One male's flipper tags were turned in by a 
fisherman who stated that it had drowned in his gill net. 
Another weanling was missing a few days after the annual 
fishing opener, in May. For the preceding 5 months it had 
occupied an area that was heavily fished during the opener. 
Presumably, it died and its carcass drifted beyond the study 
search area. A third male's radio and remains were recovered 
a few km down current from the same area during the middle of 
the fishing season. A fourth male became missing near Valdez 
following an unusually heavy period of gillnetting in the area 
where he was last seen. 


Three other study otters, one an adult female, became 
entangled in fishing nets but were released unharmed. Three 
fishermen reported that otters had snagged their button tags 
on the strands of the net. In one case it was thought that 
the weanling drowned because it snagged near the bottom of 
the net and could not surface. It is not possible to say what 
the role of the tags was in the deaths of these animals. 
Button tags are no longer used at this study site. It would 
seem prudent not to use button tags at any location where 
there is a chance that otters may encounter commercial fishing 
gear. This includes gill nets, seines, trawls or crab pots. 


Radio contact was lost with male weanlings more 
frequently than it was lost with female weanlings (7:1). It 
is possible that radio contact was lost because those 


180 


individuals left the study area. In this study male sea otter 
weanlings did travel more extensively than their female 
counterparts following weaning (Chapter 8). Several of the 
missing males were last seen near the edge of the study area 
after traveling as far as 123 kn. 


Survival and juvenile movements.--The difference in the 
fates of individuals was striking depending upon whether they 
were travelers or not. Approximately 90% (28 of 31) of the 
known or suspected weaning locations were located in three 
Bays: Sheep Bay, Simpson Bay or Port Gravina. Of the 12 
weanlings that remained in those bays: 2 survived, 1 was 
missing and 9 died. Of the 16 weanlings that left those 
areas: 9 survived, 4 were missing and 3 died. Even if it is 
assumed that the missing individuals died, which biases the 
data against a difference in outcome, survival was lower for 
weanlings that remained in the weaning areas (Chi square = 
3.93, 1 DE Peas 05). If the same comparison is made, 
considering only survival until spring, the relationship is 
stronger (Chi square = 8.18, 1 DF, p < .05). If it is assumed 
that the missing weanlings did not die, the relationship is 
stronger since 4 of 5 missing weanlings were travelers. The 
argument is also strengthened by the fact that the 3 weanlings 
that were known to have been weaned in bays other than Sheep, 
Simpson or Port Gravina all survived. Individuals that stayed 
in the latter bays appeared to be dying mostly of predation 
or starvation during the fall or winter. However, those that 
left the weaning areas died mostly in the spring and summer 
due to human-related activities. 


Survival rates 


Two sets of survival-related calculations are given 
(Table 9.4). Sea otters that died of all causes are included 
in one set. In the second set, deaths caused by human 
activities are treated as transmitter failures. That is, the 
otter days are included but the death is not. Mortality that 
was probably related to the fishery was quite high during this 
project. Such deaths may be rare in other areas. It is hoped 
that the alternate sets of calculations will make the data 
more general. 


The only pup to die during dependency was a male pup that 
died after being hit by a boat propeller. Thus, the survival 
rate for dependent pups is 1.0 when human caused deaths are 
not considered but is slightly lower when they are. When the 
boat strike was included, the survival rate for dependent 
males seemed low at 0.70. It is, perhaps, more insightful to 
consider that 20 of 21 (95%) male pups survived the average 
of 48 days between being instrumented and being weaned. All 
females (N = 15) survived the comparable period. 


181 


TABLE 9.4 - Survival rates of sea otter pups in Prince William 
Sound. 


ALL MORTALITY NON-HUMAN MORT. 


FEMALES MALES FEMALES MALES 
DAY YR DAY YR DAY YR DAY YR 
DEPENDENT PUPS 
11.0. 109439990; 270 ‘l HOe aie. 0 a! SONU HIS0 
(657) (1012) (657) (1012) 
INDEPENDENT JUVENILES 
ASSUMP. 1 H9969W 2) ne 9982 oS) mec 9969ea320) BOOSOhL eo 
(2917) (2720) (2917) (2721) 
ASSUMP. 2 .9966 .29 .9923 .20 .9966 .29 .9966 .30 
(2917) (2721) (2917) (2721) 
ASSUMP. 3 .9974 .38 .9982 .52 .9974 .38 .9989 .68 
(3410) (2815) (3410) (2815) 
ASSUMP. 4 N9O7Aw 34) 19957 624) 997 es 4ey 99 74e es 
(3410) (2815) (3410) (2815) 


* Numbers in parentheses are “otter days". 


ASSUMPTIONS : 
Weaned 1: Died day of last sighting, missing pups are 
alive. 
Weaned 2: Died day of last sighting, missing pups are 
dead. 


Weaned 3: 
Weaned 4: 


Died day carcass found, missing pups are alive. 
Died day carcass found, missing pups are dead 


The survival rates of weanlings are lower than those of 
dependent pups. Female rates are the least variable, ranging 
from) O29) .—9 O39 depending upon assumptions made in 
calculations. The exclusion of human-caused deaths has no 
effect on rates since no females died as a result of the 


fishery. As noted, females were caught in fishing gear, but 
all were released. Male survival rates are quite variable, 
0.21 - 0.70, depending upon the assumptions made in 


calculations and whether human-caused deaths are included. 
Changes in frequency of monitoring had little effect. 
However, potential mistakes in determining weanling survival 
(i.e. incorrect assumptions about the status of missing 
individuals) could lead to a change in survival rates by a 
factor of 2. 


DISCUSSION 
The results of this study indicated that there exists 
substantial variation in the timing of births, growth rates 


and dependency periods between individual sea otters. Males 


182 


grew more rapidly than females, were weaned at slightly 
heavier weights and thus, apparently required more parental 
resources (i.e. food) than did females. However, male and 
female births were timed similarly, and dependency periods 
were of approximately equal duration. 


Data indicated that young sea otters were capable of 
surviving independently long before they reach the typical 
age of weaning (i.e. approximately 5-6 months), even when they 
were weaned prematurely by human interference. In general, 
however, weanling survival rates were low during the first 
year of independence. Male weanlings were more likely to 
survive their first year of independence than were female 
weanlings. Observed differences between male and female 
weanlings in rate and timing of mortality existed and appeared 
to be strongly related to whether weanlings established their 
post-weaning home range in male areas or female areas. 


These data offer some support for the body of theory that 
suggests that sea otters should make a larger parental 
investment in male offspring. 


PI and growth rates 


The faster growth exhibited by males suggests that 
females may allocate more resources to dependent sons than to 
dependent daughters. Based on the arguments of Trivers and 
Willard (1973), it is possible that only the females that can 
best do so, produce males. Such females might be those that 
were older, larger or more experienced. This hypothesis 
requires the implicit assumption that there exists some 
mechanism whereby these individuals would tend to conceive 
sons, or possibly abort females (see Clutton-Brock and Albon, 
1982, for a discussion of such mechanisms). Substantial 
variation in pup sex ratios was observed in this study, but 
it is not possible to say whether such variation was related 
to maternal factors in any systematic way. 


Another way in which a female could allocate more 
resources to a son would be trade off future reproductive 
potential in order to produce an adequate male, once one is 
conceived (reviewed in Clutton-Brock and Albon, 1982). The 
faster growth of sea otter males during dependency suggests 
that male pups may require greater investment from their 
mothers than do female pups. It seems unlikely that males 
assimilate nutrients more effectively than females, since 
growth rates of male mammals are usually more strongly 
affected by food shortages than those of females (Widdowson, 
1976). Unless food is unlimited, a male's greater need for 
milk or solid food could negatively affect its mother's 
health, both by affecting her overall condition and/or by 
causing imbalances of critical elements at the time of 


183 


weaning. These "costs" to the mother could be manifested in 
effects on future reproductive potential through, for example, 
skipped or aborted pregnancies and reduced life span. It is 
possible that such "costs" exist for females that produce 
sons, especially if they live in habitats that have been 
heavily exploited. For example, higher rates of in utero 
mortality and lower overall rates of reproduction have been 
observed in sea otter populations that have over-exploited 
their food supplies (Schneider 1978b; reviewed in 
Simon-Jackson and Rotterman 1987). 


Timing of parturition 


Other information on sex differences in birth dates is 
not available. However, the observed late May peak of pupping 
is consistent with observations from other locations within 
Alaska. Barbash-Nikiforov, et al. (1978) observed a peak 
during May-June in the western Pacific. Schneider (1978b) 
reported a May peak in the central and western Aleutian 
Islands. 


Dependency period 


Dependency period lengths were quite variable for pups 
in this study and for pups in other studies at different 
locations throughout the sea otter's range. Data from research 
carried out along the eastern coast of the Soviet Union 
(Barabash-Nikiforov, et al., 1978), and in the western and 
central Aleutian Islands (Kenyon, 1969; Schneider, 1978b) 
suggested that females in some populations nursed pups for 
approximately one year. Other studies in Alaska and 
California have reported dependency periods averaging 5-7 
months (Garshelis, et al., 1984; Wendell, et al., 1984; Payne 
and Jameson, 1984). Dependency periods in this study ranged 
from 2.5-11 months. 


Several variables might contribute to variation in 
dependency periods. The underlying food supply could affect 
dependency periods both by influencing the development of 
young animals directly, and by affecting them. indirectly 
through the mother. It is likely that pups must reach a 
minimum size and minimum level of experience to be able to 
survive, in a given environment, upon independence. Food 
abundance should affect growth rates and, hence, the time it 
takes to reach that minimum size. If food was abundant, 
shorter dependency periods could result. Also, food abundance 
might affect the minimum size and experience requirements. 
If food was abundant, less precocious weanlings might be able 
to feed more effectively, mistakes would be less critical and 
thus, pups could be weaned earlier. If females were in good 
condition, they might be able to raise a pup more quickly or, 


184 


conversely, better afford to continue to support one. The 
same might be true for older, more experienced, females. 


Sex ratio 


No data are available from this study that would give any 
insight into the reason for the variation in pup sex ratio 


that was observed. However, a number of maternal or 
environmental factors are believed to be correlated with 
biased offspring sex ratios in mammals. These include: 


maternal age, parity, reproductive history, dominance status, 
size or nutrition and birthdate, litter size or timing of 
conception. Recent reviews of these factors, and relevant 
theories, are available in Clutton-Brock and Albon (1982) or 
Clutton-Brock and Iason (1986). 


The only data that are available on the sex ratio of 
young sea otters, other than this study, are from Amchitka 
during the 1960's (Kenyon 1969). Kenyon (p. 206) reported 
that the sex composition of 117 recently "deserted" or 
dependent "juveniles" (individuals of less than one year of 


age) was 58 males, 58 females, and one unknown. Schneider 
(1978b) and Kenyon (1969) reported fetal sex ratios from 
studies in the central and western Aleutian Islands. 


Combining their data, of 319 fetuses, 171 (57%) were females 
and 138 (43% males). 


Survival rates and causes of mortality 


Parental investment theory suggests that sex differences 
in post-weaning behavior, especially movement patterns, could 
lead to male and female offspring being subjected to different 
risks. Different selection pressures could, in turn, lead to 
different mortality patterns. Mortality rates and patterns 
were different for male and female sea otter offspring. 
Unexpectedly, females (the nondispersing sex) exhibited lower 
survival rates, during the first year after weaning, than did 
males. The trend toward a higher rate of male post-weaning 
survival in this study was similar to the trend in juvenile 
survival in California (Chapter 2). 


Comprehensive discussions of causes of sea otter 
mortality are available elsewhere for California (Riedman, 
1986) and Alaska (Kenyon, 1969; Simon-Jackson and Rotterman, 
1987). These include: starvation, disease, parasitism, 
predation, shark attacks, accidents during research projects, 
entanglement in fishing gear, adverse weather conditions 
(storms and icing), boat strikes, and injuries received from 
conspecifics during fighting or mating. 


The results of questionnaires (Simon-Jackson, 1985) and 
surveys (Matkin and Fay, 1980; Simon-Jackson, 1986) have 


185 


indicated that there is a significant mortality of sea otters, 
incidental to the salmon fishery in Prince William Sound and 
on the nearby Copper River Delta. Incidental take of sea 
otters in commercial fisheries is not unique to, but is 
especially significant in the Cordova vicinity (Simon-Jackson, 
1986; and reviewed in Simon-Jackson and Rotterman, 1987). Sea 
otter encounters with fishing gear are frequent and increasing 
on the Copper River delta and within the Prince William Sound. 
It is known that many untagged otters have died in fishing 
gear, either from drowning or from fishermen killing them 
directly. 


Why is first winter survival high in the male area? 


The male area is mostly contained within shallow, 
protected channels within estuarine mud flats that contain 
abundant shellfish species. Such habitat is scarce in the 
weaning areas. Food was considered to be abundant in the 
local male area when otters first recolonized it (Garshelis, 
et al., 1986) and apparently still is. Otters in the male 
area are frequently killed by natives, fishermen, and others. 
In such cases, recovered carcasses were examined and found to 
have significant subcutaneous fat. Moreover, stomachs usually 
contained large quantities of clam and/or crab tissues (pers. 
obs.). In addition to abundant food, the local male area 
offers young otters other benefits. There are large 
aggregations of males of all ages, so opportunities for 
learning from adults also exists (Garshelis, et al., 1984). 
Moreover, large predator-free, sand bars are exposed at 
moderate tides. These are heavily used as haul-out areas, 
especially during the winter. The energetic advantages of 
such behavior in cold water environments could be substantial. 


Why are pups weaned in bays where they are likely to die? 


Some weanlings apparently died of starvation during the 
first few months after weaning. The apparent maladaptive 
practice, of females weaning pups in areas where they have 
little chance of over-winter survival, may be a consequence 
of changing, conflicting needs during the development of the 
pup. On one hand, the deep, protected northern bays offer 
shelter and abundant pup food in the blue mussels (Mytilus 
edulis) which cling to the steep rock walls and grow in dense 
beds in the shallows. Opportunities exist for pups to learn 
self feeding in protected channels and lagoons. On the other 
hand, these bays tend to be deep and steep-sided so other 
larger prey may be relatively unattainable to small, newly 
independent otters whose diving abilities are usually limited. 
Although it seems probable that small juveniles can survive 
on a diet of mussels alone, it is not clear whether large 
juveniles can. 


Predation appears to be a significant source of mortality 
of young sea otters in the weaning areas. Moreover, the rate 
of predation could be exacerbated by juveniles' tendency to 
haul out when subject to food stress, as would occur under the 
preceding argument. Predation on hauled sea otters has not 
been noted before this study and may be uncommon. If so, it 
would be unlikely that adult sea otters would be prepared to 
respond to the possibility when choosing sites for weaning 
pups. 


Recolonization and future changes 


The Prince William Sound sea otter population is still 
recovering from near extirpation, which took place at the end 
of the 18th century (Lensink, 1962). The bays in which 
weaning was observed in this study, Simpson Bay, Sheep Bay 
and Port Gravina, were reoccupied by large rafts of males 
between 1970-1979 (Garshelis, et al., 1986). 


While it is risky to predict the patterns that will 
develop for otters in this area when the population approaches 
equilibrium, a few speculations can be made. It is likely 
that food resources in the three bays have been depressed as 
a result of the relatively recent occupancy by large rafts of 
males (Garshelis, et al., 1986). However, it is possible that 
food levels will eventually rise above current levels, after 
the population reaches equilibrium (e.g. Estes, et al., 1978; 


Estes, et al., 1982). If food levels do rise, juvenile 
starvation may decrease in those bays. Moreover, deaths from 
predation may also decrease. Predation can occur when 


individuals haul out on beaches. Moribund sea otters usually 
haul out (Kenyon 1969), which would be the case if food was 
scarce and they were starving. Thus, if food became more 
abundant the frequency of hauling-out and, hence, predation 
might decline. Conversely, it might be predicted that food 
abundance in the male area will decline, at least initially, 
after a longer period of sea otter occupancy. If so, 
starvation would increase. Predation, however, might remain 
unaffected since numerous safe haul out sites are available 
on sand bars in that area. 


LITERATURE CITED 


Ames, J. A.; R. A. Hardy; and F. E. Wendell. 1983. Tagging 
materials for sea otters, Enhydra lutris. Calif. Fish 
and Game 69:243=-252. 


Barabash-Nikiforov, I. I., S. V. Marakov and A. M. Nikolaev. 


1968. The kalan or sea otter. Nauka Press, Leningrad. 
184 pp. 


187 


Calkins, D. G. and P. G. Lent. 1975. Territoriality and 
mating behavior in Prince William Sound sea otters. J. 
Mamm. 26:528-529. 


Clutton-Brock, T. H.; P. H. Harvey and B. Rudder. 1977. 
Sexual dimorphism, socionomic sex ratio and body weight 
in primates. Nature 269:797-799. 


Clutton-Brock, T. H. and S. D. Albon. 1982. Parental 
investment in male and female offspring in mammals. In: 
Current Theories in Sociobiology. (eds. Kings College 
Sociobiology Group), Cambridge. 


Clutton-Brock, T. H. and G. R. Iason. 1986. Sex ratio 
variation in mammals. Quarterly Review of Biology 
61:339-374. 


Estes, J. A.; N. S. Smith and J. F. Palmisano. 1978. Sea 
otter predation and community organization in the western 
Aleutian Islands, Alaska. Ecology 59:822-833. 


Estes, J. A.; R. L. Jameson and E. B. Rhode. 1982. Activity 


and prey selection in the sea otter: influence of 
population status on community structure. Am. Nat. 
120:242-258. 

Garshelis, D. L. 1983. Ecology of sea otters in Prince 


William Sound, Alaska. Ph.D. dissertation. University 
of Minnesota, Minneapolis, Minnesota. 330 pp. 


Garshelic (o9D. Le and’™D. 75.) /Sinatt- 1983. Evaluation of 
radio-transmitter attachments for sea otters. Wildlife 
Society Bulletin 11:378-383. 


Garshelis, D. L. and J. A. Garshelis. 1984. Movements and 
management of sea otters in Alaska. J. Wildl. Manage. 
48:665-678. 


Garshelis, D. L.; A. M. Johnson and J. A. Garshelis. 1984. 
Social organization of sea otters in Prince William 
Sound, Alaska. Canadian Journal of Zoology 62: 
2648-2658. 


Garshelis, D. L., J. A. Garshelis and A. T. Kimker. 1986. 
Sea otter time budgets and prey relationships in Alaska. 
Journal of Wildlife Management 50:637-647. 


Greenwood, P. J. 1980. Mating systems, philopatry and 
dispersal in birds and mammals. Animal Behavior 28: 
1140-1162. 


188 


Glucksman, A. 1974. Sexual dimorphism in mammals. 
Biological Review 49:423-475. 


Glucksman, A. 1978. Sex determination and sexual dimorphism 
in mammals. London, Wykeham. 


Hamilton, W. D. 1964. The genetical evolution of social 
behaviour. I. J. Theoret. Biol. 7:1-16. 


Kenyon, K. W. 1969. The sea otter in the eastern Pacific 
Ocean. USFWS. North American Fauna, No. 68, 352 pp. 


Lensink, Cc. J. The history and status of sea otters in 
Alaska. Ph.D. dissertation, Purdue University, 
Lafayette, Indiana. 186 pp. 

Loughlin, T. R. 1977. Activity patterns, habitat 
partitioning, and grooming behavior of the sea otter, 
(Enhydra lutris), in California. Ph.D. thesis, 


University of California, Los Angeles, CA. 110 pp. 


Matkin, C. 0. and F. H. Fay. 1980. Marine mammal - fishery 
interactions on the Copper River and in Prince William 
Sound, Alaska, 1978. U. S. Dept. of Commerce, National 
Technical Info. Serv. PB 80-159536. 71 pp. 


Payne, S. F. and R. J. Jameson. 1984. Early behavioral 
development of the sea otter, (Enhydra lutris). J. Mamn. 
65:527-531. 


Reiter, J.; N. L. Stinson and B. J. Le Boeuf. 1978. Northern 
elephant seal development: the transition from weaning 
to nutritional development. Behav. Ecol. Sociob. 
3:336-367. 


Riedman, M. L. 1986. Draft environmental impact statement of 
proposed translocation of southern sea otters. Volume 
II: Technical support documents. United States Fish 
and Wildlife Service and University of California, Santa 
Cruze Ca. 


Schneider, K. B. 1978a. Sex and age segregation of sea 
otters. Alaska Department of Fish and Game, Final 
Report, Federal Aid Wildlife Restoration Projects W-17-4 
to W-17-8. 


Schneider, K. B. 1978b. Reproduction in the female sea otter 


in the Aleutian Islands. Unpublished Report, Alaska 
Department of Fish and Game. 44 pp. 


189 


Simon-Jackson, T. 1985. Fishermen opinions of sea 
otter/fisheries issues in Alaska. USFWS, unpubl. rept. 


17 pp. 
Simon-Jackson, T. 1986. Sea otter survey, Cordova, Alaska 
1986, (Orca Inlet to Cape Suckling). USFWS, unpubl. 
rept. 


Simon-Jackson, T. and L. M. Rotterman 1987. The sea otter in 
Alaska (Enhydra lutris): Species account with research 
and management recommendations. Prepared for the Marine 
Mammal Commission, Washington, D. C. 


Siniff, D. B., T. D. Williams, A. M. Johnson and D. L. 


Garshelis. 1982. Experiments on the response of sea 
otters, Enhydra lutris, to oil contamination. 


Biological Conservation 2:261-272. 


Trent, T. and T. and O. J. Rongstad. 1974. Home range and 
survival of cottontail rabbits in southwestern Wisconsin. 
J. Wildl. Manage. 38:459-472. 


Trivers, R. L. 1972. Parental investment and sexual 
selection. In: Sexual selection and the descent of man, 
1871-1971. (ed. B. Campbell), pp. 136-179. Chicago: 
Aldine. 


Trivers, R. L. and D. E. Willard. 1973. Natural selection 
of parental ability to vary the sex ration of offspring. 
Science 179:90-91. 


Wendell, F.E.; J. A. Ames and R. A. Hardy. 1984. Pup 
dependency period and length of reproductive cycle: 
estimates from observations of tagged sea otters, 


(Enhydra _lutris), in California. Calif. fish and Game 
70:89-100. 


Williams, T. D.; A. L. Williams and D. B. Siniff. 1981. 
Fentanyl and azaperone produced neuroleptananalgesia in 


the sea otter (Enhydra lutris). Jie A Walaa) Dilsi 
17:337-342. 


Widdowson, E. 1976. The response of the sexes to nutritional 
stress. Proc. Nutr. Soc. 35:1175-1180. 


190 


CHAPTER 10 


A SIMULATION MODEL FOR ASSESSING THE RISKS OF OIL SPILLS TO 
THE CALIFORNIA SEA OTTER POPULATION AND AN ANALYSIS OF THE 


HISTORICAL GROWTH OF THE POPULATION 


A. BRODY 


NOVEMBER 30, 1988 


191 


INTRODUCTION 


The model described in this chapter is designed to 
facilitate analysis of the risk of oil spills to the 
California sea otter population. Specifically, we provide a 
simulation of aspects of otter population biology and behavior 
that will likely affect the degree of risk to the population 
associated with oil spills. We have not conducted actual risk 
analysis with the model. 


Ford, et al., (1982) describe two general categories of 
oil spill consequences that affect the degree of risk of oil 
development activities to a wildlife population: 1) immediate 
mortality from a given oil spill, and, 2) long term population 
effects. Our model addresses the first category explicitly, 
and can be used to address the second category if the long 
term effects of oil development are assumed to result only 
from mortality due to oil spills. Our model was formulated 
to answer two specific questions about a given oil spill: 1) 
how many otters will be killed, and 2) how long will it take 
for the population to recover. 


General approach 


There is a good deal of uncertainty surrounding certain 
aspects of sea otter ecology in California, and a general lack 
of data regarding most aspects of the population dynamics. 
There are also ongoing research projects being conducted by 
various agencies and organizations aimed at correcting these 
situations. A major objective in structuring our model was 
that it be flexible enough to explore a variety of conceptual 
hypotheses about sea otter ecology, and to incorporate new 
data as they become available. 


As with any simulation model, some simplifying 
assumptions about the system are necessary. Four major ones 
will be discussed in this introductory part of this chapter, 
others will be mentioned at appropriate places later on. 


1) Geometry of sea otter range.-- The sea otter range in 
California is essentially linear, consisting of an 
approximately 350 km long by 1 km wide band along the coast. 
The width of the range is dictated by the slope of the ocean 
floor, with otters generally inhabiting only areas shallower 
than 18m (USFWS 1986); in some areas of the coast the 18m 
depth contour is more than 1 kilometer offshore, and animals 
may occasionally be found at locations of deeper depth. In 
fact, telemetry data gathered as part of this project 
indicates that certain segments of the population may utilize 
offshore areas more frequently than previously realized (see 
Chapter 3). But in the model we consider the range to be one 
dimensional, a line extending up and down the coast. 


192 


California Department of Fish and Game has traditionally used 
the "as the otter swims line" (Ames, Hardy and Wendell, 
personal communication), an ordinate system coincident with 
the 5 fathom (10m) depth contour, in their census activities. 
With an origin at Coyote Point on the San Francisco peninsula, 
position along the coast is measured in 500m units south of 
the origin. We have adopted this system for the spatial 
aspects of our model. This system allows us to incorporate 
existing census data, to ignore the shape of an oil spill and 
be concerned only with the length of coast affected, and to 
facilitate the analysis and modeling of sea otter movement. 


2) Density dependence in population dynamics.--A good 
deal of controversy surrounds the role of density dependence 
and the question of non-equilibrium in the current theories 
of population dynamics. The issue is complicated by the range 
expansion that has accompanied population growth, differences 
in habitat quality at various locations within established and 
potential range, and the role of the sea otter as a "keystone 
species", able to dramatically affect the quality of its own 
habitat (Miller 1980, Estes, et al., 1982, Estes, et al., 
1986, Wendell, et ae 1986). Rather than assume density 
dependent or density independent dynamics, we have built in 
a flexibility that allows investigation of both. The 
population dynamics portion of our model nominally assumes 
density dependent growth during recovery from an oil spill. 
The equilibrium population size and a parameter governing the 
shape of the density dependence function are easily 
manipulated at the beginning of a model run, however, and 
density-independent population growth can be simulated by 
setting parameter values that result in an essentially flat 
density dependence function. 


3) Range expansion.--Related to range geometry and the 
nature of population dynamics is the question of range 
expansion. It is an especially critical question given the 
purpose of the model, as the offshore areas most likely to be 
developed for oil and gas in the future are at the southern 
periphery of the existing sea otter range (USFWS, 1986). 
Thus, if we are to make reliable predictions about the effects 
of oil spills very far into the future we must be able to make 
predictions about the extent of sea otter range in the future. 
Unfortunately, data collected during this study do not easily 
lead to such predictions. We assume in the simulation model 
that the extent of the range is static, and that otters 
distribute themselves along the coast in the same relative 
proportions regardless of total population size. We have, 
however, built a separate, small, deterministic model that 
will generate predictions of future range length, carrying 
capacity, and population size based on historical rates of 
population growth and range expansion. This model, OTRANGE, 
described elsewhere in this report, can be used to generate 


193 


range conditions for use in the population model. It only 
applies to the existing sea otter range in California and does 
not extend to the translocation of otters to San Nicholas 
Island. 


4) Impact criteria.--In line with the stated purpose of 
the model, three criteria are assumed to measure the effect 
of an oil spill on the population. The first criterion is 
the number of animals and proportion of the population killed 
by the spill. The second criterion is the number of years 
after the spill that is required for the population to recover 
to pre-spill size. The third criterion is the decrease in the 
total reproductive value of the population (Wilson and Bossert 
1971), providing a measure of how the perturbation in age and 
sex structure caused by an oil spill may affect population 
dynamics. This is calculated using the survival and 
reproduction rates operative in the population just prior to 
the spill, before density dependent adjustments to these rates 
are effected. In addition to these three criteria the model 
can be run in a "control" mode, in which an oil spill is not 
introduced, allowing graphic comparison of population dynamics 
with and without perturbation by oil spill. 


Model structure 


We model the effect of an oil spill on the sea otter 
population as being determined by the size and location of 
the spill in relation to sea otter distribution, the movement 
of individual sea otters in the vicinity of the spill, and 
sea otter population dynamics. 


The model itself consists of four submodels imbedded into 
a larger program superstructure (Fig. 10.1). The submodels, 
all of which are stochastic, operate in different temporal and 
spatial scales. The short term population submodel (OTPOP) 
operates with a time step of one month and is spatially 
independent. The long term population model (LESLIE) operates 
on an annual time scale and is also spatially independent. 
The sea otter distribution model (OTDIST) operates on a 
Spatial scale of kilometers, and is time independent. The 
short term movement and oil response model (OTMOVE) operates 
with a time step of days, and on a spatial scale of 
kilometers. Each of the submodels operates on a numerical 
scale of individual animals. 


OTPOP creates a simulated initial population, and 
iterates for three simulated years before an oil spill. In 
the month of the simulated spill, population vectors, 
consisting of the age and reproductive status of each 
individual animal in the simulated population, are passed from 
OTPOP to OTDIST. OTDIST assigns each individual a position 


194 


FIGURE 10.1 -- Schematic representation of the interrelation of 
the 4 submodels used to predict the potential effects of oil 
spills on California sea otter population dynamics. 


INITIALIZE LESLIE END 
PARAMETERS 
AND CONSTRUCT Annual 
INITIAL POPULATION Population 


Model 


population 


OTPOP 


Monthly 
Population 
Model 


vectors 


population 
vectors 


mortality 
vectors 


OTMOVE OTDIST 


coh a Short—term Rp here tb 
HERE —» | Movement ode 


Model 


along the coast, and passes these vectors to OTMOVE. OTMOVE 
introduces an oil spill, moves animals along the coast during 
the duration of the spill, and generates a mortality vector. 
This mortality vector is then returned to OTPOP, and two years 
of population recovery are simulated on a monthly basis. At 
the end of the second year after the simulated spill, the 
population vectors representing individual animals are 
collapsed into age class vectors and passed to LESLIE. LESLIE 
then simulates the future course of population growth on an 
annual basis for up to 50 years. 


While the structure of the model is essentially set, and 
can be altered only through reprogramming of the source code, 
the parameters used in the model are set each time the model 
is run. We have supplied a set of default parameters, 
representing our best estimates of the values operating in 
the real population, but these may be altered by the user to 
investigate the importance or sensitivity of parameters or to 
take advantage of revised parameter estimates that may be 
available after future research. Following is a detailed 
discussion of the structure of each of the submodels, and the 
logic that we followed in arriving at the default parameter 
estimates. 


OTPOP AND LESLIE 
Structure 


The small size of the population allows a reasonably 
efficient consideration of individual animals. The 
reproductive biology of sea otters, specifically the fact that 
pupping is spread throughout the year rather than concentrated 
into a short reproductive season, adds a complexity to the sea 
otter model that is not present in population models of most 
other large mammals. OTPOP thus iterates on a monthly basis. 


At the beginning of each run the user specifies an 
initial population size, a carrying capacity for the range, 
a maximum population growth rate, and a parameter governing 
the shape of the density dependence (population growth rate 
vs. population size) curve. The initial growth rate of the 
population is calculated on the basis of these parameters and 
used to construct an initial population vector with a stable 
age distribution. 


The age, in months, of each male in the population is 
stored in a male age vector. The age, in months, of each 
female in the population is stored in a female age vector. 
The reproductive status of each female in the female vector 
is stored in the 
corresponding element of a reproductive vector. 


196 


During each month of simulation the model loops through 
each individual in the population, drawing a random number 
that is compared against age- and sex- specific monthly 
survival rates to determine whether or not the animal survives 
the month. In the female loop the reproductive status of each 
surviving female is checked. If a female has a pup the 
survival of that pup is determined by drawing a random number 
and comparing it to monthly pup survival rates. If a female 
has no pup, she becomes pregnant with probability determined 
by her age and the month of the year. Inter-uterine mortality 
is assumed to be zero so that if a pregnant female survives 
a given month her fetus automatically survives also. 


The age of any animal that does not survive the month is 
flagged. If a pup dies, the mother is assumed to get pregnant 
again immediately. If a pup survives until weaning eas 
assigned a sex randomly and added to the appropriate age 
vector. At the end of each monthly loop through the 
population the vectors are reloaded without the animals that 
have died, and numbers are totaled and reported. 


The model is allowed to run for three simulated years 
before an oil spill is introduced, to subject the initial 
stable age distribution to stochastic fluctuations. In the 
year of the spill, OTDIST and OTMOVE are called and subject 
the population vectors to oil spill induced mortality. oOil 
spill mortality is considered after the regular loop for that 
month has been completed, but after the population vectors 
are reloaded. Thus simulated oil spill mortality is strictly 
additive to the simulated natural mortality. 


OTPOP continues to iterate on a monthly basis for two 
simulated years after the spill, with population growth rates 
determined by the density dependence function annually. Two 
years after the spill the population vectors are collapsed 
into age class vectors and passed to LESLIE. Pups and fetuses 
are grouped into age class 0, age class 1 contains animals 
between seven and 18 months of age, age class 2 contains 
animals between 19 and 30 months, etc. The numbers of animals 
of each sex in age class 0 are determined by a random draw 
from a binomial distribution assuming a 0.5 probability of 
being either sex. 


LESLIE, iterating on a yearly basis, runs much faster 
than OTPOP at the expense of seasonality. Population sizes 
are reported once a year, at the end of the month in which 
the oil spill occurred. Survival and reproductive rates based 
on the density dependence function are determined at each 
iteration, just as in OTPOP. Two sources of variation, after 
Harris, et al., (1987), are considered explicitly in LESLIE. 
The first is a "demographic" stochasticity wherein it is 
assumed that all animals of the same age and sex have the same 


197 


probability of surviving the year, but the number that 
actually do survive is determined by a random draw from a 
binomial distribution with parameters (n,,S,) where n, is the 
number of animals in age class x and s, is the annual survival 
rate for age class x. If n, is greater than 30 the normal 
approximation to the binomial is used to reduce the computer 
time required. The second source of variation is 
"environmental" stochasticity that operates simultaneously 
across all age classes to the same degree in a given year. 
This is introduced to simulate the occurrence of "good years" 
and "bad years". The bounds of this stochasticity are 
determined by the user at the beginning of the run, and p, 
the environmental stochasticity parameter, is assumed to be 
uniformly distributed between those bounds. At the beginning 
of each simulated year p is determined once by random draw 
and the survival rate, s,, of each age class for that year is 
modified by adding ps,. 


Built in to the structure of LESLIE is the ability to 
consider density independent mortality. This is included 
primarily because of the possibility that incidental gill- 
and trammel-net mortality, which has substantially affected 
population growth in recent years (Ames, et al., 1985, 
Wendell, et al., 1985, Estes, et al., 1986) operates in a 
density independent fashion. The user may set a density 
independent mortality rate, applied to all age classes, and 
also the degree to which this density independent mortality 
compensates for density dependent mortality. 


Theoretical framework for parameterization of OTPOP and 
LESLIE. 


Field data that can be used to infer the dynamics of the 
California sea otter population are scarce. Raw data 
available to us included aerial and ground censuses conducted 
by the CDFG and the U.S. Fish and Wildlife Service (1968- 
1985), records from CDFG and USFWS carcass recovery efforts 
(1968-present), monthly counts from several CDFG and USFWS 
index areas (1976-1982), and our own live-capture and 
telemetry data. In addition to this raw data, information on 
sea otter population dynamics was gleaned from several 
publications and manuscripts. 


The information available to us differed greatly in 
reliability, in relative quantity, and in the extent to which 
it was directly applicable to our purposes. Bringing all of 
the information together into a single flexible yet 
comprehensive model required the use of a strong theoretical 
framework, necessary both for evaluation of the available 
data and for estimating model parameters. The theoretical 
framework that we used was developed by Eberhardt (1985) 


198 


following work by Siler (1979), and based on the classical 
age structure models of Lotka (1907). 


Survivorship.--Siler (1979) and Eberhardt (1985) viewed 
survival to any age as a function of three "competing risks". 
The first risk is that of "early hazards", risks which are 
associated with the early years of life, but become much less 
important as the animal matures. In sea otters these early 
hazards are most likely to be associated with dispersal from 
the natal area after weaning, and with low competitive ability 
relative to older, more experienced animals. The early 
hazards risk is greatest in the first years of life, but 
essentially zero after maturity. The second competing risk 
is that of incidental, or "constant hazards", which in sea 
otters would include those due to possible predation, severe 
weather, infectious disease, and possibly accidental 
entanglement in commercial gill and trammel nets. This risk 
is seen as constant throughout an animal's life, and is the 
only important risk of mortality during the prime adult years. 
The last risk is that due to senescence, and in sea otters 
would include reduced competitive ability or increased 
susceptibility to disease due to old age. This risk is 
essentially zero through the prime adult years, but reaches 
100% by the maximum age. 


The probability of survival to any age, then, is the 
probability of surviving all three competing risks at that 
age, or the product of the three age specific survival rates 
(Fig. 10.2). Estimating the survivorship schedule inside of 
this framework allows calculation of survival rate for each 
age class without having to estimate each directly from field 
data. The number of parameters that have to be estimated is 
thus greatly reduced. Sea otters are long-lived; the data 
for directly estimating survivorship for perhaps 25 age 
classes are simply not available. Using Eberhardt's (1985) 
approach we need only to determine the form of the three 
competing risk curves. 


The equation for survivorship at age x, from Eberhardt 
(1985) is: 


1, = exp{-a;[1l-exp(-b;x)] - ax - as[exp(b3x) -1] } (1) 


where 1, is survivorship, a;, a2, az are the coefficients for 
early, incidental, and senescence risks, respectively; and 
b,, bz are parameters governing the shape of the early hazards 
and senescence curves, respectively. In this formulation the 
risk coefficients (a, a2, a3) are taken as -ln(S) where S is 
the annual survival rate against early hazards, incidental 


199 


FIGURE 10.2 -- Hypothetical survivorship curve depicting the 
relationships of the 3 competing risks of Siler (1979) and 
Eberhardt (1985). 


SURVIVORSHIP 


200 


hazards, and senescence, respectively. Separate survivorship 
curves can be generated for males and females using the basic 
equation with different parameter values. 


Reproduction. --As for survivorship, age specific 
reproduction can be modeled as a function of three component 
curves: an early reproductive function, a "prime" rate during 
adulthood, and a decrease in reproductive output during the 


years of senescence (Fig. 10.3). Again, conceptualizing 
reproductive rates in this manner greatly reduces the number 
of parameters to be estimated. If early reproduction 


increases asymptotically to the prime rate a separate term 
governing early reproduction is not necessary, and Eberhardt 
(1985) gives an equation for the reproductive curve: 


m, = A{1l-exp[-b,(x-C) ] }}exp[-azexp (b3x) -1] (2) 


where m, is the number of female offspring per year weaned by 
each female of age x, A is the maximum reproductive rate (in 
number of female offspring weaned per prime aged female), B 
is a parameter governing the rate of increase of early 
reproductive rate to the prime rate, C is the age before the 
age of first reproduction, and az and bz; are the senescence 
parameters as in (1). 


Population growth rate and density dependence. The per 
capita population growth rate is a central parameter in any 
population model. Recent investigations suggest that in 
marine mammal populations the dependence of population growth 
rate on population size is nonlinear, with the growth rate 
decreasing more and more rapidly as the population approaches 
carrying capacity (Eberhardt and Siniff 1977, DeMaster 1981, 
Fowler 1981). We generalized DeMaster's (1981) density- 
dependent relationship for survivorship to obtain a simple 
non-linear function for population growth rate: 


Yr = Ymax{1l-exp[—b(K-N) ] } (3) 


where r is the annual per capita growth rate, r,,, is the 
maximum annual per capita growth rate, N is the population 
size, K is the equilibrium population size, and b governs the 
shape of the curve. Because growth rates determined according 
to (3) decline unrealistically rapidly when the population 
size gets very far above the equilibrium level we impose an 
arbitrary floor on the growth rate at -r,,, (Fig. 10.4). 


Lett, et al., (1981) and Fowler (1981) point out that 
density dependence in age-structured populations may be 


201 


FIGURE 10.3 -- Hypothetical reproductive curve depicting the 
relationship between prime reproductive rate and senescence, 
after Eberhardt (1985). 


1.0 -—---------___ Sen. 


oe prime reproductive rate 


FECUNDITY 


0.0 


202 


FIGURE 10.4 -- The effect of the value of b on the non-linearity 
of the density dependence function used in OTPOP and LESLIE. XK 
is the carrying capacity. 


Em 
z 


b=s0.0025 


oO 


SEX ISe 


Per capita growth rate (r) 


Population size (N) 


203 


mediated through a variety of mechanisms including age at 
first reproduction, pregnancy rate, age of weaning, juvenile 
survival rate, and/or adult survival rate. Our California 
field work indicates that juvenile females are under more 
intra-specific competition pressure and suffer higher 
mortality rates than adults. Our model incorporates a 
hierarchical adjustment mechanism in the calculation of female 
parameters to achieve the annual growth rates specified by 
(3). Most density dependence is effected only through changes 
in the female early risk coefficient (a, in (1)); if the early 
survival rate drops below .40 or rises above the incidental 
survival rate, the incidental female risk coefficient (a2) is 
altered. If the incidental survival rate reaches 0.99 (i.e., 
extremely high values of r), the age at first reproduction (C- 
1 in (2)) is reduced. Senescence parameters, and male 
survival rates are held constant during a given run of the 
model. 


In an age structured population with a stable age 
distribution and a constant rate of growth the relationship 
between the survival and reproductive rates, and the growth 
rate of the population is described by Lotka's (1907) 
equation: 


lme™ = 1 (4) 


where x is age, 1, is female survivorship to age x, m, is the 
number of female offspring recruited to the population from 
each female of age x, and r is the rate of growth. Formulated 
in this manner, Lotka's equation holds strictly only for 
birth-pulsed populations (where all the reproduction takes 
place during a short discrete time period) at a stable age 
distribution. California sea otters reproduce throughout the 
year, and so are not birth-pulsed; the stochastic nature of 
the model precludes attainment of a stable age distribution, 
even at equilibrium population size, except by fortuitous 
coincidence. But we assume in the model structure and for 
purposes of parameter estimation that Lotka's equation 
provides an adequate approximation to the structure and 
dynamics of the California sea otter population. 


At the beginning of each time step in the model r is 
calculated according to (3), and then Lotka's equation is 
solved numerically by adjusting the female early hazards curve 
until the appropriate 1, schedule is found. Because of the 
violations in the assumptions of Lotka's equation the 
stochastic fluctuation in age structure, the survivorship 
schedule determined at the beginning of each time step will 
result in a population growth rate that only approximates the 
rate calculated in (3). 


204 


Parameter estimates 


Our approach in estimating population parameters was to 
begin with independently derived estimates, then vary them 
systematically to arrive at a set of parameters that: 1) were 
consistent with each other within the structural framework of 
the model, and 2) were in reasonable agreement with the 
original empirically derived estimates. 


Survival. We estimated annual survival rates from our 
telemetry data (see Chapter 2) using the method of Heisey and 
Fuller (1985). The ages at which these rates applied were 
estimated by calculating an average age for known aged 
telemetered animals weighted by the number of days each animal 
was observed: 


x x;d; / a; ;i=1,N (5) 


where x is the age at which the calculated annual survival 
rate applies, x; is the age of animal i as determined from 
tooth annuli (see Chapter 6), and d; is the number of days 
animal i was observed. The values in Table 10.1 show the 
effect on the estimates of using different age thresholds for 
distinguishing between juveniles and adults. We thus obtained 
pairs of age-specific annual survival rates for each sex that 
could be fitted to survival rates calculated from (1). 


Table 10.1. Estimates of annual survival rates of telemetered 
sea otters in California, as determined by the method of 
Heisey and Fuller (1985), 1983-1986. 


Juveniles Adults 
Age of Average Estimated Average Estimated 
adulthood age! survival age survival 
FEMALES : 
1 0.61 0.69 5.94 0.91 
2 1.41 0.78 7.06 0.92 
3 1.88 0.83 7.88 0.92 
4 2.16 0.83 8.09 0.91 
MALES: 
1 0.38 1.00 4.78 0.73 
2 1.26 1.00 6.21 0.68 
3 1.92 0.87 6.92 0.71 
4 1.94 0.76 6.94 0.77 


Separate estimates of annual survival can be obtained 
from the sample of otter teeth collected from museums and aged 


1 ‘5 2 
Weighted according to eq. (9) in text. 


205 


by counting annuli. The method of age determination and the 
problems associated with it are discussed in Chapter 6 the 
distributions of estimated ages are shown in Fig. 10.5. The 
number of animals represented in Fig. 10.5 is smaller than the 
total sample because many of the collected animals could not 
reliably be assigned a sex. The fact that, in the female 
data, the numbers of animals in the even age classes is 
consistently higher than those in the odd age classes (except 
for age 5) is unexplained. And the fact that there 

are apparently fewer animals of both sexes dying at age one 
than at age two is questionable, and may be a result of small 
carcasses not having the same chance of recovery as larger 
ones (due, perhaps to more rapid decomposition or lower 
visibility) or that incomplete dentition in some one year olds 
may have resulted in selection against them when the teeth to 
be aged were extracted from the museum collections. 


Despite the uncertainties in the tooth data, we did 
calculate annual survival rates using the "segment" method 
(Chapman and Robson 1960), which assumes constant annual 
survival after a threshold age, to compare to the estimates 
obtained from the telemetry data. Assuming a stationary 
population: 


d, = N(1, - 141) = N(1s"° - Ds FS) 
Hs 


(6) 


where N is the total population size, d, is the number of 
animals dying at age x, 1, is survivorship to age x, s is the 
constant annual survival rate, c is the threshold age, and H 
is the constant represented by: dl,(1-s)s‘s’. Using the 
Chapman and Robson (1960) regression to estimate a constant 
annual survival rate between ages four and ten (i.e. c =4 in 
(6), data for teeth older than 10 were not used) yielded a 
female rate of 0.925 (s.e. =0.045) and a male rate of 0.723 
(s.e. =0.038). If we assume that actual survivorship is as 
in (1), the ages from four to ten represent the segment that 
has survived the early hazards but is not yet greatly affected 
by senescence. If we further assume that the Chapman-Robson 
method estimates average survival during that age segment, we 
can calculate the age at which the estimate applies (as we did 
for the telemetry data) by calculating weighted average ages 
of the samples in the segment. The weighted averages imply 
that the female rate applies at age 7.63 and the male rate 
applies at age 6.40. 


The tooth data can also be used to estimate a,, the early 


hazard coefficient in (1). Ignoring the portion of the sample 
greater than 10 years old allows us to ignore the senescence 


206 


tters 


la sea oO 


-5 -- Distribution of ages of 425 Californ 


ted by tooth cementum annuli technique. 


FIGURE 10 
estima 


45 


40 


35 


Mmmm SS CMales 


ERR Females 


KAA AAK/ 
60504 


wo 
N 


YdsEWNnn 


20 


Ti, 
YL 


ALA 
O06, 


os 
13 


+ 


11. «12 


10 


ESTIMATED AGE 


207 


terms in (1), and, assuming that the early hazards are over 
by age two, the proportion of animals dying at ages one or two 
is: 


p = (1-s;s*)/(1-s-s'”) (7) 


where s,; is the survival rate against early hazards (i,e., 
-lIn(a,)) and sis the adult survival rate. Using the adult 
rates calculated above yields a female early risk coefficient 
of 0.051 (s- =0.95). The calculation for males yielded an 
illogical value of 0.270 (s;=1.31), indicating that the number 
of 1 year olds may indeed be underrepresented in the sample, 
and making the female early survival estimate suspect also. 


Reproduction.--Since m, and A in (2) are in terms of 
offspring weaned, they in turn are functions of pregnancy 
rates 
and pup survival rates. Otters give birth to a single pup, 
twinning is rare enough to be neglected in calculations. 
Loughlin et al (1981) suggest a gestation period of four to 
six months, a pup dependency period of four to eight months, 
and annual reproduction. Wendell et al (1984) report annual 
reproduction and a pup dependency period of five to eight 
months. The longest period that we observed a radioed female 
associating with a pup was about six months. We use constant 
gestation and pup dependency periods of six months each in our 
model. The pupping interval observed in our telemetered 
animals was not significantly different from one year (see 
Chapter 2). 


Our telemetry data yielded a pup survival (from birth to 
6 months) estimate of about 0.50 (see Chapter 2). Assuming 
no interuterine mortality, a gestation period of six months, 
and a pup dependency period of six months, this rate 
translates directly to an annual survival rate of fetuses. 

A separate estimate of the ratio of pup survival to adult 
survival was derived from unpublished CDFG data collected in 
index areas at monthly intervals between 1976 and 1982 
(Wendell et al 1986). These data were counts of independent 
otters and pups, and the pups were divided by size into a 
small stage (here assumed to be from 0-3 months old) anda 
large stage (here assumed to be from 3-6 months old) (Fig. 
10.6). Average relative pup survival rate was estimated by 
contrasting relative numbers of large pups with the peak 
number of small pups three months earlier. If s, is the 
average survival of independent otters in a given three month 
period, and sp is the survival from the small pup stage to the 
large pup stage, relative numbers of independents (I), large 
pups (L), and small pups (S) can be expressed as: 


L/I = (S/T) (So/S1) (8) 


208 


FIGURE 10.6 -- Relative average number of small pups and large 
pups, by month, in the CDFG index areas, 1977-1984. 


SMALL PUPS 


PUPS PER INDEPENDENT 


0.10 
0.08 a. LARGE PUPS 
0.06 z 
0.04 9 
0.02 


0.00 


PUPS PER INDEPENDENT 


and, if q is average pup survival relative to adult survival, 
q = S/S; = (L/1I)/(S/T) (9) 


From the CDFG data q = 0.439/0.607 = 0.723. The observations 
were taken three months apart; if we assume that q measures 
survival over a three month interval, the relative monthly 
survival is the cube root of gq, =0.898. This relative rate 
can be used to calculate an absolute estimate of pup survival 
once the adult survivorship schedule is determined. 


Using the pup survival estimate derived from telemetry, 
and assuming a 99% pregnancy rate for prime aged animals and 
an even sex ratio at birth, the prime weaning rate (A in (2)) 
is 0.50*0.99*0.5 = 0.247. In the model pup survival is held 
constant over all maternal age classes, and age specific 
weaning rates are achieved by varying pregnancy rate with age. 


The age of first reproduction for female sea otters in 
Alaska appears to be about four years (Schneider 1972), 
although limited observation in California suggests that some 
otters may reproduce as early as three years. If age at first 
reproduction is taken as 4 the parameter C in (2) is set at 
3. 


Senescence.--In sea otters, which are not subject to 
heavy natural predation or killing (gillnet, shooting or 
harvest) by humans, and which exhibit low reproductive rates 
and long lifespans, the senescence parameters are likely to 
be more important than for many other mammalian species. 
Siler (1979) relates a; and bz; from (2) and (5) to the modal 
age of senescence and the standard deviation around that age: 


az = exp(-T/S) (10) 


and 


b; 1/s (11) 
where T is the modal age of senescence and S is the standard 
deviation. The senescence parameters can thus be derived from 
estimates of T and S. Data from Schneider (1978) suggests 
that the modal age of senescence in Alaska otters may be 16 
years. Eberhardt (1985) found a significant correlation 
between T and S for 10 species of large mammals. A regression 
of the data presented by Eberhardt (1985) yielded: 


S = 0.161 + 0.144T ;(R°=0.84) (12) 
Using this equation and a modal age of senescence of 16, the 


standard deviation of age of senescence is 2.46, az = 0.0025 
and bz = 0.41. 


210 


Density-dependence.--Analysis of historical range 
expansion and population size data for the OTRANGE model, 
discussed elsewhere in this report, suggested values of 0.09, 
and 0.035, for Ypx and b, respectively, in (3). Given the 
present length of the range and USFWS (1986) estimates of 
substrate-specific carrying capacities, the model uses a 
default value of 1920 animals for K. 


Reconciliation of estimates.--A "spread-sheet" type 
program was set up to allow testing of adjustments in the 
parameter estimates. Equations (1) and (2) contain a large 
number of parameters relative to the number of "data points" 
available to use in estimating them; unstructured numerical 
solutions to (1) based on annual survival rates at only two 
ages undoubtedly would be degenerate. The "Shape" parameters 
(b,, bs, ba in (1) and (2)) are particularly difficult because 
of their abstractness and their potentially large effect on 
the survival and/or reproductive rates at certain ages. We 
thus structured our search for the best parameter estimates 
as follows: 


Siler (1979) describes the time constant at which 
maturity is approached (i.e., the rate at which the early risk 
becomes asymptotic) as 1/b;. In five mammal populations 1/b, 
was always less than one, with the value of b, ranging from 
1.06 to 3.84. We conservatively set b, equal to one for both 
males and females, giving a relatively slow approach to 
maturity, and emphasizing the apparently intense interspecific 
competition and reduced survival that young female sea otters 
experience. Similarly, b,, which governs the rate at which 
the prime reproductive rate is approached, was set equal to 
one. The relationship between the senescence parameters, bz 
and az, was held initially at the relationship suggested by 
Eberhardt (1985), and (12) above, and modal age of senescence 
was varied to affect the shape of the survivorship and 
reproductive curves through the adult ages. The prime 
reproductive rate was set at 0.25 and held constant to start 
with, and a;, az, and the modal age of senescence were varied 
to find estimates combinations of survivorship and 
reproductive schedules that: 1) were in reasonable agreement 
with estimated annual survival rates, 2) produced sex ratios 
of independent animals in the simulated population in 
reasonable agreement with those seen in the wild (i.e., 
female-biased), and, 3) provided reasonable survivorship 
schedules at all population growth rates between r,,, and - 
IYmxe ANnual survival rates were calculated from constructed 
survivorship schedules by : 


Sx xa, xt (13) 


where s, is annual survival rate at age x and 1, is 
survivorship to age x. 


211 


Following this approach, we arrived at the parameters 
given in Table 10.2; these are the default model parameters. 
With these parameters exp(-al) (early survival rate) in (1) 


Table 10.2. Default parameters used in OTPOP and LESLIE. 


Parameter Equation in text Value 

Maximum per capita Tey ee Tiea( 3) 0.085 
growth rate 

Non-linearity of density Voy “altey -(())) 0.020 
dependence 

Equilibrium population K in (3) 1720 
size 

Adult female risk a2 in (1) -1n(0.93) 
coefficient 

Female modal age of ud Ma Bo Wa OD) neS) 
senescence 

Standard deviation of female Sra meals) 2.46 
age of senescence 

Female senescence risk az; in (1) -1n(0.9977) 
coefficient 

"Shape" parameter for female b,; in (1) and (2) ait 
early hazards risk 

"Shape" parameter for female bz in (1) and (2) 0.41 
senescence risk 

Prime reproductive rate A in (2) 0.25 
"Shape" parameter for approach ba in (2) 1 
to prime reproduction 

Age before first reproduction Cuan; 1(2))s 3 

Adult male risk a2 in (1) 1n(0.87) 
coefficient 

Male modal age of T in (10) 9 
senescence 

Standard deviation of male Sian (asl) 3.5 
age of senescence 

Male senescence risk az in (1) -1n(0.9264) 
coefficient 

Male early hazards risk a, in (1) 0 
coefficient 

"Shape" parameter for male bz in (1) and (2) 1 


senescence risk 


for females varies from 0.96 when the population growth rate 
is 0.09 to 0.41 when the population growth rate is -Yyax; 
exp(-a2) (incidental survival rate) for females is 0.93, 
dropping to 0.91 when r = -Ypax, and increasing to 0.97 when 
Y = Ymx (Fig. 10.7). The age of first reproduction is lowered 
to three years when r > 0.06 (Fig. 10.8). The modal age of 
senescence for females remains constant at 15 years. The fit 
of the various estimates of female survival rates to the 
default model rates for r=0 is shown in Fig. 10.9. 


For males, the early hazards survival rate is set at one 
(a, =0, if a male survives weaning there are no additional 
hazards associated with youth, the value of b, for males is 
thus inconsequential), but the incidental survival rate (exp(- 
ao) is 0.87, and the modal age of senescence is nine years. 
A modal age of senescence of nine years implies, by (12), a 
standard deviation of 1.46 years around that age, values of 
az; and bz; calculated as such and combined with the previously 
determined rates for a, and ,2 gave a good fit to the two 
survival estimates, but led to there being no males over 12 
years of age in the population. This is obviously 
unrealistic, as four of the 219 male teeth were estimated to 
be 12 years or older. Adjusting the standard deviation of 
senescence upwards to 3.5 years led to a good fit of the 
annual survival estimates and ensured that male otters could 
survive until 16 years of age. The low modal age of 
senescence combined with a relatively large standard deviation 
results in the effects of senescence being manifest in the 
male survivorship curve at an early age (Fig. 10.10). MThis 
may seem anomalous, but if the concept of "risk of senescence" 
in males is stretched to include the risks associated with 
holding a breeding territory against younger animals, and if 
these risks include a mortality rate that increases with the 
years that an animal is not able to hold a territory, the 
curve may be biologically justified. 


With a population per capita growth rate (r) of zero, the 
predicted sex ratio in the independent population under this 
parameterization is 0.755 males per female (Fig. 10.11). When 
r =0.9 the predicted sex ratio is .572 males per female; when 
r=-0.09 the predicted sex ratio is 1.06 males per female. 


Using the survivorship schedules described above with 


r=0, the average annual survival rate of independent otters 
is calculated as: 


Sa (1yy/ lyy) ¥S, 
=0.844 (14) 


taking the 12th root yields an average monthly survival rate 
of 0.986. Using the relative pup survival rate calculated 


213 


FIGURE 10.7 -- Age-specific female survivorship curves and annual 
survival rates at different per capita growth rates, under the 
default population parameters used in OTPOP and LESLIE. 


or 
SURVIVORSHIP (¢ 


ANNUAL SURVIVAL (---) 


t 4 J 
Oo 1 304 so Geno) 9 10 11 12 13 14 15 16 17 18 19 20 21 22 25 


AGE 


214 


FIGURE 10.8 -- Age-specific reproductive rates under the default 
population parameters used in OTPOP and LESLIE. 


2 
Le) 
as 


rst eProduction at age 3 


0.08 


First reproduction at age 4 


Fi 


WEANED FEMALES PER FEMALE 


10 11 12 13 14 15 16 17 18 19 20 21 22 23 


AGE 


Py By Ch G7 tye) 


Oo 1 


2515) 


FIGURE 10.9 -- Age-specific female California sea otter annual 
survival rates calculated under default model parameters and a 
per capita growth rate=0 compared to survival rates estimated 
from field data. See text for explanation of estimates. 


ANNUAL SURVIVAL RATE 


0.9 Chapman—Robson estimate 
+/— 1 3.e.) 
(Applies at ages 4—10) 
0.8 - 
A Heisey—Fuller estimate 

0.7 (adults > 2 years old) 
0.6 | Heisey—Fuller estimate 

E (adults > 3 years old) 
0.5 - 
0.4 
0.3 
0.2 - 
0.1 


0 12 3 4 5 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 


AGE 


216 


FIGURE 10.10 -- Age-specific male California sea otter annual 
survival rates calculated under default model parameters and 
a per capita growth rate=0 compared to survival rates 
estimated from field data. See text for explanation of estimates. 


Chapmon—Robson estimate 
(+/— 1 s.e. 
(Applies at ages 4—10) 


q A Helsey—Fuller estimate 
(adults > 2 years old) 


0 Helsey—Fuller estimate 
(adults > 3 years old) 


ANNUAL SURVIVAL RATE 


0123 45 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 


AGE 


217 


FIGURE 10.11 -- Male and female California sea otter 
survivorship curves under the default population parameters 
used in OTPOP and LESLIE and a per capita growth rate=0. 


Female 


SURVIVORSHIP - 


012 3 4 5 67 8 9 1011 12 13 14 15 16 17 18 19 20 21 22 23 


AGE 


218 


above in (9), the absolute monthly pup survival rate is then 
0.986*0.898=0.885, and an annual rate (assuming six months 
dependency, six months gestation, and no  interuterine 
mortality) of 0.482. This is not too distant from the 0.50 
derived from the telemetry data when it is realized that the 
index areas, being areas of relatively high pup abundance, 
certainly contain a higher proportion of females than in the 
population as a whole, and the average adult survival rate as 
calculated above thus underestimates the average survival of 
adults in the index areas. By contrast, the average annual 
female survival rate (calculated as in (14) but ignoring 
males), is 0.87, and the annual pup survival rate calculated 
with that rate is 0.499. 


It should be reemphasized at this point that the 
parameter estimates supplied above are the model default 
values. They may be easily changed by the user to examine the 
effects of alternative parameterizations, or if future 
research allows refinement of the estimates. 


Seasonality.--The CDFG index area counts indicate a 
pronounced seasonality in the abundance of pups; this 
additional complexity in sea otter reproduction is taken into 
account in the model. The abundance of small pups in the 
index areas (Fig. 10.6) and the pooling across years (Fig. 
2.1, Chapter 2) indicated a peak of pup production in March 
and a low point in June or July; the biological basis for 
this seasonality is as yet unknown, it may be due to favorable 
pup-rearing conditions in the spring and summer leading 
directly to a synchronous breeding season, or it may be that 
pups born in the fall suffer high mortality and the females 
rebreed immediately. We model the seasonality descriptively, 
assuming a constant monthly pup mortality, and a constant 
monthly base rate of pupping, adding a seasonal pupping rate 
to give the observed spring peak. 


With a basic year round rate yielding 0.05 pups per 
independent as a starting point, and assuming that the small 
pup stage lasts three months the basic monthly rate is 
0.05/3.0 = 0.0167. An increased seasonal rate for December 
to April was then fit to the data in Fig. 10.6 by a numerical 
least squares procedure. This added increment was 0.02, 
making the monthly pupping rate for the December through April 
0.0167+0.020 = 0.0367. Since these rates were based on 
relative numbers of pups and independents the ratio of peak 
monthly pup production to basic monthly pup production, 
0.0367/0.0167 = 2.21 is the important parameter. The model 
uses this ratio and the age-specific pregnancy rates to 
determine the probability of conception in each month. The 
annual pregnancy rate is decomposed into the two relative 
pupping rates by numerically solving: 


(1-p) = (1-u,)’(1-u,)° (15) 


(i.e., the probability of not conceiving during the year is 
the product of the monthly probabilities of not conceiving) 
where p is the age-specific annual pregnancy rate, u,, is the 
basic monthly pupping rate, uz is the peak monthly pupping 
rate, and up = 2.21u,;. Intrauterine mortality is again assumed 
to be zero, and gestation is again assumed to be six months, 
so that the conception rate for November through May is u, and 
the conception rate for June through October is up). 


Age specific pregnancy rates for use in OTPOP are 
calculated according to: 


Px = 2m, / {1-(1-s,) —(1-Sp) +(1-S,) (1=So) } (16) 


where p, is the annual pregnancy rate for females of age x, 
m, is the reproductive rate from (2), s, is the annual survival 
rate for females of age x (as in (13)), and Sg is the annual 
pup survival rate. This calculation assumes an even sex ratio 
at birth and assumes that the number of animals weaned in a 
year (2m,) will be the number of pregnancies minus the number 
of females that die during the year (1-s,) minus the number of 
pups that die during the year (1-S)). Because pup death is 
not independent of maternal death (if a mother dies her pup 
dies also) the interaction term, (1=-s,) (1-S9), is added. The 
age specific pregnancy rates are used to calculate age 
specific monthly conception rates according to (15). The 
pattern of monthly pup production Generated by OTPOP appears 
aliny JaaIe?4 BLOBS MA 


In LESLIE maternal deaths are not considered in the 
calculation of pregnancy rates because in the life table 
formulation using an annual time step only the animals that 
survive the year are available to reproduce, so pregnancy 
rates are: 


Px = 2m,/So (17) 


Constructing an initial population 


The population growth rate for the first year of 
simulation is determined according to the density dependence 
function (3) given the initial population size and the 
carrying capacity supplied by the user. Survivorship 
schedules for each sex are calculated according to (1) using 
the population growth rate and the sex-specific risk 
parameters and senescence functions. The proportion of the 


220 


FIGURE 10.12 -- Pattern of monthly pup abundance produced by 
OTPOP. Crosses represent the mean of 85 years of simulation. 


Total pups 

170 
150 
Zz 
Lid 
Z 
4G 150 
Oo 
Ld 
OQ 
Z 110 Small pups 
oO 
oO 
O 30 
= 

N 
om 
LJ 
Qo 70 
io) 
oO 
2 50 Large pups 
N D J F M A M J J A S 0 N 
-MONTH 


population of each sex in each year class is then determined 
by: 


Vt Cha lana Camel, ;x=1,W, y=1,2 (18) 


where V,, is the proportion of the population of age x and sex 
ye 


Initial conditions require a distribution of ages in 
months within each age class, and an initial distribution of 
female reproductive state. Deterministic simulations 
convinced us that, given the relative conception rates 
described above, the distribution of reproductive status (and 
thus of month of birth) converges to a stable distribution 
from any initial distribution within a small number of 
iterations. This stable distribution depends on both the 
basic conception rate (u;) and the pup survival rate. In any 
run of OTPOP the pup survival rate is constant, but the 
conception rate varies by age according to (2) and (16). An 
average basic conception rate, weighted by the initial stable 
age distribution, is calculated; beginning with an initial 
uniform distribution of reproductive states the distribution 
is simulated deterministically for 15 years on a monthly 
basis. From the final (assumed stable) distribution of 
reproductive status a distribution of month of birth is 
extrapolated, and then these distributions are converted to 
cumulative density functions. 


The age in months of each independent animal in the 
initial population is determined by a random draw from the 
month of birth function. The number of pups and fetuses in 
the initial population is extrapolated from (18) given the 
initial population size and assuming a stable age 
distribution. The required number of pups and fetuses are 
distributed through the female age classes in proportion to 
the elements of the l,m,e™ vector, and the age of each pup or 
fetus within each female age class is determined by a random 
draw from the cumulative density function of reproductive 
status. Initial conditions generated in this manner alleviate 
the need for long runs of the model prior to introduction of 
an oil spill. 


OTDIST 
Structure 


OTDIST distributes the simulated population along the 
coast. It differs from the other submodels in the fact that 
it is not dynamic. In OTDIST we assume that the position of 
an individual otter along the coast is a function of the 
animal's sex and reproductive status, the month of the year, 


empirically derived population density functions, and 
empirically derived local sex ratios. 


We have digitized the coast and the associated five 
fathom ordinate system for use in the model (Appendix 10.1). 
For certain aspects of the model it is useful to consider the 
coast in discrete divisions; in these cases we have used the 
40 coastal segments delineated by the CDFG and USFWS in their 
carcass recovery efforts. 


Two density functions are used as input for OTDIST, one 
giving the density of otters at any point along the coast in 
May, the second giving the density in December. Similarly, 
functions giving the sex ratio in each of the 40 coastal 
segments in May and December are required input for the model. 
The first step in the distribution procedure is to determine 
local sex ratios and a density function for the desired month 
of the year by linear interpolation between the May and 
December functions. Then the density function and local sex 
ratios are combined to arrive at density functions for each 
sex by multiplying the density at each point along the coast 
by the proportion of each sex at that point. 


The female density function is then converted to a 
cumulative distribution function, and the program loops 
through the female age vector, determining each animal's 
position by generating a random number and finding a 
corresponding location in the cumulative function. 


After the females have been distributed OTDIST 
distributes territorial males. First potential territories 
are set up along the entire coast. Territory size is assumed 
to be normally distributed, the location of potential 
territories are determined by generating lengths from a normal 
distribution with empirically derived parameters, and stacking 
them along the coast. Coastal substrates are classified as 
either rock or sand, as determined from U.S.G.S. topographic 
maps; only rocky areas are allowed to be potential 
territories, as no territorial males have been observed in 
sandy areas in California (B. Hardy, pers. commn). 


Once the locations of potential territories are 
determined, the territorial status of each male in the 
population is determined. Males six years old and older are 
potentially territorial. The probability that a potentially 
territorial male will actually be on a territory is viewed as 
a function of month of the year, and determined as follows: 


P(t) = P(a) (1-P(d)) (19) 


where P(t) is the probability of being on a territory, P(a) 
if the probability of having arrived on a territory, and P(d) 


223 


is the probability of having departed a territory. P(a) and 
P(d) are assumed normally distributed with empirically derived 
parameters. The program then loops through the male 
population vector; when a male of 6 years or greater is 
encountered a random number is generated and compared to P(t) 
to determine territorial status. If the male is territorial 
the location of its territory is determined by generating 
another random number and finding a corresponding location in 
the cumulative distribution function for females. The length 
of coast encompassed by this territory is then made 
unavailable to non-territorial males by setting the male 
density function equal to 0 at all points within the 
territory. If a male draws a territory that is already 
occupied by a territorial male he is moved to the closest 
available unoccupied territory. 


The last step in the program is to distribute the non- 
territorial males. The male density function, as modified by 
territoriality, is converted into a cumulative density 
function. Another loop through the male population vector is 
executed; if a male does not have a territory a random number 
is generated and the corresponding location in the male 
cumulative distribution function is found. 


Once the positions of all animals in the population have 
been determined, the population and location vectors are 
passed to OTMOVE, the short term movement and oil response 
model. The locations of males occupying territories are 
flagged as they are passed to OTMOVE. 


Parameterization of OTDIST 


Biologically, the spatial distribution of animals ina 
population can be viewed as a function of the distribution of 
resources and social interactions amongst the individuals in 
the population. Our understanding of the actual mechanisms 
that produce an observed distribution of animals from the 
underlying distribution of resources and social system is very 
incomplete, so we must be content with modeling the 
distribution in an essentially descriptive manner, 
incorporating few mechanisms. OTDIST, the distribution model, 
has been structured to utilize what data is available on the 
distribution of sea otters within their range in California. 


Density functions.--Required inputs for OTDIST include 
density vectors for winter and summer, representing the number 
of animals in each 500m segment of the five fathom ordinate 
system in November and May, respectively. CDFG and USFWS 
census data were used to construct these vectors. Dates and 
methods of the censuses for which we had access to the raw 
data are given in Table 10.3, census methods are described 
Wendell, et al., (1986). 


Interpretation of the census data suffered from the 
questions of range expansion mentioned at the beginning of the 
report. For our analysis the locations of animals recorded 
on field maps during the censuses were translated to five 
fathom line ordinates and summed by census and 500m segment. 
Contiguous blocks of 20 500m segments, representing 


Table 10.3. CDFG and USFWS censuses used in analysis of 
California sea otter distribution. 


Date Average 
Year Month Total Count Group Size Method 


1968 Aug 311 5.27 Air 
1968 Nov 659 4.92 Air 
1968 Dec 409 3.56 Air 
1969 Jan 986 5.94 Air 
1969 Feb 685 3.26 Air 
1969 Mar 942 4.34 Air 
1969 Apr 654 5.03 Air 
1969 May 315 4.32 Air 
1969 Jun 1013 5.30 Air 
1969 Aug 528 3.74 Air 
1969 Sep 404 3}, Sal, Air 
1969 Oct 485 3.13 Air 
1970 May 902 5.43 Air 
1970 Sep 607 4.40 Air 
1971 Feb 719 3.95 Air 
1971 Apr 901 3} oa Air 
1971 Jul 957 4.65 Air 
1971 Oct 712 4.07 Air 
1972 Jan 1064 3.81 Air 
1972 Apr 772 2.81 Air 
1973 Dec 936 3.38 Air 
1974 Mar 956 2.14 Air 
1975 Jun 1040 2.30 Air 
1976 Jun 1148 2.12 Air 
1979 Jun 808 2.32 Air 
1982 Nov 1334 akg Ia Ground 
1983 Oct 1222 2.06 Ground 
1984 Jun 1567 2.32 Ground 
1985 May/Jun 1287 2.14 Ground 
1985 Oct/Nov 1212 2.20 Ground 


10 km of coast each, were grouped together and totaled by 
census. The percent of the total count of each census in each 
block was then calculated and used in an analysis of variance. 
Since sea otter distribution is generally considered to change 
on a seasonal basis (USFWS 1986) we grouped the census within 


225 


each year by season: winter consisting of November through 
April, summer consisting of May through October (data were too 
sparse to attempt analysis by month). ANOVA (Table 10.4) 
showed significant main effects of both year and location, and 
of season within year. There was also significant interaction 
between year or season and location. The situation is 
represented graphically in Figs 10.13 and 10.14. 


SS ____eese__......_._.______ Ee 


Table 10.4 -- Analysis of variance in CDFG and USFWS 
California sea otter census data, 1968-1985. Dependent 
variable is the proportion of census total along a 10 km 
section of coast. 


Source d.f ss F 

Year 13 -06065 3.98 0.0001 
Season within Year 6 -01985 2.82 0.0100 
Location 40 - 41554 8.87 0.0001 
Year * Location 362 - 18867 0.45 0.9900 
Season & Location 127 »10161 0.68 0.9900 
Error 240 ~28108 

Total Model 548 - 78632 1.23 0.0300 


Model R=0.737 


Variation in distribution due to location and season are 
easily handled by the structure of OTDIST, but, as mentioned 
previously, OTDIST is time-independent, and annual changes in 
distribution (aside, of course, from pure stochastic effects) 
are not explicitly considered. Distribution of sea otters 
throughout the range in California has undoubtedly changed 
since the censuses began. We thus decided that the best 
parameterization of the density functions would be direct 
incorporation of the most recent census data. A separate 
analysis showed a highly significant effect of census method 
on both total count and on average group size. Ground counts 
appear to enumerate a greater proportion of the population, 
particularly that part of the population that is solitary on 
the day(s) of the census. The relative worth of ground counts 
versus ground-truthed aerial counts for estimating population 
size has been the subject of some debate, but for present 
purposes, that of determining relative distribution, it seems 
as if the method that enumerates the higher proportion of the 
population will better estimate relative densities. This 
contributed to our decision to use only the most recent 
censuses, aS coordinated ground counts did not begin until 
1982. 


Census data from 1984 and 1985 were converted to 


probability density functions for each census by dividing each 
segment total by the total number of animals recorded in the 


226 


FIGURE 10.13 -- Contour diagram indicating annual changes in 
sea otter density in California from 1968-1985. Y-axis 
represents space (5 fathom line ordinate system, increasing, 
generally from North to South, in 500 meter increments along 
the five fathom line), x-axis represents time. Diagram was 
constructed for CDFG and USFWS census data (Table 3). 
Individual census counts were totaled by 20km section of coast, 
and the proportion of the total count in each section 
calculated. These proportions were then averaged by year to 
get the values used to produce the diagram. Contour interval 
is 3% of the individual census total. Expansion of the range 
is evident at the north and south ends of the diagram. 


200 


LOCATION ALONG 5 FATHOM LINE 


YEAR 


227) 


FIGURE 10.14 -- Contour diagram indicating monthly changes in 
sea otter density in California from 1968-1985. Y-axis 
represents space (5 fathom line ordinate system, increasing, 
generally from North to South, in 500 meter increments along 
the five fathom line), x-axis represents time. Diagram was 
constructed from CDFG and USFWS census data (Table 3). 
Individual census counts were totaled by 20km section of coast, 
and the proportion of the total count in each section 
calculated. These proportions were then averaged by month to 
get the values used to produce the diagram. Contour interval 
is 3% of the individual census total. Seasonal contraction of 
the range, presumably due to migrations by males, is evident 
at the north and south ends of the diagram between February and 


July. 


LOCATION ALONG 5 FATHOM LINE 


228 


respective census. These density functions were then averaged 
by season to get the empirical distributions that are used in 
the model (Fig. 10.15). Approximately 30% of the range is 
inaccessible by road, and thus is counted by air, even during 
the "ground" counts. Geibel and Miller (1984), Wendell et al 
(1986), and Hardy (CDFG, personal communication) report that 
aerial observers typically enumerate from 50-80% of the 
animals seen by ground truth observers; the density functions 
derived from the raw census data are thus further modified by 
multiplying the densities in portions of the range that are 
counted from the air are multiplied by 1.3. 


It is assumed that differences between the two seasonal 
distributions are due to seasonal movements of animals, 
particularly adult males migrating between male areas in the 
winter and breeding territories in the female areas in the 
summer. The censuses are timed to reflect the peak of 
congregation in the male areas and the peak of territoriality, 
respectively; therefore density functions for each month are 
obtained by linear interpolation between the summer and winter 
functions. 


The censuses are an integral part of CDFG's and USFWS's 
sea otter research programs, and are scheduled to continue to 
be conducted twice yearly in the future. The density 
functions are stored in an external file so that they can be 
easily updated after each census and incorporated into the 
model. USFWS has been supplied with a computer program that 
allows rapid digitizing of raw census data and outputs the 
data in an appropriate form for the model. 


Sex ratios.--The census data provide distributions for 
the population as a whole, but provide very little information 
about the distribution of each sex throughout the range. Some 
local sex ratio information is available from the carcass 
recovery data. The location of each recovered carcass is 
recorded by recovery area (Ames et al 1985). Each recovery 
area is about 12km long. Sex ratios for each recovery area 
were calculated, combining all data from December through 
April and from May through November (Table 10.5). 


The carcass recovery data suffers from small sample size 
in many of the areas, from sampling problems discussed by Ames 
et al (1984), and, again, from changes in distribution that 
have occurred as the range expanded. As an alternative we 
used a delphic technique, asking field biologists from CDFG, 
USFWS, and other institutions who had been frequent 
participants in the semi-annual censuses to estimate present 
seasonal sex ratios in each of the recovery areas. Averaging 
the responses from the questionnaire gave the sex ratios in 
Table 10.5 and Fig. 10.16. Each of the respondents was 
familiar with the carcass recovery data, and most said that 


229 


FIGURE 10.15 Density functions used in OTDIST for the 
location of independent sea otters in California in June 
(dashed line) and December (solid line). Functions derived 
from CDFG and USFWS censuses conducted in 1985. OTDIST works 
at a resolution of 50m, but for clarity densities were totaled 
by 20km sections to construct this diagram. The numerical 
value of the locations along the five fathom line generally 
increases as one moves to the south along the California Coast. 


ry 


0.12 ra 
Zz 1\ 
O i t 
~ 53 
< 0.10 - i) ! \ 
! 1 
3) 2 \ iN i \ 
3 Teoog einer 
0.08 \ \ 
ral I | i \ [= : \ 
Le i { ! ; /\ 
© 0.06 - es i visleoaty 
FL, \ i \ \ 
Oo Se \ 
~ { ‘ 
be 0.04 - | | x 
oO A \ 
QO. 7N > \ 
\ > S 
o 0.02 + /,” . 7 § 2 
ra / \i/ E 
y \f 2 = 
0.00 


LOCATION ALONG THE 5 FATHOM LINE 


23C. 


\ 


=~ 
~ 


220 260 300 340 380 420 460 500 540 580 620 660 700 740 780 820 860 900 940 


FIGURE 10.16 -- Local proportion of California sea otters that 
are female in June (dashed line) and December (solid line), 
used in OTDIST. Data, from questionnaires distributed to 
knowledgeable biologits, was collected by CDFG carcass recovery 
area and assumed to apply over the entire section of coast 
within each recovery area. The numerical value of the 
locations along the five fathom line generally increases as one 
moves to the south along the California Coast. 


0.8 


Lif 
—! 0.6 
=e 
= 
Ld 
Le 
Zz 
O 0.4 
kK 
or 
O 
oO 
© 
o. 0.2 
o 
s 
o 
= 
0.0 


200 300 400 500 600 700 800 900 


LOCATION ALONG THE 5 FATHOM LINE 


Zoi 


1000 


they used that data as a guideline in making their estimates. 
We feel that their estimates are more realistic than those 
that could have been attained by a purely statistical analysis 
of the sparse data, and therefore used their estimates in the 
model. 


Table 10.5. California sea otter sex ratios of recovered 
carcasses (1968-1985) and as subjectively estimated by field 
biologists, by season and CDFG carcass recovery area. 


CDFG Recovered carcasses Subjective estimates 
recovery May-Oct Nov-Apr May-Oct Nov-Apr 
area® M F M F M/F M/F. 
<7 3 (0) 1 0) 4.5 16.7 
7 1 0 3 3 4.5 16.7 
8 1 0 5 0 4.5 16.7 
9 1 0 0) 3 4.5 16.7 
10 2 0) 2 1 4.5 16.7 
11 42 8 27 6 4.5 16.7 
12 26 24 6 1 1.0 3.0 
13 36 24 35 30 0.5 0.9 
14 20 20 22 13 0.5 0.9 
15 2 6 4 5 0.5 0.9 
16 16 11 13 23 0.3 0.3 
17 2 5 9 16 0.3 0.3 
18 11 30 25 48 0.4 0.4 
19 3 17 14 16 0.4 0.4 
20 1 7 7 13 0.5 0.4 
21 1 5 1 1 0.5 0.4 
22 2 4 2 4 0.5 0.4 
23 (0) 0 2 5 0.5 0.4 
24 (0) 1 1 (0) 0.5 0.4 
25 0 1 5 1 0.5 0.4 
26 (0) 1 0 0 0.4 0.3 
27 3 10 5 6 0.4 0.3 
28 16 10 15 17 0.4 0.3 
29 18 31 12 30 0.3 0.3 
30 10 6 8 2 0.3 0.3 
31 16 9 5 7 0.8 0.6 
32 70 36 27 18 0.8 0.6 
33 25 16 25 12 0.8 0.6 
34 2 2 2 2 1.0 3.0 
35 5 1 2 0 1.0 3.0 
36 22 2 5 3 4.5 16.7 
37 33 7 12 5 4.5 16.7 
38 4 1 2 0 4.5 16.7 
39 0 0 1 0 4.5 16.7 
40 2 0 1 2 5.0 17.0 
See Appendix 1 for location of CDFG carcass recovery areas. 


232 


As for the seasonal distributions, the seasonal sex 
ratios are assumed to be the result of male migrations to and 
from breeding territories, and sex ratios for the each month 
are obtained by linear interpolation. Probability density 
functions for each sex in a given month are then obtained by 
multiplying the population density function for the 
appropriate month by the appropriate sex ratio: 


Sijk = AjRPijx (20) 


where s;;, is the density of sex i, in month j in 500m segment 
k, Gj, is the population density in month j in segment k, and 
Pijk 1S the proportion of sex i in month j in segment k, and 
sex ratio in a recovery area is assumed to apply to all 500m 
segments in that recovery area (Fig. 10.17). 


Parameters pertaining to territoriality.--The model divides 
the coast into potential territories. Territory length is 
assumed normally distributed, parameters of the distribution 
were obtained from Jameson (1987), who reported mean male 
territory length near Piedras Blancas during 1978-1984 at 1.1 
km (s.d. =0.43km, N=13), and our telemetry data. Additional 
parameters required by the model are mean territory arrival 
and departure dates, and associated standard deviations. 
Jameson (1987) gives mean arrival date as 22 May (s.d. =33.6 
days, N =16) and mean departure date as 21 December (s.d. 
=38.1 days, N =18). The proportion of males over the age of 
six that are territorial at any given time is assumed to vary 
seasonally between 0.25 and 0.75. 


The data from Jameson (1987) provides default values for 
the male territoriality parameters in OTDIST; but our 
telemetry data suggests that the highly seasonal pattern of 
male movements observed by Jameson (1987) during 1978-1984 may 
not be occurring at the present time or throughout the entire 
range (see Chapter 3). The user can change the values of the 
territoriality parameters at the beginning of a model run; 
setting large standard deviations of arrival and departure 
dates and/or small differences between the minimum and maximum 
proportions of males that are territorial will reduce the 
amount of seasonal variation in male territoriality in the 
model. 


Expanded sea otter range 


As mentioned previously, the size of the sea otter range 
and its carrying capacity are fixed for the duration of any 
run of the model. Since, however, the peripheral areas of the 
current range are the most susceptible to oil spills, and it 
is very possible that the range will continue to expand in the 


233 


FIGURE 10.17a -- Density functions used in OTDIST for the 
location of male (dashed line) and female (solid line) sea 
otters in California in June. The numerical value of the 
locations along the five fathom line generally increases as one 
moves to the south along the California Coast. 


0.14 
Zz 012 
2) 
K- 
yore 
=) 
au 
O “\ \ 
OL 0.08 7 \ " , I 

atee at NV 1\ \ f 
O 7\ | \ F \ a \ ! . 
=> 9.06 ee F \ ! Ny \\ / _ 
O J \ I \ 1 \\ / \ 
f= Sys: Lopdl! sl 7 \ 
/ 
004-4, i: Se " 
O I 7 . : 
QO. U \ > [e} \ 
O 0 oO : 
(v 0.02 - Vig ° a 
ij = E a 

i. WE 8 

0.00 


220 260 300 340 380 420 460 500 540 580 620 660 700 740 780 820 860 900 940 


LOCATION ALONG THE 5S FATHOM LINE 


234 


FIGURE 10.17b -- Density functions used in OTDIST for the 
location of male (dashed line) and female (solid line) sea 
otters in California in December. The numerical value of the 
locations along the five fathom line generally increases as one 
moves to the south along the California Coast. 


0.12 


PROPORTION OF POPULATION 


> 
2 
o 
ra 
: 
6 
= 


0.00 = 
220 260 300 340 380 420 460 500 540 580 620 660 700 740 780 820 860 300 3940 


LOCATION ALONG THE 5 FATHOM LINE 


2335) 


near future, we have constructed a simple deterministic model 
of sea otter range expansion, OTRANGE, which was described 
earlier. The output from OTRANGE includes projected 
population sizes, range boundaries, and carrying capacities, 
which can be used to initialize OTPOP. Obviously, the census 
data required as input for OTDIST will not be available for 
the peripheral parts of the simulated range as expanded by 
OTRANGE. Thus we provide a sub-routine, EXPRAN, that, in the 
event that the north and south ends of the census data used 
as input do not coincide with the range boundaries input by 
the user, will predict the distribution of otters in the 
expanded range. 


After examination of the data in Figs. 10.13,10.14,10.15, 
and 10.16, we arbitrarily divided the range into a southern 
periphery, south of the 840 ordinate on the five fathom line, 
a northern periphery, north of the 340 ordinate on the five 
fathom line, and a center, between the 340 and 840 ordinates. 
Using the density functions constructed by OTDIST from the 
census data, EXPRAN calculates the average density of otters 
of each sex in the center of the range (Fig. 10.18, point A), 
and assuming that the densities at the endpoints of the 
censused range are zero, calculates the average slope of the 
distribution through the north and south peripheries (Fig. 
10.18, slope S). Density is then extrapolated from the end 
of the central range into the peripheral range a distance 
equal to the length of expansion using the central range 
average density combined with censused deviations from the 


slope of the peripheral density (Fig. 10.18, line A-B). In 
this manner the central range is thus considered to have 
extended into existing peripheral range. Densities in the 


peripheral part of the range beyond the expanded central part 
of the range are calculated by adding the difference in 
density implied by the slope of the peripheral density and the 
length of expansion (Fig. 10.18, line B=-C). Density in the 
expanded part of the range is made cumulatively equal to the 
area of CDE in Fig. 10.18, but weighted at any point according 
to the type of substrate. USFWS (1986) estimates that rocky 
habitats can support 3.1 times as many otters as_ sand 
habitats, this ratio is used in the determining the densities 
in the expanded range. The heavy line in Fig. 10.18 indicates 
the new densities calculated by EXPRAN. 


OTMOVE 
Structure 

OTMOVE simulates the movements of the animals in the 
population on a daily basis for up to 30 days, and checks 
animal positions against the location of a simulated oil 


spill. It iterates on a daily time step, and considers 


236 


FIGURE 10.18 -- Schematic representation of the algorithm 
used in EXPRAN to predict sea otter densities in expanded 
range. See text for explanation. 


RELATIVE 
OTTER DENSITY 


Existing: (Expanded Range periphery Range center 
Range 


a ip 


Range periphery Range center 


POSITION ALONG COAST 


Expanded: 


237 


position at a spatial resolution of 50 meters. The location 
vectors generated by OTDIST are the initial positions of the 
animals, the identity of animals dying as a result of contact 
with the spill are passed to OTPOP. 


As in OTDIST, spatial considerations are simplified by 
conceptualizing the system as one dimensional. Otters are 
located and move upcoast and downcoast on the five fathom line 
ordinate system, and oil spills are 1 dimensional also. At 
the beginning of a run the date, duration (in days), and 
boundaries of the oil spill are input. Since the purpose of 
the movement model is to determine the numbers of animals that 
die as a result of the spill, it runs for only as many days 
as the duration of the spill. The movements of each animal 
are assumed independent of the movements (but not the spatial 
distribution) of the other animals in the population. This 
allows a structural efficiency of looping days within animals 
rather than animals within days; the movements of each animal 
are simulated for the duration of the spill and its fate 
decided before the next animal is considered. 


Each otter in the simulated population is considered to 
have a home range (or a territory, for territorial males), the 
center of which is the position assigned to the animal in 
OTDIST. OTDIST assigns each animal to one of six classes 
depending on sex and reproductive status: 1) juvenile male, 
2) adult, non-territorial male, 3) adult territorial male, 4) 
juvenile female, 5) adult female without pup, 6) adult female 
with pup. Juveniles are animals younger than the age of 
sexual maturity that is used in the population model, males 
over the age of six years are potentially territorial. Three 
categories of daily movements are considered in the model: 1) 
"routine" movements around the home range or territory center, 
2) seasonal migrations by territorial males, and 3) movements 
in response to oil spills. The position of an animal at 
the end of a simulated day is calculated as: 


Xp = Xe-1 + At (21) 


where t indexes days, X is the position along the five fathom 
line, and d, is the daily movement. Negative values of d, 
indicate movement up the coast (i.e., toward the origin of the 
five fathom ordinate system), positive values indicate 
movement down the coast. 


Routine daily movements.--Routine daily movements are 
modeled as a function of displacement from the home range or 
territory center, and the magnitude and direction of the 
previous day's movement: 


A, = bydy.y+b2(Xp—-C) +2, (22) 


238 


where b, is’ the autoregressive parameter, bz is the 
displacement parameter, C is the location of the home range 
or territory center, and Z is a normally distributed random 
error with mean 0. The parameters b, and bz and the standard 
deviation of Z vary with class. 


Migratory movements.--Migratory movements by adult males 
are simulated at appropriate times of the year. Territory 
arrival and departure dates are assumed to be normally 
distributed, empirically derived means and standard deviations 
around those dates are used to calculate the probability of 
a male arriving or departing a territory on each day of the 
simulation. For territorial males on each day of simulation 
a random number is compared with the probability of departing 
a territory in that day For non-territorial but potentially 
territorial males a random number is compared against the 
probability of arriving on a territory; to account for travel 
time to the territory, the mean of the probability 
distribution is set three days before the actual mean; thus 
the distribution gives the probability of non-territorial male 
departing for a territory. 


Class two (non-territorial adult) males that are determined 
to depart their present home range for a territory are 
assigned destination territories using the cumulative 
distribution of female positions derived from the female 
location vector constructed by OTDIST. Since territorial male 
density is thought to be negatively correlated with pup 
density (Jameson 1987, USFWS 1987), only the locations of 
mature females without pups are used to construct the 
distribution. The destination is compared to the list of 
potential territories and their statuses (occupied or not 
occupied) (also generated by OTDIST). If the territory 
originally assigned as a destination is occupied, the closest 
(to the original destination) available territory becomes the 
destination territory. Once a destination territory has been 
determined, the male moves according to: 


Ap = i* |ViaxtZe | (23) 


where Z, is as in (22), Vmx is the maximum daily rate of 
movement for a class 2 animal and i = +1 if the destination 
is down the coast from present position and -1 if the 
destination is up the coast. Once the male has reached the 
destination territory it moves routinely according to class 
three parameters. 


Class three animals that are determined to leave their 
territories are assigned destination home ranges by choosing 
from a cumulative density function constructed from the 
locations of males generated in OTDIST. Simulated male 
density in each 500m segment is squared before constructing 


239 


the cumulative density function in order to accentuate the 
aggregation of non-territorial males. Once a destination has 
been chosen the male moves according to (23) until he reaches 
the territory, at which point he moves routinely according to 
(22) with class two parameters. 


Movements in response to oil spill.--Each time that a daily 
movement can bring an otter into contact with the oil spill 
a series of "decisions" on the part of the otter are 
simulated, conditional on the spatial relationship between the 
animal's home range, its present location, and the oil spill. 
If the animal's home range center is inside the spill 
boundaries it may elect to abandon its home range and 
establish a new range outside of the spill with daily 
probability PE. If the animal's home range center is not 
within the spill boundaries, or if it is within the spill 
boundaries but the animal has elected not to abandon the home 
range, it may attempt to avoid the spill with probability PA. 
If an animal avoiding a spill with a present location outside 
of the spill "bounces" off of the spill boundary it moves a 
distance: 


Br = -(d,-D,;) (24) 


where d,; is the predicted daily movement according to (22), 
B, is the distance bounced, and D, is the distance to the oil 
spill boundary. If an animal elects to attempt to avoid the 
oil after it is already inside of the spill boundaries it 
moves according to (23) with the value of V,,, and the standard 
deviation of Z appropriate to its class, and the sign of i is 
randomly assigned with equal probability. © 


If, at the end of a simulated day, an animal is inside of 
the spill boundaries (and thus exposed to oil), it dies with 
probability PM. 


Assuming for a moment that PM is very low, a number of 
behavior patterns in relation to oil on the part of individual 
animals may occur depending on the values of PE and PA and the 
size of the spill. Animals might depart the spill area 
immediately and not return during the life of the spill; 
animals might spend a few days in the spill and then depart; 
animals might move routinely outside of the spill; avoiding 
it by bouncing off when routine movements would ordinarily 
bring them inside the spill; animals might continue to move 
routinely on the edge of the spill entering it occasionally; 
animals might move in a routine manner within the spill; 
animals might move long distances up and down the coast inside 
the spill in a "panic". 


240 


Parameterization of OTMOVE 


Parameters used in the movement model were derived from our 
telemetry data. The daily locations of radioed animals, 
recorded in the field on an x,y coordinate system, were moved 
to the five fathom line ordinate system to simplify the 
analysis of movement patterns and to derive a parameterization 
applicable to the single spatial dimension used in the model. 
Graphic traces of the movements of each animal along the five 
fathom line are presented in Chapter 3. Two of these are 
reproduced here for illustrative purposes (Fig. 10.19). 


Routine movements.--After original examination of the 
traces represented by Fig. 10.19 we attempted to analyze the 
movements of each animal as an autoregressive time series, 
modeling each day's movement as a function of previous days' 
movements and/or correlated error terms. This analysis led 
to good predictive equations for the movements of many of the 
animals, but the equations for individual animals often 
differed dramatically in form and degree. With no biological 
basis on which to decide upon the efficacy of one form of 
equation over another, incorporation of these equations into 
the movement model was unjustifiably complicated. 


We thus opted for the simple regression equation (22) 
mentioned previously. One autoregressive parameter is 
maintained in the equation, but the major factor in the 
equation is the displacement term. The horizontal lines in 
the traces of Fig. 10.19 mark the mean position over all days 
for which data was obtained on each animal. Movements of many 
animals were characterized by long periods of time above or 
below the overall mean, but localized oscillations around 
short term means (Fig. 10.20). Since the movement model is 
designed to run for at most 30 days, we ignore patterns that 
occur on a longer time scale, and model the oscillation around 
short term means. The series of locations for each animal 
were divided into arbitrary non-overlapping 30 day segments, 
and regression parameters calculated for each segment 
(segments wherein an animal was not able to be located for 
more than 10 of the 30 days were not included, a total of 383 
segments were used in the analysis). This parameterization 
was encouraging, as the displacement terms in all equations 
were negative, and most were highly significant. We then 
restratified the analysis, grouping the segments according to 
the six classes of animals described earlier (30 day segments 
in which the reproductive status of an adult female changed 
or was unknown were discarded). The regression parameters and 
the standard deviation of the errors are given in Table 10.6; 
these are the default values used in the model. 


241 


FIGURE 10.19a -- Daily locations of a juvenile female 
California sea otter (#35) as determined by telemetry, 1985- 
1986. Location is given in 50m units south of San Francisco, 
along the 5 fathom line ordinate system. Horizontal line is 
the mean position of all daily locations. Julian date 1 is 1 
January 1984. The numerical value of the locations along the 
five fathom line generally increases as one moves to the south 
along the California Coast. 


4000 


qn 
o 
c—) 
c—} 


6000 


mam Bomar ns wi NQzZor>zr =zonaTSoor 
— 
oe 
o 
eo 


8000 
400 500 600 700 800 900 1000 1100 
JULTAN DATE 


242 


FIGURE 10.19b -- Daily locations of a juvenile female 
California sea otter (#29) as determined by telemetry, 1985- 
1986. Location is given in 50m units south of San Francisco, 
along the’ 5 fathom line ordinate system. Horizontal line is 
the mean position of all daily locations. Julian date 1 is 1 
January 1984. The numerical value of the locations along the 
five fathom line generally increases as one moves to the south 
along the California Coast. 


erie 
en 


ao 
wo 
[=] 
Oo 


~~ 
Ss 
oS 
— 


~ 
_ 
ao 
o 


71200 


MSHRr- ZomtHAS sy on Qzor> SOM H SOOT” 


7300 


7400 


7500 
600 700 800 900 1000 
JULIAN DATE 


243 


Leora, 
i 


1100 


FIGURE 10.20a -- Daily locations of juvenile female California 
sea otter #35, as in Fig. 10.19a. Horizontal lines are the 
mean positions during arbitrary 30-day segments. The 
numerical value of the locations along the five fathom line 
generally increases as one moves to the south along the 
California Coast. 


6300 


~r 
_ 
a 
[—J 


an 
on 
(—J 
[—] 


an 
x 
c—J 
(—J 


maser ZrortHere nn] on Qzo~> =zZorH> Oor” 
arn 
3 3 
= = 


6900 


600 700 800 900 1000 
JULTAN DATE 


244 


1100 


FIGURE 10.20b -- Daily locations of juvenile male California 
sea otter #29, as in Fig. 10.19b. Horizontal lines are the 
mean positions during arbitrary 30-day segments. The 
numerical value of the locations along the five fathom line 


generally increases as one moves to the south along the 
California Coast. 


: AY | 
Nl \ i" 


on 
a 
o 
o 
—_ 
[— 
| 
a 


M=zHR- SOEASTN HW QZFOory BONA SOOM 
(2) 
w 
o 
o 


6000 


600 700 800 900 1000 
JULIAN DATE 


245 


1100 


Vmax, the maximum daily rate of travel was estimated for 
each class of animal by considering the maximum distance 
between locations taken at least 24 hours apart for each 
animal, and calculating maximum net daily movement: 


Vmax = MAX{ | (dx/dp) |*24 } (25) 


where d, is the distance between two successive locations and 
d, is the time, in hours, between the locations. If an 
animal's class changed during the study periods of different 
classes were considered separately. These values were then 
grouped and averaged by class to get the values in Table 10.6. 
Since the greatest values of V,,, are seen in juveniles 
(classes one and four), and it is unlikely that an otter's 
swimming speed decreases in adulthood, V,,, for classes one- 
three is set equal to 48.6 500m segments/day in the model, and 
Vmax f£0r Classes four and five is set equal to 37.5 500m 
segments/day. Pups likely restrict female movements, SO Vmax 
for class six animals remains at 8.4 500m segments/day in the 
model. 


ee —_.._._ 


Table 10.6. Parameters used in short-term otter movement 
model. AR and CE are regression parameters discussed in text. 
sd is standard deviation of regression errors, R° given for 
regressions. Vmax is mean maximum daily movement, derivation 
discussed in text. 


Class Status AR CE sd R? Vmax 
1 Juvenile male -0.045 -0.290 8.56 (0.13) 48.9 


2 Adult non-territorial 
male 0.105 -0.815 4.64 (0.37) 40.2 


3 Adult territorial 
male 0.042 -1.044 1.93 (0.43) 36.2 


4 Juvenile female 
0.367 -0.163 8.09 (0.10) Bi D 


5 Adult female w/o 
pup -0.025 -0.406 6.39 (0.21) 20.0 


6 Adult female 
w/pup -0.009 -0.706 2.95 (0.38) 8.4 


Migratory movements by adult males.--The probabilities 
of migratory movements for males of class two and three are 
calculated using the data from Jameson (1987), as in the 
distribution model. Probabilities are calculated on a daily 
basis assuming a normal distribution of arrival and departure 


246 


times. Since Jameson (1987) gives mean date of territory 
arrival, calculation of the mean date upon which males leave 
for their territories requires consideration of the transit 
time to the territory. We have arbitrarily set that at three 
days, so that the mean date of departure for a territory is 
19 May. 


The dates given by Jameson are far enough apart that 
there is no overlap in simulated territory arrival and 
departure. The same caveats about the seasonality in male 
migratory movements that were mentioned in the discussion of 
OTDIST apply here. The same values of the territorial 
parameters that are set for OTDIST are used in OTMOVE; 
seasonality less pronounced than that described by Jameson 
(1987) and as indicated by our telemetry data, can be 
simulated by setting large standard deviations of arrival and 
departure dates. 


Movements in response to oil spill.--The parameters PM, 
PA, and PE are delphic parameters; very little data from which 
to estimate their values are available. Costa and Kooyman 
(1982) found that otters oiled over 25% of their surface will 
die of hypothermia if not cleaned, suggesting that PM could 
be very high. Ford and Bonnell (1986) use values of 30%-90% 
as most likely mortality rates in their simulations, depending 
on the condition of the oil, but allowed the possibility of 
mortality varying between 10 and 100%. Siniff, et al., (1982) 
found that captive sea otters did not avoid areas of the 
holding tanks experimentally contaminated with oil, suggesting 
that PA, the probability of localized movements to avoid the 
spill, may be very low. It is also likely that PE, the 
probability of leaving the spill area to establish a new home 
range, is a good deal smaller than PA. 


The values of these parameters are set by the user at 
runtime, facilitating evaluation of the relative importance 
of these parameters in determining the amount of mortality 
from a spill within the structure of the model. Additionally, 
the values of these parameters are set independently for each 
day of the spill, allowing consideration of the effects of 
weathering on oil (i.e., PM decreasing with time), or possibly 
learning on the part of the animals (i.e., PA increasing with 
time), or other scenarios. We anticipate that much of the 
model's usefulness will be due to its ability to simulate 
different oil spill response scenarios. 


Fig. 10.21 traces simulated movements of otters in the 
vicinity of an oil spill presumed to occur on December 1, and 
lasting 15 days, at the southern end of the Monterey Bay and 
eastern side of the Monterey Peninsula. The rangewide 
population in these simulations was set at 1600 animals. In 
Fig. 10.21la all delphic parameters are set to 0.0, that is, 


247 


the spill has no effect on the animals' behavior, and the 
movements thus reflect a "normal" situation. In Fig 10.21b 
PM and PA were set to 0.8 for the duration of the spill, and 
PE was set to 0.5 for the duration; in Fig. 10.21c PM was set 
to 0.1 for the duration, PA was set to 0.5 for the duration, 
and PE was set to 0 for the duration. The same initial 
population and distribution, and the same random number seeds, 
were used in all 3 simulations. 


MODEL OUTPUT 


A log file, recording the user-input parameter settings, 
is generated each time the model is run. Six files of raw 
output data are written as the model runs. One contains the 
Simulated population sizes for control runs (runs without 
introduction of oil spills), and one contains the simulated 
population sizes for runs with oil spills. In each of these 
the numbers of males, females, and pups are recorded once a 
year, at the end of the month in which the oil spill occurs. 
A third file records the numbers of oil spill-caused deaths 
by class ( juvenile male, adult non-territorial male, etc.) 
and by day of spill. The fourth file records the total number 
of deaths of males, females, and pups due to the spill in each 
run. The fifth file records the total population size just 
prior to the spill, and the number of simulated years that 
pass before the population recovers to that size. The sixth 
file records the total reproductive value of the population 
just before and just after the spill. The reproductive value 
of a female is the relative number of female pups she is 
expected to wean during the remainder of her life. For a 
female of age x, Fisher (1930) and Wilson and Bossert (1971) 
give the formula for reproductive value (v,): 


Vn = (2/15) 6 am, |; y=x,W (26) 


Fig. 10.22 illustrates the reproductive values of females 
under default parameter settings and a per capita population 
growth rate = 0. The total reproductive value of the 
population is: 


v7 = nyv, ¢ =X=1,W (27) 


where n, is the number of females of age x in the population. 
The reduction in total reproductive value may provide a 
measure of how the perturbations in age and sex structure of 
the population caused by an oil spill effect population 
recovery. 


248 


FIGURE 10.21a -- Simulated movements of sea otters around an oil 
spill in Monterey Bay beginning 1 December and lasting 15 days. 
Boxed area (heavy dark line) represents spill location, between 
Fort Ord (365) and Point Pinos (390). Each trace represents a 
different simulated individual otter. Total range wide population 
at the time of the spill was set at 1600; OTDIST positioned 56 
animals in the area covered by the diagram but for clarity only 25 
were chosen, at random, for representation. Movement parameter 
settings (see text) were PM=0.0, PA=0.0, PE=0.0, for all days of 
the spill, thus simulating no oil-caused mortality or effect on 
behavior. The relatively few numbers of animals at the north end 
of the diagram reflect the much lower density of otters in the 
sandy habitat of Monterey Bay relative to the rocky habitat of the 
Monterey Peninsula. The numerical value of the locations along the 
five fathom line generally increases as one moves to the south 
along the California Coast. 


0 1 DharteneS tietares 4 eevee) OmereineeriS Se Oma UA nerlics 


345 


SDD 


365 


SYS 


385 


SEIS) 


405 


LOCATION ALONG 5 FATHOM LINE 


415 <2 aaa —— 
OM Si? Mae ae nome er One OL Bi 13) 14 


DAY OF SPIRE 


249 


FIGURE 10.21b -- Simulated movements of sea otters around an 
oil spill in Monterey Bay beginning 1 December and lasting 15 
days, as in Fig. 10.21a, except that in this simulation 
movement parameters were set at PM=0.8, PA=0.8, PE=0.5. A 
star (*) indicates an otter death. The numerical value of the 
locations along the five fathom line generally increases as 
one moves to the south along the California Coast. 


0 1 2 OB, VA Seth COE eS ATO RRO IE POTS IS 
345 


355) 


S199) 


SS 


385 


LOCATION ALONG 5 FATHOM LINE 


Onn Det Rayne 6 Ai Sh Ake Ova NERA 1s 
DANO Or SP iige 


14 


15 


FIGURE 10.21c -- Simulated movements of sea otters around an 
oil spill in Monterey Bay beginning 1 December and lasting 15 
days, as in Fig. 10.2la, except that in this simulation 
movement parameters were set at PM=0.1, PA=0.5, PE=0.0. A 
star (*) indicates an otter death. The numerical value of the 
locations along the five fathom line generally increases as 
one moves to the south along the California Coast. 


Cy ee! 2 ot Oph tim TD ee ene: She eal Ojes ylulbet ele al 


LOCATION ALONG 5 FATHOM LINE 


8 ao en 1D 13 
DAY OF SPILL 


251 


14 


14 


FIGURE 10.22 -- Age-specific reproductive values of female 
California sea otters under default parameter settings and a 
per capita growth rate=0. 


—_ 
on 


a 
[=) 


9 
ui 


REPRODUCTIVE VALUE (pups per female) 


0 5 6 7 8 9 1011 12 13 14 15 16 17 18 19 20 21 22 23 


AGE 


A data summary and analysis program, OTPROC, has been 
written to summarize the raw output files and write a short 
report on the run. Additionally, OTPROC writes output files 
specifically designed to be read by LOTUS123. A LOTUS123 
macro worksheet is supplied that will generate simple graphics 


illustrating the outcome of the model runs. Detailed 
discussion of the processing of model output can be found in 
the user's manual (Appendix 10.2). Examples of model output 


appear in Fig. 10.23. 
OTRANGE 


In an effort to be able to make predictions about the 
future status of the California sea otter population, and to 
aid in the estimation of some of the parameters used in OTPOP 
and LESLIE, we undertook an analysis and modeling of the 
historical growth of the population. Because of the type of 
data used and the speculative nature of the modeling, this 
work was conducted separately from the oil spill population 
model. 


Background 


The growth of the California sea otter population 
following almost complete elimination by the turn of the 20th 
century is described by CDFG (1976), Ralls, et al., (1983), 
and USFWS (1986). Information on the size and range of the 
population before 1968, when CDFG began intensive study of sea 
otters, is scanty and generally anecdotal. A remnant 
population of probably less than 100 individuals grew at what 
appeared to be a steady rate of approximately 5% per year 
until the mid 1970's, expanding the occupied range in the 
process. By 1969, when regular counts began, an estimated 
1,390 otters inhabited the coast between Seaside and Point 
Estero. In 1976 the highest population estimate was recorded, 
1,789 animals, and occupied range extended from Rio Del Mar 
to Pecho Rock. Population estimates have been relatively 
constant since 1976, while the range has continued to expand 
slightly. 


Reasons for the arrest in population growth have been the 
subject of much debate centering around whether density 
dependent or density independent factors have been most 
influential in bringing about the decline. Miller (1980) 
maintained that the population had reached an equilibrium with 
habitat carrying capacity, and that starvation was the 
principal cause of mortality. Estes et al (1986), based on 
comparisons of time budgets with an Alaska population assumed 
to be food limited, concluded that the California population 
was not at carrying capacity, and that density independent 
processes, particularly accidental entanglement in fishing 
nets, were the primary factors limiting population growth. 


253 


FIGURE 10.23a - Log file from a run of the model. A large 
oil spill is introduced for 10 days beginning June 15 along 
the 50km section of coast between Marina and Yankee Point. 
Initial population size is set at 1600, carrying capacity set 
at 1600. See user's manual for explanation of parameters. 


DATE AND TIME USED TO GENERATE RANDOM NUMBER SEEDS 
20,"NUMBER OF YEARS PER RUN" 
100,"NUMBER OF RUNS WITH OIL SPILL" 
100,"NUMBER OF CONTROL RUNS" 
1600,"INITIAL POPULATION SIZE" 
6,"MONTH OF SPILL" 
15,"DAY OF SPILL" 
10,"DURATION OF SPILL" 
350,"NORTH BOUNDARY OF SPILL" 
450,"SOUTH BOUNDARY OF SPILL" 
1600,"EQUILIBRIUM POPULATION SIZE" 
201,"NORTH BOUNDARY OF RANGE" 
955,"SOUTH BOUNDARY OF RANGE" 
0.090,"MAXIMUM PER CAPITA ANNUAL GROWTH RATE" 
0.005,"NON-LINEARITY OF DENSITY DEPENDENCE" 
0.000,"DENSITY INDEPENDENT MORTALITY RATE" 
1.000,"DEGREE OF COMPENSATION" 
0.930,"ADULT FEMALE SURVIVAL RATE" 
15.000,"MODEL AGE OF FEMALE SENESCENCE" 
0.250,"PRIME REPRODUCTIVE RATE" 
0.530,"PUP SURVIVAL RATE" 
0.870,"ADULT MALE SURVIVAL RATE" 
9.000,"MODEL MALE AGE OF SENESCENCE" 
5.000,"PERCENT VARIATION IN ADULT SURVIVAL" 
5.000,"PERCENT VARIATION IN PUP SURVIVAL" 
ww MAR" "CE" "STGMA", "VMAX" 
“JUVENILE FEMALES" 0.367-0.163 8.09037.500 
“ADULT FEMALES W/PUP"-0.025-0.406 6.39307.500 
"ADULT FEMALES W/O PUP"=0.009-0.706 2.950 8.400 
"JUVENILE MALES"-0.045-0.290 8.56048.900 
“ADULT NON-TERRITORIAL MALES" 0.105-0.815 

4.64048.900 
“ADULT TERRITORIAL MALES" 0.042-1.044 1.93048.900 
6,"AGE AT WHICH POTENTIALLY TERRITORIAL" 
60.000,"MAXIMUM % OF POTENTIALS THAT HOLD 
TERRITORIES" 
20.000,"MINIMUM % OF POTENTIALS THAT HOLD 
TERRITORIES" 
8.110,"MEAN TERRITORY LENGTH" 
0.440,"S.D. OF TERRITORY LENGTH" 
5,"MEAN ARRIVAL DATE MONTH" 
23,"MEAN ARRIVAL DATE DAY" 
11,"S.D. OF ARRIVAL DATE IN DAYS" 
12,"MEAN DEPARTURE DATE MONTH" 
1,"MEAN DEPARTURE DATE MONTH" 
15,"S.D. OF DEPARTURE DATE IN DAYS" 
"DAY", 

"P (MORTALITY) ","P (AVOIDANCE) ",""P (EMIGRATION) " 


N 


TABLE 23a. (continued) 


1 1.000 0.000 0.000 
2 1.000 0.000 0.000 
3 1.000 0.000 0.000 
4 1.000 0.000 0.000 
5 1.000 0.000 0.000 
6 1.000 0.000 0.000 
7 1.000 0.000 0.000 
8 1.000 0.000 0.000 
9 1.000 0.000 0.000 
10 1.000 0.000 0.000 


FIGURE 10.23b - Report file generated by OTPROC after the run in Fig. 10.23a. 


CONTROL RUNS: 


YEAR FEMALES MALES PUPS 
-4 1038.0¢1038-1038, 0.0) 566.0(566-566, 0.0) 195.5¢€175-220, 10.7) 
-3 1029.8¢ 997-1064, 13.6) 559.4(529-589, 12.8) 244.8(220-272, 11.8) 
-2 1047.1¢1002-1105, 17.7) 579.5(529-615, 18.4) 228.5(193-262, 12.3) 
-1 1056.4¢1015-1110, 18.6) 590.6(542-632, 19.3) 215.4¢(170-248, 13.8) 
0 1058.1¢1002-1123, 24.7) 597.0(536-652, 20.9) 214.3(¢185-239, 12.6) 
1 1035.7¢€ 979-1094, 23.4) 603.6¢560-658, 20.0) 212.5¢177-248, 12.9) 
2 1021.3¢ 978-1066, 1938) 606.9(558-650, 18.2) 213.8¢181-249, 14.5) 
3 997.1¢€ 916-1049, 22.2) 593.9(526-645, 21.0) 214.7(188-238, 9.5) 
4 997.9( 938-1063, 24.7) 582.2(520-640, 24.5) 211.0¢167-240, 11.1) 
5 998.8¢( 935-1058, 23.1) 573.3(509-628, 23.8) 208.8(189-235, 9.2) 
6 1007.0¢ 947-1071, 23.3) 568.7(506-623, 24.6) 207.6(175-234, 10.9) 
7 1010.4¢ 947-1074, 25.0) 563.4(508-627, 22.9) 210.9(¢186-238, 10.7) 
8 1018.3¢ 955-1084, 25.5) 559.8(504-619, 23.3) 213.1¢190-245, 11.3) 
9 1026.5¢ 969-1085, 24.7) 558.3(507-615, 24.4) 215.2¢183-250, 12.9) 
10 1031.2¢ 965-1090, 27.3) 557.3¢499-627, 26.6) 216.7¢179-244, 11.1) 
11 1037.4¢ 973-1098, 27.3) 555.8(508-623, 25.8) 219.1(194-242, 11.4) 
12 1038.5¢ 980-1103, 28.4) 558.0(484-617, 24.6) 220.9(187-255, 10.3) 
13. 1039.4¢ 964-1136, 28.1) 560.9(508-622, 21.1) 219.8(196-241, 9.9) 
14 1039.3¢ 965-1115, 26.0) 564.7(494-619, 21.6) 220.2(201-246, 9.4) 
15 1039.2¢ 980-1086, 23.1) 565.9(512-616, 21.2) 219.5¢(200-250, 9.9) 
16 1036.3¢ 976-1103, 23.3) 565.3¢(516-618, 21.8) 220.9(¢186-254, 12.2) 
17 1034.8¢( 972-1091, 23.4) 565.8(503-623, 23.4) 220.3¢188-240, 11.2) 
18 1039.0¢ 988-1108, 24.0) 565.9(510-636, 23.8) 218.2¢190-248, 11.1) 
19 1034.8¢ 975-1109, 25.6) 564.6(518-642, 22.6) 219.1€¢191-250, 11.0) 
20 1032.3¢ 966-1081, 23.8) 564.9(506-616, 23.0) 219.4¢(195-250, 10.6) 


256 


10.23b (continued) 


FEMALES 
-4 1038.0 (1038-1038, 


0.0) 

955-1061,13.0) 
983-1090,19.3) 
978-1108,26.3) 
972-1116,29.8) 


713- 860,29.4) 
740- 899,31.5) 
784- 949,28.7) 
836- 984,27.4) 


884-1019,26.6) 
894-1048,28.0) 
933-1085,26.6) 
983-1119,26.2) 


(1008-1128,23.5) 
(1029-1135,22.3) 
(1022-1146,26.2) 
(1011-1124,24.4) 


968-1110,27.7) 
963-1107,25.8) 
973-1126. 3) 
972-1096,27.9) 
924-1101,27.8) 
896-1132,29.6) 
892-1088,27.2) 
902-1082,28.3) 


FIGURE 
OIL SPILL RUNS: 
YEAR 
-3 1030.4 ¢ 
-2 1044.0 ¢ 
-1 1049.3 ¢ 
-0 1047.4 ¢ 
+0 785.3 ¢ 
1 837.5 ¢ 
2 887.3 ¢ 
3 923.5 ¢ 
4 956.7 ¢ 
5 989.5 ¢ 
6 1019.8 ¢ 
7 1045.6 ¢ 
8 1063.4 
9 1073.8 
10 1077.9 
11 1068.6 
12 1060.1 ¢ 
13° 1047.4 (¢ 
14 1039.0 ¢ 
15 1033.9 ¢ 
16 1027.7 (¢ 
17 1025.3 ¢ 
18 1020.4 ¢ 
19 1018.4 ¢ 
20 1021.8 ¢ 


956-1075,24.7) 


NUMBER OF DEATHS FROM OIL SPILL: 


CLASS MEAN s.D 
JUVENILE MALES 13.4 4.0 
ADULT MALES 18.7 3.2 
JUVENILE FEMALES 75.7 8.3 
ADULT FEMALES 186.4 12.1 
PUPS 48.4 6.6 
TOTAL ANIMALS 342.5 19.0 


RECOVERY AFTER OIL SPILL: 


REPRODUCTIVE VALUE BEFORE SPILL 
REPRODUCTIVE VALUE AFTER SPILL 
REDUCTION (4) 

YEARS TO RECOVERY 


**ON 25 OF 


PRE-SPILL SIZE 
TIME TO RECOVERY CALCULATED ONLY FOR RUNS THAT DID RECOVER. 


100 RUNS (¢ 25.0%) THE POPULATION 


MALES 
566.0(566-566, 0.0) 
559.1(530-589,12.9) 
573.5(531-617,18.8) 
585.2(531-624,21.2) 
590.2(¢541-652,24.2) 
558.1(508-612,23.9) 
540.3(472-612,26.3) 
530.0(468-609,28.6) 
515.4(462-575,26.6) 
501.0¢440-563,27.2) 
496.4(418-566,25.0) 
496.0(¢431-571,24.8) 
500.7(435-579,25.9) 
512.8(¢453-582,24.3) 
527.7¢6458-578,25.6) 
542.5(490-615,24.2) 
556.1(511-608,23.8) 
567.4(519-634,23.5) 
574.8(519-643,24.6) 
578.6(531-638,24.6) 
579.1(528-666,27.4) 
578.6(532-638,24.5) 
577.4(520-642,22.8) 
572.3¢(524-619,22.3) 
568.3(514-628,23.8) 
568.1(511-626,25.1) 


PUPS 


196.4(165-217,11.1 


232. 
219. 
211. 
211. 
162. 
169. 
176. 
181. 
188. 
197. 
206. 
214. 
221. 
227. 
229. 
230. 
228. 
227. 
225. 
223. 
219. 
215. 
215. 
214. 
213. 


7¢195-277,16 
0(¢188-254,16 
5(187-251,13 
0¢180-251,12 
6(134-194,11 
6(140-208,12 
0¢151-205,12 


3(150-201,10. 


4(157-210,11 


1¢169-230,10. 
11132231, 10+ 


3(¢186-241,11 


0¢200-244,10. 
6(204-255,10. 


9¢201-259,10 


1(196-263,12. 
4(189-258,12. 


4(202-265,11 
5(189-262,11 
3(200-256,10 
4(196-246,11 
2(¢184-252,11 
4(166-238,11 
0¢165,-241,11 
7(188-240,10 


PERCENT OF POPULATION 


RANGE MEAN  S.D. 
oc @5) 4.96 1.48 
2 74h) 5.86 1.01 
- 100 26.88 2.65 
- 222 24.35 1.58 
SONG, 22.95 2.94 
- 383 18.54 1.06 

MEAN $.D. 
1316.3 39.5 
987.7 39.4 
24.971 1.52641 
9.800 1.71270 


257 


RA 
1.90 
3.40 

22.00 
20.80 
15.20 
16.20 


RAN 


NGE 
7 


GE 


-90 
8. 
37. 
0) 4 
32. 
21. 


90 
20 
40 
00 
00 


1223.2-1409.9 

889.9-1073.3 
22.271-28.831 
7.000-17.000 
DID NOT RECOVER TO 


-0) 
-0) 
-3) 
5) 
-8) 
5&2) 
at) 
3) 
-4) 
5) 
6) 
-6) 
5) 
9) 
-5) 
3) 
5) 
-8) 
-5) 
-7) 
oI} 
-7) 
4) 
-0) 
-7) 


v 


FIGURE 10.23c -- Trace of the total simulated population size 
for, runs) in Fig. 10.23a-.) Meanvand Gange sot) 1 OO suns Ousl: 
spill occurs at year 0. 


1800 
WY) 
4 
FE 
O high 
ue 1600 
ZZ 
Li 
oO = 
FZ 
Lil \ 
O 1400 + | mee 
OQ 
z 
LL low 
O 
1200 
oO 
LJ 
faa) 
= 
=) 
= 1000 


Sy <@Q@ © 2 4 6 8 ‘Om Wecatke «AG 18 


YEAR OF SIMULATION 


FIGURE 10.23d -- Trace of the total simulated population size 
for the control (no oil spill) runs in Fig. 10.23a. Mean and 
range of 100 runs. 


NUMBER OF INDEPENDENT OTTERS 


1800 


1600 


1200 


1000 


=i 


high 
mean 


low 


— 2) 0 2 4 6 8 10 12 14 16 18 


YEAR OF SIMULATION 


20 


FIGURE 10.23e -- Trace comparing the mean values from Figs. 
10.23c and 10.23d. 


NUMBER OF INDEPENDENT OTTERS 


1800 


control 
1600 


1400 


oil spill 


1200 


1000 
—4 -—2 0 2 4 6 8 10 12 14 


YEAR OF SIMULATION 


16 


18 


20 


CUMULATIVE NO. OF DEATHS 


FIGURE 10.23f -- Mean cumulative number of otter deaths due 
to oiling for the run in Fig. 10.23a. 


300 
independents 

240 

180 rn 

120 
pups 

60 
0 - 
0 2 4 6 8 10 
DAY OR TSPIFE 


261 


Relatively high levels of commercial gill- and trammel-net 
fishing along the central coast began in the early 1970's, 
coincident with the decline in sea otter population growth. 
Wendell, et al., (1985) estimated that approximately 80 sea 
otters drowned accidentally in nets each year between 1973 
and 1983. Ames, et al., (1985) suggested that density 
dependent processes are most important in the central part of 
the range, where the population has been established for the 
longest period of time and has depressed prey populations 
below pre-recolonization levels, while at the range 
peripheries, which still have relatively low numbers of 
otters, density independent processes, particularly accidental 
entanglement and great white shark attacks are the most 
important factors limiting the growth of the population. The 
time budget data we gathered using telemetry in areas of the 
central part of the range tended to support the suggestion 
that the juvenile females may be suffering from food 
limitations in the central part of the range (see Chapter 4), 
but the evidence is not conclusive and does not address the 
situation at the range peripheries. 


The degree of density dependence in the dynamics of a 
particular population is a question that must be addressed by 
any attempt to model that population. The uncertainty 
surrounding this issue in the California sea otter population 
led us to build in to our model the capability of simulating 
any degree of density dependence. This approach allows 
Simulation of the effect of oil spills under different 
conceptual hypotheses about the dynamics of the population, 
but does little towards providing a set of parameters most 
applicable to a timely and realistic risk analysis. In an 
effort to better understand the dynamics and estimate 
pertinent parameters we have attempted to simulate the 
historical growth of the California population with a model 
(OTRANGE) that incorporates a feedback mechanism between 
population size and range size. It is hoped that a model that 
fits the historical data will be of value in predicting future 
population sizes and range extent. Because of the speculative 
nature of the mechanisms incorporated in the model, and the 
amount of computer time required to fit model parameters to 
the historical data, OTRANGE is completely deterministic. It 
is intended only as an aid in determining the values of user- 
input parameters in the main risk analysis model, particularly 
when considering spills in the future. 


Structure 


The density dependence function used in the population 
dynamics portion of OTRANGE is the same as that used in OTPOP 
lolol IHSNGIGA ((¥eig (3), Imskef, al@odt, eyerel rely IO) 25) o Range 
expansion is incorporated with the assumption that range 
expansion is density dependent, and that range expansion 


262 


results in an increased population-wide carrying capacity 
(Fig. 10.24). Biologically, this means that as the population 
approaches its carrying capacity it can respond by both 
reducing its growth rate, and by increasing its carrying 
capacity through an increase in the area of occupied habitat. 

The density dependence function for range 

size is simply taken as the mirror image of the density 
dependence function for population size (Fig. 10.25): 


kK = Kpax~Kmax{1-exp{-b (K-N) ]} (27) 


where k is the annual rate of change in carrying capacity, 
Knax iS the maximum annual rate of change of k, K is the 
carrying capacity, N is the population size, and b governs 
the shape of the curve. As in the population dynamics portion 
of the model, an arbitrary ceiling on the size of k is imposed 


at Kmax- 


The fact that occupied range has expanded faster to the 
south than to the north, and the fact that the range continued 
to expand during the period of apparent population decline in 
the late 1970,s, led us to modify (27) in the model. First, 
range expansion to the north and to the south are considered 
separately. Secondly, allowance is made for density 
independent range expansion. Biologically, density 
independent range expansion may be the result of natural 
dispersal of young animals, regardless of the equilibrium 
status of the population, superimposed on a limited geographic 
range. OTRANGE thus uses 


Kg Kgs + Kmax,s7Kmax,s{1-exp{-b (K-N) ] } and 
Ky Kan + Knaxin@ Kua, n{ 1-exp{-b(K-N) ] } (28) 


where k, and k, are annual range expansion rates to the south 
and to the north, respectively, ky, and ky, are density 
independent rates of range expansion ‘to the south and to the 
north respectively, Kmax,s and Kpax,, are the maximum rates of 
range expansion to the south and to the north, respectively, 
and b, K, and N are as in (27). 


til 


Translating k, which is terms of numbers of animals, into 
range size requires an estimate of the number of animals that 
can be supported in a given area of habitat. Ford and Bonnell 
(1986), using USFWS census data, estimated that maximal 
densities were 4.7 otters per km? over rocky substrate and 1.3 
otters per km, over sandy substrate. By digitizing the coast, 
assuming that otter habitat extended from the coast to the 20 


263 


FIGURE 10.24 -- Schematic representation of OTRANGE. At each 
iteration, the quantity K-N determines the next iteration's 
population size and carrying capacity. 


KN 


Density 
dependence 


264 


FIGURE 10.25 -- Density dependence functions used in OTRANGE. 
The parameter b affects the shape of the curve (see Fig. 
10.4). 


Kegs 


a LJ 
<l = 
(eZ <l 
(ee 
ma 
= FE, 
= 6 
ra Z 
O < 
ae 
: a 
(als 
<L Lod 
O O 
on x 
Oo Cie 
population below e) population above 
carrying capacity carrying capacity 


N—K 


fathom depth contour, coding substrate from USGS maps, and 
using Ford and Bonnell's (1986) estimates, we calculated the 
carrying capacity of each 0.5 km segment of the 5 fathom line 
ordinate system. This allowed calculation of historical 
values of K given the length of occupied range at any point 
in the past. Using the historical range length data in USFWS 
(1986, Table 1.3) we calculated historical carrying capacities 
based on Ford and Bonnell's (1986) estimates. The calculated 
carrying capacities were well below historical population 
estimates, leading us to believe that Ford and Bonnell's 
census-based estimates underestimated actual carrying 
capacity, and forcing us to include the per hectare carrying 
capacities of rock and sand substrates in the group of 
parameters to be estimated. 


Parameterization 


We estimated OTRANGE parameters by means of a 2-stage 
numerical search. First, we chose what seemed to be 
reasonable bounds on the value of each parameter in the model, 
and a testing interval for each parameter between those 
bounds. Then the model was run under every possible 
combination of parameter values within the bounds, using the 
1914 historical estimates of population size and carrying 
capacity as initial conditions. Goodness of fit to historical 
data was calculated for each run by comparing the modeled 
population size and range-wide carrying capacity to the 
historical estimates of those values for each year that 
historical data were available. Total sum of squares of the 
aifference between the modeled and historical values was the 
goodness of fit criterion. 


A total of over 196,000 combinations of parameter values 
were tested in the first stage of estimation. The 20 
parameter combinations that gave the best fit to the 
historical data by each of the goodness of fit criteria were 
saved, new bounds chosen from those combinations, and the 
analysis repeated with a smaller testing interval for each 
parameter. Over 40,000 parameter combinations were tested in 
this second stage of estimation. The parameter combination 
that gave the best fit is listed in Table 10.7, and a trace 
of the model run under this parameterization is shown in Fig. 
10.26. 


266 


Table 10.7. Parameters giving the best fit of OTRANGE to 
historical data. See text for explanation of parameters. 


OTRANGE without OTRANGE with 
density independent density independent 
Parameter mortality mortality 
Maximum per capita 
growth rate (Ymax) 0.085 0.077 
Non-linearity of 
density dependence (b) 0.020 0.030 


Maximum density dependent 
rate of expansion: 
North (Kmex,n) * 7.5 8.5 
South (Kpax,s) : 20.7 27.5 
Density independent rate 
of expansion: 


North (kg py) : 8.0 2.6 
South (Kg ,) : 10.0 15 
Per hectare carrying 
capacity 
Rock substrate: 0.26 6.5 
Sand substrate: 0.78 6.75 
Density independent 
mortality rate (m) -- 0.03 


Incorporation of density independent mortality. 


The parameterization of the above model that produced the 
best fit to the historical data produced oscillations in 
population size once the population approached 1600 animals. 
This may imply that the stabilization of population size that 
occurred in the mid-1970's is a consequence of the internal 
dynamics of the population and its habitat, and that 
stabilization would have occurred even without gill net 
mortality. In other words, it implies that gill-net mortality 
completely compensates for natural mortality, and that if 
animals would not have been killed in the nets they would have 
died from natural causes. Certainly natural mortality is 
greater than 5% per year (see Figs. 10.7,10.9, and 10.10), and 
compensatory density independent mortality of the magnitude 
due to gill-nets is mathematically possible. 


We decided to investigate the possibility that gill-net 
mortality is completely additive to natural mortality. We 
thus modified the population growth equation: 

Yr = Ymax{1-exp[-b(K-N) ]}-m (29) 
where m is the density independent mortality rate. We then 


estimated parameters in the same manner as above, including 


267 


FIGURE 10.26 -- Fit of OTRANGE output to historical data using 
the parameters in Table 10.7 without density independent 
mortality. Solid line traces population size, dashed line 
traces carrying capacity. 


2400 
Historical A 
population size A 
Bolo Historical a 
carrying capacity s 


1600 


1200 


800 


NUMBERS OF OTTERS 


400 


1915 1925 19355 1945 1955 1965 1975 1985 


YEAR 


268 


FIGURE 10.27 -- Fit of OTRANGE output to historical data using 
the parameters in Table 10.7 incorporating density independent 
mortality. Solid line traces population size, dashed line 
traces carrying capacity. 


2400 
Historical 
population size 
& 
fe 2000 Historical a " 
carrying capacity oe oe 
E < 
1 
oO 600 
Le 
O 
1200 
”Y 
or 
a 
800 
= 
=) 
Fz, 


400 


1915 3925 1935 1945 1955 1965 1975 1985 


YEAR 


269 


m in the list of parameters that were estimated. Density 
independent mortality was incorporated in the model only after 
1972, to simulate the effect of gill-net mortality. A trace 
of the simulated dynamics that produced the best fit appears 
Wg IPSS ALO) 6 BW 


Caveats 


The model incorporating density independent mortality 
produced a better fit to the historical data than did the 
basic model (total sum of squares =870,616 vs. 1,121,621). 
In the parameterization giving the best fit, m =0.03, less 
than the 5% gill net mortality estimated by Wendell et al 
(1985), perhaps indicating that gill net mortality is partly 
compensatory. 


Neither of the models fits the historical data 
particularly well, and there is no guarantee that the 
mechanisms in the model mimic those in the natural population. 
There are two levels on which our approach should be 
criticized. First, we made no effort to investigate different 
functional forms of density dependent range expansion. As far 
as we know no other attempts have been made to model density- 
dependent changes in carrying capacity in any situation; we 
did not have the advantage of theoretical precedents. We used 
the mirror image of the population growth function as a 
convenient starting point, but there is no reason to assume 
that it is correct. Secondly, of course, parameter estimates 
are only as good as the data they are based on. For example, 
Geibel and Miller (1980) and Wendell, et al., (1986) described 
the problems associated with the estimation of sea otter 
population size. 


Relationship between OTRANGE and OTPOP 


The analysis of OTRANGE under different parameterizations 
gave the default parameter values for r,,,, b, and K used in 
OTPOP and LESLIE (Table 10.2). Because the proposed oil and 
gas developments that are of concern will not occur until the 
1990's, at the earliest, and because much of that development 
is proposed for what is now peripheral to occupied sea otter 
range, we have supplied OTRANGE to MMS in a form that can be 
used to predict future population size, range size, and 
carrying capacity. These predictions can then be used as 
input to the main model to simulate initial conditions in the 
future. 


LITERATURE CITED 


Ames, R. A., F. E. Wendell, and J. J. Geibel. 1985. Sea 
otter mortality in California. Draft unpublished 


270 


report. Marine Resources Branch, California Department 
of Fish and Game. 49pp. 


California Department of Fish and Game. 1976. A proposal for 
sea otter protection and research, and request for return 
of management to the state of California. Unpubl. 
report, January 1976. 2 Vols. 


Chapman, D. G. and D. S. Robson. 1960. The analysis of a 
catch curve. Biometrics 16:354-368. 


Costa, D. P., and G. L. Kooyman. 1982. Oxygen consumption, 
thermoregulation, and the effect of fur oiling and 
washing on the sea otter, Enhydra lutris. Can J. 
Zoology 60:2761-2767. 


DeMaster, D. P. 1981. Incorporation of density dependence 
and harvest into a general population model for seals. 
Pp. 389-402 in C. W. Fowler and T. D. Smith, eds., 
Dynamics of large mammal populations. J. Wiley and Sons, 
New York. 476pp. 


Eberhardt, L. L. 1986. Assessing the dynamics of wild 
populations. J. Wildl. Manage. 49:997-1012. 


Eberhardt, 9b.) L., ands "Ds 9B. Sinver. OMe Population 
dynamics of the Pribilof fur seals. Pp 197-220 inc. W. 
Fowler and T. D. Smith, eds., Dynamics of large mammal 
populations. J. Wiley and Sons, New York. 476pp. 


Estes, J. A., R. J. Jameson, and E. B. Rhode. 1982. Activity 
and prey selection in the sea otter: influence of 
population status on community structure. Am. Nat. 
120:242-258. 


Estes, J. A., K. E. Underwood, and M. J. Karmann. 1986. 
Activity-time budgets of sea otters in California. J. 
Wildl. Manage. 50:626-636. 


Fisher, R. A. 1930. The genetical theory of natural 
selection. Clarendon, Oxford. 


Ford, R. G. 1985. A risk analysis model for marine mammals 
and seabirds: a Southern California Bight scenario. 
Final Report. Minerals Management Service Pacific OCS 
Region Contract No. 14-12-0001-30224. Los Angeles. 
236pp. 


Ford, R. G., J. A. Weins, D. Heinemann and G. L. Hunt. 1982. 


Modeling the sensitivity of colonially breeding marine 
birds. co) Old Spills: Guilemots and /kittiwake 


271 


populations on the Pribilof Islands, Bering Sea. J. 
Applied Ecology 19:1-31. 


Ford, R. G., and M. L. Bonnell. 1986. Analysis of the risk 
of oil spills to sea otters-methodology. Technical 
Support Document 3, Draft Environmental Impact Statement 
for Proposed Translocation of Southern Sea Otters. US 
Fish and Wildlife Service. 


Fowler, C. W. 1981. Comparative population dynamics in large 
mammals. Pp 437-456 in C. W. Fowler and T. D. Smith, 
eds., Dynamics of large mammal populations. J. Wiley and 
Sons, New York. 476pp. 


Garshelis, D. L. 1983. Ecology of sea otters in Prince 
William Sound, Alaska. Ph. D. Thesis. Univ. of 
Minnesota, Minneapolis. 321pp. 


Geibel, J. J., and D. J. Miller. 1984. Estimation of sea 
otter, Enhydra lutris, population, with confidence 
bounds, from air and ground counts. California Fish and 
Game 70:225-233. 


Green, B. 1978. Sexual maturity and senescence of the male 
California sea otter (Enhydra lutris). M. S. Thesis. 
San Jose State University, California. 


Harris, R. B., L. A. Maguire, and M. L. Shaffer. 1987. Sample 
sizes for minimum viable population estimation. Conserv. 
Biol. 1:72-76. 


Heisey, D. M., and T. K. Fuller. 1985. Evaluation of 
survival and cause-specific mortality rates using 
telemetry data. J. Wildl. Manage. 49:668-674. 


Jameson, R. J. 1987. Movements, home range, and territories 
of male sea otters in central California. Manuscript in 
review. 22pp. 


ett Push waeRaeks Mohn; gandsDemroEGray., 196i. sensitiv 
dependent processes and management strategy for the 
northwest Atlantic harp seal populations. Pp. 135-158 
in Cc. W. Fowler and T. D. Smith, eds., Dynamics of large 
mammal populations. J. Wiley and Sons, New York. 476pp. 


Lotka, A. J. 1907. Relation between birth and death rates. 
Science, N.S. 26:21-22. 


Loughlin, T. R., J. A. Ames, and J. E. Vandevere. 1981. 
Annual reproduction, dependency period, and apparent 
gestation period in two California sea otters, Enhydra 
lutris. Fishery Bull. 79:347-349. 


272 


Miller, D. J. 1980. The sea otter in California. Pp. 79-81 
in CalCOFI Rep. Vol. XXI. 


Ralls, K., J. Ballou, and R. L. Brownell. 1983. Genetic 
diversity in California sea otters: theoretical 
considerations and management implications. Biol. 
Conserv. 25:209-232. 


Reed, M., D. French, J Calambokidis, and J. Cubbage. 
Simulation modeling of the effects of oil spills on 
population dynamics of northern fur seals. Final Report. 
Minerals Management Service Alaska OCS Region. Contract 
NO. 14-12-0001-30145. Anchorage, AK 139pp. 


Schneider, K. B. 1972. Reproduction in the female sea otter. 
Federal Aid in Wildlife Restoration Project W-17-1. 
Project progress report. 26pp. 


Schneider, K. B. 1978. Sex and Age segregation in sea 
otters. Federal Aid in Wildlife Restoration Project W- 
17-4 and W-17-5. Final Report. Alaska Dept. of Fish and 
Game. 45pp. 


Siler, W. 1979. A competing-risk model for animal mortality. 
Ecology 60:750-757. 


Siniff, D. B., T. D. Williams, A. M. Johnson, and D. L. 
Garshelis. 1982. Experiments on the response of sea 
otters, Enhydra lutris, to oil contamination. Biol. 
Conserv. 2:261-272. 


Wendell, F. E., J. A. Ames, and R. A. Hardy. 1984. Pup 
dependency and length of reproductive cycle: estimates 
from observations of tagged sea otters, Enhydra lutris, 
in California. Calif. Fish and Game 70:89-100. 


Wendell, F. E., J. A. Ames, and R. A. Hardy. 1986. A review 
of California sea otter, Enhydra lutris, surveys. Marine 
resources Technical Report No. 51. 42pp. 


Wendell, F. E., R. A. Hardy, and J. A. Ames. 1985. 
Assessment of the accidental take of sea otters, Enhydra 
lutris, in gill and trammel nets. Draft Report. 
California Dept. of Fish and Game. Morro Bay, CA. 30pp. 

Wilson, E. O., and W. H. Bossert. 1971. A primer of 
population biology. Sinauer Associates,Inc. Sunderland, 
MA 192pp. 


United States Fish and Wildlife Service. 1986. Summary of 
information on the biology of the southern sea otter. 


273 


Technical Support Document 1, Draft Environmental Impact 
Statement for Proposed Translocation of Southern Sea 
Otters. 7Opp. 


APPENDICES 


Appendix 2.1 - Reproductive data on individual females. 


SEA OTTER 6 


DATE DAYS CUM STATUS 
BETWEEN DAYS 
05-JUL-84 NO PUP 
07-JUL-84 2 2 NO PUP 
08-JUL-84 1 3 NO PUP 
17-JUL-84 9 12 NO PUP 
18-JUL-84 1 13 NO PUP 
19-JUL=84 1 14 NO PUP 
21-Jul-84 2 16 NO PUP 
22-Jul-84 1 17 NO PUP 
23-Jul-84 1 18 NO PUP 
25-Jul-84 2 20 NO PUP 
26-Jul-84 1 21 NO PUP 
27-Jul-84 1 22 NO PUP 
28-Jul-84 1 23 NO PUP 
29-Jul-84 1 24 NO PUP 
30-Jul-84 1 25 NO PUP 
08-Aug-84 9 34 NO PUP 
13-Aug-84 5 39 NO PUP 
21-Aug-84 8 47 NO PUP 
31-Aug-84 10 57 NO PUP 
01-Sep-84 1 58 NO PUP 
01-Sep-84 fo) 58 NO PUP 
02-Sep-84 1 59 NO PUP 
03-Sep-84 1 60 NO PUP 
05-Sep-84 2 62 NO PUP 
06-Sep-84 1 63 NO PUP 
08-Sep-84 2 65 NO PUP 
12-Sep-84 4 69 NO PUP 
15-Sep-84 3 72 NO PUP 
17-Sep-84 2 74 NO PUP 
14-Nov-84 58 132 NO PUP 
29-Nov-84 15 147 NO PUP 
21-Dec-84 22 169 NO PUP 
26-Dec-84 5 174 NO PUP 
15-Jan-85 20 194 NO PUP 
20-Feb-85 36 230 NO PUP 
21-Feb-85 1 231 NO PUP 
06-Mar-85 13 244 NO PUP 
12-Mar-85 6 250 NO PUP 
13-Mar-85 1 251 NO PUP 
14-Mar-85 1 252 NO PUP 
02-Apr-85 19 271 NO PUP 
16-Apr-85 14 285 NO PUP 
18-Apr-85 ae 287 NO PUP 
02-May-85 14 301 NO PUP 
11-Jun-85 40 341 NO PUP 


21-Jun-85 10 351 NO PUP 


SEA OTTER 9 


DATE 


02-Mar-85 
07-Mar-85 
08-Mar-85 
19-Mar-85 
01-Apr-85 
02-Apr-85 
11-Apr-85 
15-Apr-85 
20-Apr-85 
22-Apr-85 
23-Apr-85 
04-May-85 
05-May-85 
07-May-85 
15-May-85 
16-May-85 
17-May-85 
28-May-85 
10-Jun-85 
23-Jun-85 
24-Jun-85 
25-Jun-85 
27-Jun-85 
29-Jun-85 
02-Jul-85 
03-Jul-85 
20-Jul-85 
29-Jul-85 
09-Aug-85 
12-Aug-85 
15-Aug-85 
16-Aug-85 
23-Aug-85 
24-Aug-85 
03-Sep-85 
04-Sep-85 
06-Sep-85 
12-Sep-85 
07-Oct-85 
23-Oct-85 
31-Oct-85 
15-Nov-85 
12-Dec-85 
15-Dec-85 
08-Jan-86 
10-Jan-86 


DAYS 
BETWEEN 


PRR b 
WWPPPONPRPPNULUP 


[ed 


Pr 
DNPRPORPNRPWWRPUOYIRPWNNEF EF 


r 


N 
oO 


16 


CUM 


DAYS 


277 


SEA OTTER 9 (cont.) 


DATES 


14-Jan-86 
23-Jan-86 
05-Feb-86 
08-Feb-86 
10-Feb-86 
13-Feb-86 
06-Mar-86 
17-Mar-86 
29-Mar-86 
18-Apr-86 
14-May-86 
18-Jun-86 
05-Jul-86 
25-Aug-86 


DAYS 
BETWEEN 


CUM 


DAYS 


318 
327 
340 
343 
345 
348 
369 
380 
392 
412 
438 
473 
490 
541 


278 


SEA OTTER 11 


DATE DAYS CUM STATUS 
BETWEEN DAYS 
17-May-85 NO PUP 
07-Sep-85 113 113 PUP 


22-Jul-86 318 431 NO PUP 


279 


SEA OTTER 14 


DATE 


18-Mar-85 
19-Mar-85 
20-Mar-85 
27-Mar-85 
02-Apr-85 
08-Apr-85 
14-Apr-85 
20-Apr-85 
21-Apr-85 
22-Apr-85 
01-May-85 
14-May-85 
14-May-85 
14-May-85 
15-May-85 
15-May-85 
16-May-85 
18-May-85 
28-May-85 
10-Jun-85 
12-Jun-85 
23-Jun-85 
24-Jun-85 
25-Jun-85 
29-Jun-85 
01-Jul-85 
02-Jul-85 
03-Jul-85 
20-Jul=85 
27-Jul-85 
29-Jul-85 
05-Aug=85 
12-Aug-85 
15-Aug-85 
16-Aug-85 
20-Aug-85 
22-Aug-85 
23-Aug-85 
24-Aug=85 
27-Aug-85 
31-Aug-85 
01-Sep=85 
04-Sep-85 
05-Sep=85 
06-Sep-85 
10-Sep-85 
11-Sep-85 


DAYS 
BETWEEN 


rR r 
PNWONFPORPOOWUWOPPANDADYPHE 


=) 


ra 


PPPPWUPPBUPPNAPUNUIYNNYIYGVPPNRPP 


DAYS 


280 


SEA OTTER 14 (cont.) 


DATES 


19-Sep-85 
25-Sep-85 
30-Sep-85 
23-Oct-85 
28-Oct-85 
31-Oct-85 
15-Nov-85 
19-Nov-85 
09-Dec-85 
10-Dec-85 
14-Dec-85 
28-Dec-85 
18-Jan-86 
29-Jan-86 
06-Mar-86 
09-Mar-86 
10-Mar-86 
13-Mar-86 
17-Mar-86 
01-Apr-86 
26-Apr-86 
05-Jul-86 
06-Aug-86 
27-Aug-86 
30-Aug-86 
12-Sep-86 
03-Jan-87 
14-Feb-87 
16-Mar-87 


DAYS CUM 
BETWEEN DAYS 
8 185 
6 191 
5 196 
23 219 
5 224 
3 227 
15 242 
4 246 
20 266 
1 267 
4 271 
14 285 
21 306 
11 317 
36 353 
3 356 
1 357 
3 360 
4 364 
15 379 
25 404 
70 474 
32 506 
21 527 
3 530 
13 543 
113 656 
42 698 


281 


SEA OTTER 15 


DATE 


21-Mar-85 
22-Mar-85 
24-Mar-85 
26-Mar-85 
27-Mar-85 
28-Mar-85 
29-Mar-85 
30-Mar-85 
31-Mar-85 
01-Apr-85 
02-Apr-85 
06-Apr-85 
08-Apr-85 
11-Apr-85 
12-Apr-85 
13-Apr-85 
14-Apr-85 
15-Apr-85 
16-Apr-85 
17-Apr-85 
18-Apr-85 
20-Apr-85 
22-Apr-85 
28-Apr-85 
30-Apr-85 
02-May-85 
03-May-85 
04-May-85 
05-May-85 
06-May-85 
07-May-85 
07-May-85 
11-May-85 
12-May-85 
13-May-85 
15-May-85 
16-May-85 
17-May-85 
23-May-85 
24-May-85 
25-May-85 
26-May-85 
28-May-85 
02-Jun-85 
06-Jun-85 
08-Jun-85 


DAYS 
BETWEEN 


NVPUNPRPPOAPPNPPROPPRPRPPNNONNEPRPRPRPEPPPUNAPRPPPPRPERPNNE 


CUM 


DAYS 


282 


SEA OTTER 15 (cont.) 


DATE 


10-Jun-85 
11-Jun-85 
12-Jun-85 
24-Jun-85 
25-Jun-85 
29-Jun-85 
01-Jul-85 
03-Jul-85 
06-Jul-85 
08-Jul-85 
25-Jul-85 
25-Jul-85 
27-Jul-85 
29-Jul-85 
31-Jul-85 
01-Aug-85 
02-Aug-85 
05-Aug-85 
14-Aug-85 
16-Aug-85 
17-Aug-85 
19-Aug-85 
20-Aug-85 
23-Aug-85 
24-Aug-85 
27-Aug-85 
31-Aug-85 
01-Sep-85 
04-Sep-85 
05-Sep-85 
06-Sep-85 
07-Sep-85 
10-Sep-85 
12-Sep-85 
13-Sep-85 
19-Sep-85 
24-Sep-85 
25-Sep-85 
28-Sep-85 
10-Oct-85 
14-Oct-85 
15-Oct-85 
16-Oct-85 
18-Oct-85 
21-Oct-85 
23-Oct-85 
26-Oct-85 


DAYS 
BETWEEN 


a 


NNWNNPRPNRPEHN 


Rr 


WNWNPRPENWPUDPNWPPPWUPARWPUWUPNENWOWPPNNNO 


SEA OTTER 15 (cont.) 


DATES 


31-Oct-85 
07-Nov-85 
11-Nov-85 
12-Nov-85 
15-Nov-85 
19-Nov-85 
20-Nov-85 
08-Dec=-85 
10-Dec-85 
13-Dec-85 
14-Dec-85 
28-Dec-85 
03-Jan-86 
08-Jan-86 
09-Jan-86 
13-Jan-86 
18-Jan=86 
24-Jan-86 
25-Jan-86 
28-Jan-86 
03-Feb=86 
07-Feb=86 
08=Feb=-86 
13-Feb=-86 
15-Feb=-86 
19-Feb-86 
06-Mar-86 
10-Mar-86 
12-Mar-86 
13-Mar-86 
24-Mar-86 
10-Apr-86 
18-Apr-86 
29-Apr-86 
18-May-86 
10-Jun-86 
24-Jun-86 
05-Jul-86 
28-Jul-86 
05-Aug-86 
08-Aug-86 
27-Aug-86 
12-Sep-86 
03-Oct-86 
09-Oct-86 
02-Nov-86 
08-Nov-86 


DAYS 
BETWEEN 


=) 
PPUNORPhWP EY U 


Pr 
OPNURFPEAWRPAUNPPUOD 


Rr 


PP 
OIP PN > 


CUM 
DAYS 


224 
231 
235 
236 
239 
243 
244 
262 
264 
267 
268 
282 
288 
293 
294 
298 
303 
309 
310 
313 
319 
323 
324 
329 
331 
335 
350 
354 
356 
357 
368 
385 
393 
404 
423 
446 
460 
471 
494 
502 
505 
524 
540 
561 
567 
591 
597 


284 


SEA OTTER 15 (cont.) 


DATE 


26-Nov-86 
13-Dec-86 
15-Dec-86 
07-Jan-87 
06-Mar-87 


DAYS 
BETWEEN 


18 
17 

2 
23 
58 


CUM STATUS 
DAYS 

615 NO PUP 
632 NO PUP 
634 PUP 
657 PUP 
715 PUP 


285 


SEA OTTER 16 (cont.) 


DATE 


08-Apr-85 
09-Apr-85 
11-Apr-85 
13-Apr-85 
15-Apr-85 
20-Apr-85 
25-Apr-85 
06-May-85 
16-May-85 
01-Jul-85 
06-Jul-85 
16-Aug-85 
25-Sep-85 
15-Nov-85 
11-Dec-85 
25-Mar-86 
30-Apr-86 
12-Jun-86 
04-Jul-86 
01-Nov-86 
23-Nov-86 
29-Nov-86 


DAYS 
BETWEEN 


SEA OTTER 19 


DATES DAYS CUM STATUS 
BETWEEN DAYS 

08-Apr-85 

09-Apr-85 a 1 PUP 
11-Apr-85 2 3 PUP 
11-Apr-85 0 3 PUP 
22-Apr-85 5 14 NO PUP 
25-Apr-85 3 sy) NO PUP 
02-May-85 7 24 NO PUP 
06-May-85 4 28 NO PUP 
16-May-85 10 38 NO PUP 
17-Jun-85 32 WO NO PUP 
01-Jul-85 14 84 NO PUP 
06-Jul-85 6 ve ~89 NO PUP 
22-Aug-85 47 136 NO PUP 
07-Sep-85 16 152 NO PUP 
14-Nov-85 68 220 PUP 
15-Nov-85 1 221 PUP 
19-Nov-85 4 225 PUP 
11-Dec-85 22 247 PUP 
03-Jan-86 23 270 PUP 
07-Feb-86 35 305 PUP 
28-Apr-86 80 385 NO PUP 
06-Jul-86 69 454 NO PUP 
28-Jul-86 22 476 NO PUP 
23-Sep-86 57 533 PUP 
28-Sep-86 5 538 PUP 
27-Oct-86 29 567 PUP 
01-Nov-86 5 572 PUP 


06-Nov-86 5 577 NO PUP 


SEA OTTER 22 


DATES DAYS CUM STATUS 
BETWEEN DAYS 
12-Apr-85 NO PUP 
20-Apr-85 8 8 NO PUP 
09-May-85 19 27 NO PUP 
10-May-85 1 28 NO PUP 
12-May-85 2 30 NO PUP 
13-May-85 1 a NO PUP 
15-May-85 2 33 NO PUP 
16-May-85 1 34 NO PUP 
24-May-85 8 42 NO PUP 
08-Jun-85 15 57 NO PUP 
10-Jun=85 2 59 NO PUP 
12-Jun-85 2 61 NO PUP 
13-Jun-85 it 62 NO PUP 
17-Jun-85 4 66 NO PUP 
26-Jun-85 9 75 NO PUP 
29-Jun-85 3 78 NO PUP 
04-Jul-85 5 83 NO PUP 
06-Jul-85 2 85 NO PUP 
21-Jul-85 15 100 NO PUP 
22-Jul-85 i 101 NO PUP 
28-Jul-85 6 107 NO PUP 
29-Jul-85 1 108 NO PUP 
04-Aug-85 6 114 NO PUP 
16-Aug-85 12 126 NO PUP 
17-Aug-85 1 27) NO PUP 
25-Aug-85 8 135 NO PUP 
31-Aug-85 6 141 NO PUP 
10-Oct-85 40 181 NO PUP 
13-Nov-85 34 215 NO PUP 
30-Nov-85 17 232 NO PUP 
03=Dec-85 3 235 NO PUP 
12-Dec-85 9 244 NO PUP 
15-Jan-86 34 278 NO PUP 
16-Jan-86 st 279 NO PUP 
22-Jan-86 6 285 NO PUP 
10-Feb-86 19 304 NO PUP 
19-Feb-86 9 313 NO PUP 
01-Jun-86 102 415 NO PUP 
09-Sep-86 100 515 NO PUP 
22-Oct-86 43 558 NO PUP 


12-Nov-86 21 579 NO PUP 


288 


SEA OTTER 25 


DATES 


20-Apr-85 
22-Apr-85 
23-Apr-85 
28-Apr-85 
01-May-85 
06-May-85 
07-May-85 
11-May-85 
16-May-85 
17-May-85 
28-May-85 
10-Jun-85 
23-Jun-85 
25-Jun-85 
29-Jun-85 
01-Jul-85 
02-Jul-85 
03-Jul-85 
04-Jul-85 
0-Jul-85 

29-Jul-85 
31-Jul-85 
01-Aug-85 
05-Aug-85 
15-Aug-85 
16-Aug-85 
17-Aug-85 
18-Aug-85 
19-Aug-85 
19-Aug-85 
20-Aug-85 
21-Aug-85 
22-Aug-85 
23-Aug-85 
24-Aug-85 
27-Aug-85 
31-Aug-85 
01-Sep-85 
02-Sep-85 
03-Sep-85 
04-Sep-85 
04-Sep-85 
05-Sep-85 
06-Sep-85 
07-Sep-85 
10-Sep-85 
12-Sep-85 
15-Sep-85 
18-Sep-85 


DAYS 
BETWEEN 


r 
PRPUPRPUWUPRN 


PR 
WW 


WWNWPPRPOPPPPRUPPPRPRPOPPPRPOBPNUDPPERPNEAN 


DAYS 


STATUS 


NO PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 


SEA OTTER 25 (cont.) 


DATES 


24-Sep-85 
25-Sep-85 
28-Sep-85 
30-Sep-85 
03-Oct-85 
08-Oct-85 
10-Oct-85 
14-Oct-85 
15-Oct-85 
23-Oct-85 
29-Oct-85 
30-Oct=85 
31-Oct=85 
01-Nov-85 
07-Nov-85 
11-Nov-85 
12-Nov-85 
15-Nov-85 
20-Nov-85 
29-Nov-85 
04=Dec-85 
08=Dec-85 
14-Dec=85 
15=Dec=85 
27=Dec=85 
03-Jan-86 
o08s=Jan-86 
13-Jan-86 
15-Jan-86 
18-Jan-86 
19-Jan-86 
23-Jan-86 
04-Feb-86 
08-Feb-86 
11-Feb-86 
13-Feb-86 
15-Feb-86 
19-Feb-86 
01-Mar-86 
06-Mar-86 
10-Mar-86 
12-Mar-86 
13-Mar-86 
17-Mar-86 
24-Mar-86 
25-Mar-86 
10-Apr-86 
28-Apr-86 
29-Apr-86 


POPPRPAOPARNUWNUWE OV 


a) 


a 


kr 


PR 
POHDPIPPNARUOANNWEN PP WNUNINE RHE UOUWP 


CUM 


DAYS 


157 
158 
161 
163 
166 
171 
173 
177 
178 
186 
192 
193 
194 
195 
201 
205 
206 
209 
214 
223 
228 
232 
238 
239 
251 
258 
263 
268 
270 
273 
274 
278 
290 
294 
297 
299 
301 
305 
315 
320 
324 
326 
327 
331 
338 
339 
355 
373 
374 


290 


PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 
PUP 


SEA OTTER 25 (cont.) 


DATES 


31-May-86 
10-Jun-86 
24-Jun-86 
05-Jul-86 
07-Jul-86 
06-Aug-86 
08-Aug-86 
12-Sep-86 
21-Sep-86 
03-Oct-86 
16-Oct-86 


DAYS 
BETWEEN 


CUM 


DAYS 


406 
416 
430 
441 
443 
473 
475 
510 
519 
531 
544 


291 


SEA OTTER 27 


DATES 


04-Oct-85 
07-Oct-85 
09-Oct-85 
11-Oct-85 
16-Oct-85 
17-Oct-85 
23-Oct-85 
28-Oct-85 
30-Oct-85 
31-Oct-85 
11-Nov-85 
1Li-Nov-85 
16-Nov-85 
29-Nov-85 
10-Dec-85 
14-Dec-85 
15-Dec-85 
20-Dec=85 
27-Dec-85 
28-Dec-85 
30=-Dec-85 
03-Jan-86 
08-Jan-86 
09-Jan-86 
15-Jan-86 
18-Jan-86 
22-Jan-86 
23-Jan-86 
28-Jan-86 
08-Feb-86 
11-Feb-86 
13-Feb-86 
06-Mar=86 
10-Mar-86 
12-Mar-86 
24-Mar-86 
02-Apr-86 
18-Apr-86 
29-Apr-86 
05-May-86 
23-May-86 
18-Jun-86 
24-Jun-86 
05-Jul-86 
28-Jul-86 
29-Aug-86 
17-Sep-86 
27-Sep-86 
28-Sep-86 


DAYS 
BETWEEN 


PR 
PNPIUP ERP WUOPPNUDPUNN WwW 


N Rr 
WONNPFPRPNWRUP HE WOAP YU 


rR 


PR 
PO 


SEA OTTER 27 


DATES 


04-Nov-86 
07-Nov-86 
12-Nov-86 
28-Mar-87 


(cont. ) 


DAYS 
BETWEEN 


37 
3 

5 
136 


CUM 


DAYS 


396 
399 
404 
540 


293 


STATUS 


NO PUP 
NO PUP 
NO PUP 
NO PUP 


SEA OTTER 31 


DATES 


16-Oct-85 
21-Oct-85 
23-Oct-85 
28-Oct-85 
30-Oct-85 
07-Nov-85 
15-Nov-85 
16-Nov-85 
10-Dec=-85 
12-Dec-85 
20-Dec-85 
28-Dec-85 
03-Jan-86 
21-Jan-86 
23-Jan-86 
24-Jan-86 
28-Jan-86 
05-Feb=-86 
08-Feb-86 
10-Feb-86 
13-Feb-86 
15-Feb-86 
19-Feb-86 
21-Feb-86 
01-Mar-86 
05-Mar-86 
06-Mar-86 
10-Mar-86 
12-Mar=86 
27-Mar-86 
01-Apr=86 
06-Apr-86 
10-Apr-86 
18-Apr-86 
24-Apr-86 
29-Jun-86 
08-Aug-86 
29-Aug-86 
12-Sep-86 
09-Oct-86 
23-Oct-86 
08-Jan-87 
05-Feb-87 
02-Mar-87 
16-Mar-87 
18-Mar-87 
07-Apr-87 
28-Apr-87 


DAYS 
BETWEEN 


iS) 


ray 
AOR UNUNUNFLRPRPOAONPMENWNWORPNADDANAPHPAODONUN WU 


ra 


294 


SEA OTTER 33 


DATES 


03-Jul-84 
23-Oct-85 
29-Oct-85 
30-Oct-85 
31-Oct-85 
02-Nov-85 
07-Nov-85 
12-Nov-85 
14-Nov-85 
17-Nov-85 
19-Nov-85 
20-Nov-85 
29-Nov-85 
08-Dec-85 
12-Dec-85 
14-Dec-85 
03-Jan-86 
10-Jan-86 
13-Jan-86 
22-Jan-86 
23-Jan-86 
24-Jan-86 
25-Jan-86 
28-Jan-86 
07-Feb-86 
09-Feb-86 
11-Feb-86 
13-Feb-86 
15-Feb-86 
19-Feb-86 
23-Feb-86 
24-Feb-86 
01-Mar-86 
08-Mar-86 
10-Mar-86 
12-Mar-86 
13-Mar-86 
25-Mar-86 
24-Apr-86 
10-Jun-86 
24-Jun-86 
06-Aug-86 
08-Aug-86 
29-Aug-86 
03-Oct-86 
01-Nov-86 
17-Dec-86 
23-Mar-87 
28-Mar-87 


DAYS 
BETWEEN 


PNNAIUONRPPRPNNNNOWRPRPRPUOWIONRPUOUPRPNWNUUNEFE O 


295 


STATUS 


NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 


NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 
NO PUP 


SEA OTTER 33 (cont.) 


DATES 


05-Apr-87 
24-Apr-87 


DAYS 
BETWEEN 


8 
19 


529 
548 


296 


STATUS 


PUP 
NO PUP 


SEA OTTER 36 


DATES DAYS CUM STATUS 
BETWEEN DAYS 

18-Nov-85 NO PUP 
19-Nov-85 1 1 NO PUP 
22-Nov-85 3 4 NO PUP 
25-Nov-85 3 7 NO PUP 
06-Dec-85 11 18 NO PUP 
09-Dec-85 3 21 NO PUP 
20-Dec-85 11 32 NO PUP 
15-Jan-86 26 58 NO PUP 
21-Jan-86 6 64 NO PUP 
05-Feb-86 15 79 NO PUP 
11-Feb-86 6 85 NO PUP 
27-Feb-86 16 101 NO PUP 
08-Apr-86 40 141 NO PUP 
14-Apr-86 6 147 NO PUP 
21-May-86 37 184 PUP 
22-May-86 1 185 PUP 
06-Jun-86 15 200 PUP 
11-Jun-86 5 205 PUP 
13-Jun-86 2 207 PUP 
16-Jun-86 3 210 PUP DEAD 
18-Jun-86 2 212 PUP GONE 
19-Jun-86 1 213 NO PUP 
16-Jul-86 27 240 NO PUP 
17-Jul-86 1 241 NO PUP 
22-Jul-86 5 246 NO PUP 
01-Aug-86 10 256 NO PUP 
08-Aug-86 7 263 NO PUP 
15-Aug-86 7 270 NO PUP 
29-Aug-86 14 284 NO PUP 
04-Sep-86 6 290 NO PUP 
10-Sep-86 6 296 NO PUP 
12-Sep-86 2 298 NO PUP 
03-Oct-86 21 319 NO PUP 
07-Oct-86 4 323 NO PUP 
10-Nov-86 34 357 NO PUP 
16-Nov-86 6 363 NO PUP 
18-Nov-86 2 365 NO PUP 
25-Nov-86 7 372 NO PUP 
30-Nov-86 5 377 NO PUP 
12-Dec-86 12 389 NO PUP 
17-Dec-86 5 394 NO PUP 
24-Dec-86 7 401 NO PUP 
03-Mar-87 69 470 NO PUP 
20-Mar-87 17 487 PUP 
25-Mar-87 5 492 NO PUP 


Appendix 2.2 Data on instrumented otters used for estimating survival 
rates. 
OTTER SEX CAPTURE LAST DATE STATUS DAYS LAST DATE 
NUMBER DATE TX HEARD TRANSMITTING RECOGNIZED 
BY TAGS 
1 AM 0O7-MAR-84& 28-MAY-85 HISSING 447 
2 AM 16-MAR-84 15-SEP-84 MISSING 183 
3 AM 21-MAR-84 28-AUG-85 TX EXPIRED 525 
4 AM 21-MAR-84 28-AUG-85 TX EXPIRED 523 26-Nov-86 
6 AF 03-JUL-84 23-JUN-85 TX EXPIRED 355 
7 AM 15-FEB-85 10-NOV-86 MISSING 633 
9 AF 01-MAR-85 27-AUG-86 MISSING 544 
10 AM 01-MAR-85 12-NOV-86 MISSING 621 
11 AF 15-MAR-85 26-NOV-86 MISSING 621 
13 JM 16-MAR-85 O05-JAN-87 MISSING 660 11-Aug-87 
14 AF 16-MAR-85 30-MAR-87 MISSING 744 
15 AF 20-MAR-85 11-DEC-86 TX EXPIRED 631 06-Mar-87 
16 AF O3-APR-85 02-DEC-86 MISSING 608 
17 AM 0O3-APR-85 28-MAR-86 MISSING 359 
19 AF O3-APR-85 11-NOV-86 MISSING 587 
21 AF 10-APR-85 13-APR-85 MISSING 3 
22 AF 10-APR-85 16-NOV-86 MISSING 585 
23 AM 10-APR-85 27-APR-85 DEAD 17 
25 AF 13-APR-85 20-OCT-86 MISSING 555 
26 AF O8-MAY-85 O05-JUN-85 DEAD 28 
27 AF 04-OCT-85 31-OCT-86 TX EXPIRED 392 28-Mar-87 
28 AF 04-G€T-85 04-NOV-85 TX FAILED 31 
29 JF 11-OCT-85 14-MAR-87 MISSING 519 
30 JM 11-OCT-85 29-JUL-87 MISSING 656 
31 AF 11-OCT-85 24-JUL-87 MISSING 651 
33 AF 18-OCT-85 26-SEP-87 MISSING 708 
34 AM 19-OCT-85 23-SEP-86 MISSING 339 
35 JM O8-NOV-85 21-MAR-87 MISSING 498 
36 AF O08-NOV-85 07-OCT-87 MISSING 698 
37 JF 22-NOV-85 17-OCT-86 MISSING 329 
38 JF 22-NOV-85 O02-JAN-86 MISSING 4j 
39 JF 22-NOV-85 25-MAR-87 MISSING 488 
40 JF 17-DEC-85 03-DEC-87 TRANSMITTING 716 
41 JM 17-DEC-85 13-APR-87 DEAD 482 
42 JF 17-DEC-85 09-OCT-87 MISSING 661 
43 JM 18-DEC-85 10-NOV-87 MISSING 692 
44 JF 18-DEC-85 29-JUN-86 DEAD 193 
45 JF 18-DEC-85 22-MAR-87 MISSING 459 
46 JF 18-DEC-85 25-DEC-87 TRANSMITTING 737 
47 JF 30-DEC-85 28-DEC-87 TRANSMITTING 728 


298 


Appendix 2.3 -- Tag loss information for instrumented sea otters in 

California as of 

10 July 1987. 

OTTER DATE RIGHT TAG LEFT TAG DAYS FROM TAGGING 

NUMBER TAGGED LAST SEEN MISSING LAST SEEN MISSING TO DATE LAST SEEN 
OR MISSING 


RIGHT TAG LEFT TAG 


1 O7-MAR-84 28-MAY-85 28-MAY-85 447 447 
2 16-MAR-84 15-SEP-84 15-SEP-85 183 183 
3 21-MAR-84 28-AUG-85 28-AUG-85 525 525 
4 21-MAR-84 26-NOV-86 26-NOV-86 980 980 
6 O3-JUL-84 21-FEB-85 06-MAR-85 21-JUN-85 246 353 
7 15-FEB-85 06-AUG-86 06-AUG-85 537 537 
9 01-MAR-85 07-OCT-85 23-OCT-85 05-FEB-86 29-APR-86 236 424 
10 0O1-MAR-85 14-NOV-85 06-APR-86 11-SEP-85 14-NOV-85 401 258 
11 15-MAR-85 07-SEP-85 07-SEP-85 176 176 
13° 16-MAR-85 31-AUG-85 31-AUG-85 168 168 
14° 16-MAR-85 25-SEP-85 15-NOV-85 29-JUN-85 24-JUL-85 244 130 
15  20-MAR-85 06-MAR-87 06-MAR- 87 716 716 
16 O3-APR-85 16-AUG-85 11-SEP-85 16-AUG-85 11-SEP-85 161 161 
17 O3-APR-85 11-NOV-85 11-NOV-85 222 222 
19 0O3-APR-85 01-NOV-86 11-SEP-85 11-DEC-85 577 252 
21 10-APR-85 13-APR-85 13-APR-85 3 3 
22 10-APR-85 22-OCT-86 22-0cT-85 560 560 
23. 10-APR-85 27-APR-85 27-APR-85 17 17 
25  13-APR-85 21-SEP-86 03-OCT-86 24-JUN-86 06-AUG-86 538 480 
26 O08-MAY-85 06-JUN-85 06-JUN-85 29 29 
27 04-OCT-85 28-MAR-87 28-MAR-87 540 540 
28 04-0CT-85 15-OCT-85 15-OCT-85 11 11 
29 11-OCT-85 30-OCT-85 11-FEB-86 30-OCT-85 11-FEB-86 123 123 
30 11-OCT-85 30-MAY-87 30-MAY-85 596 596 
31 11-OCT-85 18-APR-86 02-MAY-86 16-MAY-87 203 582 
33. 18-OCT-85 07-JUL-87 07-JUL-87 627 627 
34 19-OCT-85 13-AUG- 86 06-JUL-86 13-AUG-86 298 298 
35 O8-NOV-85 03-MAR-87 03-MAR-87 480 480 
36 O08-NOV-85 01-JUL-87 06-JUL-87 600 605 
37 22-NOV-85 17-JUL-86 17-JUL-86 237 237 
38 22-NOV-85 18-DEC-85 18-DEC-85 26 26 
39 22-NOV-85 28-MAR-87 28-MAR- 87 491 491 
40 17-DEC-85 06-JUN-86 17-JUL-86 09-JUL-87 212 569 
41 17-DEC-85 17-OCT-86 17-OCT-86 304 304 
42 17-DEC-85 25-JUN-86 17-JUL-86 25-JUN-86 17-JUL-86 212 212 
43 18-DEC-85 28-JAN-86 27-FEB-87 11-APR-87 41 479 
44 18-DEC-85 04-JUL-86 04-JUL- 86 198 198 
45 18-DEC-85 11-JUN-86 17-JUL-86 17-JAN-87 211 395 
46 18-DEC-85 09-NOV-86 29-MAR-87 09-NOV-86 29-MAR-87 466 466 
47 30-DEC-85 01-JUL-87 01-JUL-87 548 548 


APPENDIX 3.1 -- Analysis of variance for the distance 


between successive locations of individual instrumented sea 
Log transformed data, base 2. 


otters. 


A. LOCATIONS 18-36 HOURS APART 


Adult females 
Among individuals 
Error 


Juvenile females 
Among individuals 
Error 


Adult males 
Among individuals 
Error 


Juvenile males 
Among individuals 
Error 


Age/sex classes 
Among classes 
Error 


df MS F 
12 20.1 37.4 
3450 0.537 
9 8.8 15.8 
1918 0.557 
7 1.6 4.0 
1401 0.398 
4 5.0 6.8 
964 0.735 
3 185.6 342.2 
7733 0.542 


B. LOCATIONS MORE THAN 36 HOURS APART 


Adult females 
Among individuals 
Error 


Juvenile females 
Among individuals 
Error 


Adult males 
Among individuals 
Error 


Juvenile males 
Among individuals 
Error 


Age/sex classes 
Among classes 
Error 


df MS F 
12 15.8 18.9 
1441 0.837 
9 8.1 7.6 
913 1.1 
7 33.4 22.1 
635 1.5 
4 1.6 1.0 
544 1.6 
3 131.7 94.1 
3533 1.4 


300 


Pp 
<0.001 


<0.001 


<0.001 


<0.001 


<0.001 


Pp 


<0.001 


<0.001 


<0.001 


ns 


<0.001 


APPENDIX 3.2 -- Analysis of variance for the minimum convex 
polygon daily home ranges. Log transformed data, base 2. 


df MS F p 
Adult females 
Among individuals 4 Seis 5.91 ns 
Error 4 1.41 
Juvenile females 
Among individuals 5 3.04 vA ns 
Error 5 1.10 
Adult males sample size too small for testing 
Juvenile males 
Among individuals 3 1.90 0.39 ns 
Error 7 3.41 4.87 


Age/sex classes after otters 
Among classes 2 25.95 28.51 <0.001 
Error 16 0.91 


301 


APPENDIX 3.3 -- Analysis of variance for seasonal 
differences in monthly harmonic mean home range size. Log 
transformed data, base 2. 


af MS F p 

Adult females 
Among individuals a2. 17.08 7.98 <0.001 
Season 1 0.40 0.19 ns 
Individualxseason 12 0.94 0.44 ns 
Error 227 2.14 

Juvenile females 
Among individuals 8 2.36 1.78 ns 
Season 1 1.50 1.13 ns 
Individualxseason 8 0.98 0.73 ns 
Error 137 1.33 

Adult males 
Among individuals 7 9.11 6.70 <0.001 
Season 1 0.10 0.07 ns 
Individualxseason 7 Nog Bal, 0.89 ns 
Error 96 1.36 

Juvenile males 
Among individuals 4 5.88 2.47 ns 
Season 1 5.80 2.43 ns 
Individualxseason 4 1.60 0.67 ns 
Error 719 2.38 


302 


APPENDIX 3.4 -- Analysis of variance for differences in 
monthly harmonic mean home range size among age/sex classes. 
Log transformed data, base 2. 


af MS F p 
Adult females 
Among individuals 12 17.08 8.29 <0.001 
Error 240 2.06 
Juvenile females 
Among individuals 8 2.45 N67) ns 
Error 145 0.557 
Adult males 
Among individuals a 8.47 5.92 <0.001 
Error 105 1.43 
Juvenile males 
Among individuals 4 5.89 2.47 ns 
Error 84 2.38 
Age/sex classes after otters 
Among classes 3 294.87 163.8 <0.001 
Error 574 1.80 


*This test does not include seasonal effects. The only 
age/sex group with significant seasonal effects was the 
juvenile females (see appendix 3.4). Variation due to 
season is included in the error term; thus, this test is 
conservative. 


303 


APPENDIX 3.5 - Analysis of variance for the distance between 
extreme locations of individual instrumented sea otters. 
Log transformed data, base 2. 


fob a MS F p 
Adult females 
Among individuals 12 19.8 22.56 <.001 
Error 240 0.88 
Juvenile females 
Among individuals 8 9.81 4.54 <.001 
Error 146 2.16 
Adult males 
Among individuals 7 10.13 3.58 <2 801. 
Error ~ 104 2.83 
Juvenile males 
Among individuals 4 2.67 1.68 ns 
Error 84 1.59 
Age/Sex classes 
Among classes 3 90.63 54.6 <.001 
Error 574 1.66 


APPENDIX 4.1 - TWENTY-FOUR-HOUR DATA BY OBSERVATION PERIOD. 


Fa ee IE ———EEE EE 
OTTER AGE/ DATE LENGTH NUMBER OF 10-MIN PERIODS 
NUMBER _ SEX HRS REST FEED OTHER 
6 AF 19-Jul-84 24 81 27 38 
6 AF 25-Jul-84. 48 145 101 62 
6 AF 07-Aug-84 24 64 66 20 
6 AF 21-Aug-84 14 39 32 14 
6 AF 31-Aug-84 24 60 58 33 
6 AF 05-Sep-84 17 35 57 15 
7 AM 05-Sep-85 24 64 71 12 
7 AM 08-Oct-85 24 105 20 20 
7 AM 06-Aug-86 24 77 53 13 
9 AFP 23-Jul-85 24 49 69 27 
9 AF 26-Aug-85 30 73 85 . 22 
10 AM 10-Sep-85 23 37 mS 17 
11 AF 05-Sep-85 24 64 58 23 
11 AF 08-Oct-85 24 61 54 28 
13 JM 24-Aug-86 24 65 45 38 
14 AFP 24-Jul-85 24 32 86 25 
14 AFP 27-Aug-85 18 51 44 12 
15 AF 20-May-85 24 79 52 15 
15 AF 30-May-85 23 68 61 6 
15 AF 03-Jun-85 24 75 45 25 
15 AF 18-Jul-85 24 75 45 26 
15 AF 12-Aug-85 72 229 128 76 
16 AF 19-May-85 24 66 52 26 
16 AF 28-May-85 48 147 98 44 
16 AF 10-Jul-85 24 75 32 37 
16 AF 10-Sep-85 23 66 54 15 
16 AF 08-Oct-85 24 75 45 18 
16 AFP 04-Nov-86 48 137 104 46 
17 AM 02-Jul-85 48 120 105 67 
17 AM 19-Jul-85 24 69 69 7 
19 AF 19-May-85 23 54 59 24 
19 AF 28-May-85 47 130 117 36 
19 AF 04-Nov-86 48 173 val 46 
19 AF 10-Sep-85 23 62 65 7 
19 AF 08-Oct-85 24 94 44 8 
22 AF 30-Jul-85 48 166 103 20 
22 AF 05-Sep-85 24 77 39 29 
22 AF 08-Oct-85 24 71 49 20 
25 AFP 07-Aug-85 48 159 152 45 
25 AFP 28-Aug-85 11 35 16 ALS) 
27 AFP 13-May-85 48 138 97 54 
29 JF 25-Jun-86 24 40 69 36 
30 JM 23-Apr-86 48 113 98 80 
34 AM 30-Jul-86 24 88 24 32 
34 AM 13-Aug-86 24 70 60 15 
35 JM 25-Feb-86 50 87 116 89 
35 JM 16-Sep-86 48 66 109 114 


36 AF 13-Mar-86 24 43 85 17 


305 


APPENDIX 4.1 (cont.) 


Se Sun ee eee ee ee ee ee EE EE ee 
OTTER AGE/ DATE LENGTH NUMBER OF 10-MIN PERIODS 
NUMBER SEX (HRS) REST FEED OTHER 

36 AFP 21-May-86 48 172 86 33 

37 JF 06-Mar-86 24 55 85 6 

39 JF 25-Feb-86 33 69 75 53 

39 JF 29-Apr-86 24 48 84 14 

40 JF 03-Apr-86 48 103 157 29 

40 JF 20-Aug-86 24 90 35 20 

41 JM 13-Mar-86 24 40 81 23 

41 JM 20-Aug-86 24 70 45 29 

41 JM 24-Sep-86 48 73 102 115 

42 JF 16-Apr-86 49 128 134 30 

43 JM 20-Aug-86 24 50 44 50 

44 JF 19-Mar-86 48 124 136 30 

45 JF 13-Nov-86 48 87 173 27 

46 JF 06-Mar-86 21 76 35 12 

46 JF 30-Apr-86 24 61 77 6 


47 JF 19-Nov-86 48 74 173 47 


306 


APPENDIX 4.2 


OTTER 
NUMBER . 


WOUNNNYANAAAAAG 


AGE/SEX 


DATE 


19-Jul-84 
25-Jul-84 
07-Aug-84 
21-Aug-84 
31-Aug-84 
05-Sep-84 
05-Sep-85 
08-Oct-85 
06-Aug-86 
23-Jul-85 
26-Aug-85 
10-Sep-85 
05-Sep-85 
08-Oct-85 
24-Aug-86 
24-Jul-85 
27-Aug-85 
20-May-85 
30-May-85 
03-Jun-85 
18-Jul-85 


DAYLIGHT DATA BY OBSERVATION PERIOD. 


NUMBER OF 10-MIN PERIODS 


REST 


FEED 


OTHER 


APPENDIX 4.2 (cont.) 


OTTER AGE/SEX DATE NUMBER OF 10-MIN PERIODS 
NUMBER REST FEED OTHER 
36 AF 13-Mar-86 20 48 10 
36 AFP 21-May-86 121 43 21 
37 JF 06-Mar-86 13 61 4 
39 JF 25-Feb-86 46 50 48 
39 JF 29-Apr-86 44 38 7 
40 JF 03-Apr-86 39 113 13 
40 JF 20-Aug-86 61 15 14 
41 JM 13-Mar-86 4 53 20 
41 JM 20-Aug-86 42 24 21 
41 JM 24-Sep-86 34 71 54 
42 JF 16-Apr-86 88 53 18 
43 JM 20-Aug-86 30 19 38 
44 JF 19-Mar-86 37 lil 15 
45 JF 13-Nov-86 3 124 11 
46 JF 06-Mar-86 41 14 3 
46 JF 30-Apr-86 38 43 6 
47 JF 19-Nov-86 16 96 27 


APPENDIX 4.3 


OTTER 
NUMBER 


OV OV OV OV OV 


AGE/SEX 


Fy Fey Fey Py ey Fy By Fay Py By ey Be 


DATE 


02-Jul-85 
19-Jul-85 
10-Sep-85 
18-Jul-85 
10-Sep-85 
26-Aug-85 
10-Sep-85 
30-Jul-85 
12-Aug-85 
24-Jul-85 
13-May-85 
21-May-86 
07-Aug-85 
19-May-85 
28-May-85 
19-May-85 
28-May-85 
20-May-85 
30-May-85 
03-Jun-85 
04-Nov-86 
06-Aug-86 
13-Aug-86 
04-Nov-86 
27-Aug-85 
25-Jun-85 
20-Aug-86 
13-Nov-86 
19-Nov-86 
06-Mar-86 
19-Mar-86 
30-Apr-86 
29-Apr-86 
06-Mar-86 
30-Apr-86 
16-Apr-86 
16-Sep-86 
25-Feb-86 
23-Apr-86 
19-Jul-84 
25-Jul-84 
07-Aug-84 
21-Aug-84 
31-Aug-84 
05-Sep-84 


VISUAL .DATA BY OBSERVATION PERIOD. 


NUMBER OF 10-MIN PERIODS 


REST FEED 
41 39 
43 17 

7 ie) 
21 10 
5 5 
1 5 
10 0) 
90 30 
110 32 
10 38 
78 21 
115 2 
71 28 
10 2 
56 28 
0) 0) 
22 6 
34 18 
38 23 
41 14 
11 ie) 
12 9 
ie) 16 
8 17 
11 0) 
4 11 
2 fe) 
0 30 
2 8 
te) al 
17 42 
35 42 
33 7 
3 27 
22 64 
63 0 
(0) 2 
0) 6 
43 1 
37 5 
77 35 
11 25 
10 14 
14 a7 
8 7 


309 


OTHER 


2 


Pr 


rR 


N 
PONRPNUONNYNFPF RP UORPRPUOONONOKHWVIOOUFENN 


APPENDIX 5.1 -- Analysis of variance for the length of dives 
made by individual instrumented sea otters in California. 
Log-transformed data, base 2. ; 


af mean square F p 
Adult females 
Among individuals 4 130.0 281.4 <0.001 
Error 2763 0.46 
Adult females with pups 
Among individuals 3 94.84 144.33 <0.001 
Error 1171 0.66 
Juvenile females . | 
Among individuals 6 76.92 180.52 <0.001 
Error 2129 0.43 
Adult males : 
Among individuals 6 10.85 14.36 <0.001 
Error 1377 0.76 
Juvenile males 
Among individuals 4 7.39 20.04 <0.001 
Error 493 0.37 


310 


APPENDIX 5.2 -- Analysis of variance for the length of the 
surface intervals made by individual instrumented otters. 
Log-transformed data, base 2. 


df Mean square F p 
Adult females 
Among individuals 4 20253 179.96 <0.001 
Error 2656 1.12 
Adult females with pups 
Among individuals 3 Sees 25.42 <0.001 
Error 1089 1.23 
Juvenile females 
Among individuals 6 60.75 42.76 <0.001 
Error 2114 1.42 
Adult males 
Among individuals 6 28.85 18.37 <0.001 
Error 1310 abo i7/ 


Juvenile males 
Among individuals 4 1.38 1.45 ns 
Error 472 0.95 


APPENDIX 5.3 - Analysis of variance for the length of dives 
made during the day and night by the individual instrumented 
otters. 


af mean square F Pp 

Adult females 
Among individuals 4 129.98 302.26 <0.001 
Day/night 1 0.10 0.23 ns 
Day/night x individual 4 19.55 45.46 <0.001 
Error 2758 0.43 

Adult females with pups 
Among individuals 3 94.80 166.37 <0.001 
Day/night 1 43.90 77.01 *<05001 
Day/night x individual 3 20.30 35.61 <0.001 
Error 1167 0.57 

Juvenile males 
Among individuals 4 7.40 21.76 <0.001 
Day/night 1 0.90 2.65 ns 
Day/night x individual 4 4.48 13.16 <0.001 
Error 488 0.34 

Juvenile females 3 
Among individuals 6 76.92 187.60 <0.001 
Day/night 1 8.20 20.00 <0.001 
Day/night x individual 6 4.50 10.98 <0.001 
Error 2122 0.41 


312 


APPENDIX 5.4 - Analysis of variance for the length of 
surface intervals made during the day and night by the 


individual instrumented otters. 


af 
Adult females 
Among individuals 4 
Day/night 1 
Day/night x individual 4 
Error 2651 
Adult females with pups 
Among individuals 3 
Day/night al 
Day/night x individual 3 
Error 1085 
Juvenile males 
Among individuals 4 
Day/night 1 
Day/night x individual 4 
Error 467 
Juvenile females 
Among individuals 6 
Day/night sl 
Day/night x individual 6 
Error 2107 


mean square 


201.33 
1.11 
5.52 
1.11 


31.23 
1.30 
1.67 
1.23 


1.38 
10.80 
9.32 
0.86 


60.75 
24.20 
8.55 
1.39 


F 


<0.001 
<0.001 


<0.001 
<0.001 
<0.001 


APPENDIX 10. 


314 


APPENDIX 10.1A -- Map of central California showing 
California Department of Fish and Game mortal ity recovery 
areas (after Ames, et al., 1983). 


APPENDIX 
California coast. 


zis. Davenport 


10.1B -- Ordinates 


for fathom line 


along the 


t 
N 


5 0 5 10 


—S ee eS 


Kilometers 


316 


Moss 


PE) 


5+) CARMEL 


: 4305 
4 : 
Pt. Lobos 
YS. 
soh 


b=) 


5 0 5 
Sas See er ae 
Kilometers 


317 


Ses 


Pfeiffer Point 


530 


Zap 


Grimes Point 


5 0 5 
[== — | 
Kilometers 


S705) John Little State Park. 


318 


Gorda 


Za! 


5 0 ee 10 


= Se | 
Kilometers 


319 


av 


5 0 5 10 
= = ee Se 
Kilometers 


320 


MORRO BAY 


Montana de Oro 


State Park 


Fa, 


$2041 GROVER CITY 


Santa Maria River 


0 5 ~___IJ@ 
= = ee ae 
Kilometers 


S2iL 


Za, 


0 5 


Kilometers 


10 


322 


APPENDIX 10.2 -- User's Manual for OTPOP: A Simulation Model 
for Assessing the Risks of Oil Spills to the California Sea 
Otter Population. 


USER'S MANUAL 
for 


OTPOP 


A Simulation Model for Assessing the Risks of 
Oil Spills to the California Sea Otter Population 


1987 


University of Minnesota 


323 


APPENDIX 10.2 -- User's Manual for OTPOP: A Simulation Model 
for Assessing the Risks of Oil Spills to the California Sea 
Otter Population. 


324 


I. INTRODUCTION. 


This manual provides information only on running OTPOP, 
it does not explain the logical structure of the model nor 
the significance of the various parameters. Many problems 
will be avoided if the user familiarizes his/her self with 
the model documentation volume before attempting to use the 
progran. 


325 


II. TECHNICAL SPECIFICATIONS 


Hardware. OTPOP is designed to run on an IBM PC,xXT, or 
AT microcomputer. An Intel 8087 or 80287 coprocessor is 
required. A hard disk is recommended. A battery operated 
clock and associated software are necessary for the random 
number generator. 


Software. OTPOP is written in FORTRAN and compiled on 
the Rand-McFarland IBM Professional FORTRAN Compiler version 
1.0. All code is ANSI FORTRAN77 compatible except for 4 IBM 
Professional extensions used extensively throughout the 
program: 1) all variables and array elements are 
automatically set to 0 at the start of the program, 2) some 
COMMON and declaration statements are included separately in 
program units through the use of INCLUDE statements, 3) 
subroutine variables are automatically saved without the use 
of SAVE statements, and 4) most integer variables are 
declared as 2 byte (INTEGER*2) to save memory. The action of 
these extensions must be considered if the program is 
transported to a different compiler. 


The menu screens used to input parameters that are set 
at runtime are generated using K&S Systems Screen Generator 


version 4.7. The memory resident portion of the screen 
generator must be loaded before running OTPOP (this is 
accomplished in the batch file OP.BAT). If OTPOP is 


transported to a different compiler a different screen 
generator interface must be used, or the data entry portion 
of the program rewritten. 


Operating environment. DOS version 3.1 or higher is 
required as an operating system, and at least 384 kilobytes 
of RAM must be available. The following statements must be 
included in the "CONFIG.SYS" file available at boot: 

DEVICE=ANSI.SYS 
BREAK ON 
FILES=25 
BUFFERS=25 


The DOS "ANSI.SYS" file must also reside on the boot disk. 


326 


III. RUNNING THE PROGRAM 


The batch file OT.BAT is supplied to easily load the 
resident portion of the screen generator, run OTPOP, process 
the raw output, and restore the proper MODE. To run OT.BAT 
type "OT" and <enter>. Fig. 1 will briefly appear on the 
screen as the screen generator is loaded, and Fig. 2, 
introducing the program, will appear as OTPOP is loaded. 


The screens pictured in Figs. 3-10 are used to set run 
environment and model parameters at runtime. Default values 
of all parameters automatically appear when the screens are 
presented. To change a default value move the cursor to the 
parameter in question and enter the new value. The program 
automatically checks for parameter values that are out of 
acceptable range or of the wrong type (for instance, entering 
a letter when a number is required, or a number with a 
decimal point when an integer is required). Move back and 
forth between screens using the Fl and F10 function keys as 
noted at the bottom of each screen. 


Following is a description of the parameters that are 
set at runtime using the input screens. 


327 


Figure 1. 


K&S Systems copyright notice for the sceeen n 
generator. 


C: \OTTERS\INIPOP>sgx 

The Screen Generator v4.47 

(C) Copyright 1982,83,84,85 K & S Systems 

(C) Copyright 1986 The West Chester Group 

PO Box 1304, West Chester, PA 19380, (215) 644-4206 


328 


Figure 2. Introductory screen for OTPOP. 


OTPOP 


A SIMULATION MODEL FOR THE ANALYSIS 
OF THE RISK OF OIL SPILLS TO THE 
CALIFORNIA SEA OTTER POPULATION 


FOR 
USDI MINERALS MANAGEMENT SERVICE 


WRITTEN AT THE 
UNIVERSITY OF MINNESOTA 


6 1987 


329 


Screen 1: Run parameters (Poet ELA 


NUMBER OF YEARS PER RUN. Enter the number of years you want 
to simulate after the oil spill. 


NUMBER OF RUNS WITH OIL SPILL. Enter the number of separate 
runs of the number of years specified above you want 
executed. 


NUMBER OF CONTROL RUNS. Enter the number of runs you want 
conducted without introducing an oil spill. The control runs 
run for the same number of years as the runs with oil spills. 


INITIAL POPULATION SIZE. Enter the desired number of 
independent otters in the simulated population at time of the 
spill. Because of the stochasticity built in to the model, 
and because the model runs for 3 simulated years before 
introducing the spill, the number of animals at the time of 
the spill may differ slightly from the inputted value. The 
initial population size will also differ between runs. 


DATE OF OIL SPILL. Enter the month (1-12) and day (1-31) to 
introduce the spill. 


DURATION OF SPILL. Enter the number of days (up to 30) that 
the spill is to affect the population. 


NORTH BOUNDARY OF SPILL. Enter the north boundary of the 
simulated spill using the CDFG 5-fathom ordinate (see 
Appendix A). 


SOUTH BOUNDARY OF SPILL. Enter the south boundary of the 
simulated spill, using the CDFG 5-fathom ordinate. This 
value must be greater than the value for the northern 
boundary entered above. 


330 


Figure 3. OTPOP Screen #1 -- Run parameters. 
Dashed lines indicate location of parameter 
edited by user. 


SET RUN PARAMETERS: 
NUMBER OF YEARS PER RUN 
NUMBER OF RUNS WITH OIL SPILL 
NUMBER OF CONTROL RUNS 
INITIAL POPULATION SIZE 
ER a DATE OF OIL SPILL (MONTH/DAY) 
DURATION OF SPILL 
NORTH BOUNDARY OF SPILL ( <SOUTH ) 


SOUTH BOUNDARY OF SPILL ( >NORTH ) 


Fl 
F10 


PROCEED 
PREVIOUS SCREEN 


331 


values 


Screen 2: Population parameters (Fig. 4): 


EQUILIBRIUM POPULATION SIZE. Enter the carrying capacity, in 
number of independent otters, of the simulated range. The 
program OTRANGE may be run separately to determine this 
value. 


NORTH BOUNDARY OF RANGE. Enter the north boundary of the sea 
otter range at the time of the spill using the CDFG 5-fathom 
ordinate. Program OTRANGE may be run separately to determine 
this value. 


SOUTH BOUNDARY OF RANGE. Enter the south boundary of the sea 
otter range at the time of the spill using the CDFG 5-fathom 
ordinate. This value must be higher than the value entered 
above for the north boundary of the range. Program OTRANGE 
may be run separately to determine this value. 


MAXIMUM PER CAPITA GROWTH RATE. Enter the maximum attainable 
per capita annual growth rate of the population, in 
animals/animal/year. This is "rpay" from equation (3) in the 
documentation volume. 


NON-LINEARITY OF DENSITY DEPENDENCE. Enter the value for "b" 
in equation (3) in the documentation volume. The higher the 
value of "b" the more rectangular the density dependence 
function (see Fig. 4 in the documentation volume). 


DENSITY INDEPENDENT MORTALITY RATE. Enter the mortality 
rate, in animals/animal/year, due to density-independent 
factors. This could be used to simulate incidental gill-net 
mortality, predation, or harvest ("m" in equation (29) in the 
documentation volume). See discussion of density independent 
mortality in the OTRANGE section of the documentation volume. 


DEGREE OF COMPENSATION. Enter the proportion of the density 
independent mortality that will compensate for density 
dependent mortality. See discussion of density independent 
mortality in the OTRANGE section of the documentation volume. 


Cy ke Gd we kd So! EDC ee eee CF TI Ee oD 
Note: Density independent growth can be simulated by setting 
the equilibrium population size very high relative to the 
initial population size (i.e., at least 10 times as high), 
and setting the nonlinearity coefficient very high (i.e., 
0.05 or greater). 

; i i i, i i i CF 


332 


Figure 4. OTPOP Screen #2 -- Population parameters. 
Lines indicate location of parameter values edited by 
user. 


SET POPULATION PARAMETERS: 


EQUILIBRIUM POPULATION SIZE 

NORTH BOUNDARY OF RANGE ( <SOUTH ) 
SOUTH BOUNDARY OF RANGE ( >NORTH ) 
MAXIMUM PER CAPITA ANNUAL GROWTH RATE 
NON-LINEARITY OF DENSITY DEPENDENCE 
DENSITY INDEPENDENT MORTALITY RATE 


DEGREE OF COMPENSATION 


Fl 
F10 


PROCEED 
PREVIOUS SCREEN 


333 


Screen 3: Survival and reproductive parameters (Fig. 5): 


ADULT FEMALE SURVIVAL RATE. Set the annual rate of survival 
of adult females against incidental risks. This is "aj" in 
equation (1) in the documentation volume. Express as a 
proportion. 


MODAL AGE OF FEMALE SENESCENCE. Set the age, in years, of 
the modal age of death due to old age for females. MThis is 
"T" in equation (10) in the documentation volume. 


PRIME REPRODUCTIVE RATE. Set the maximum yearly reproductive 
rate, expressed as weaned females per adult female per year. 
This is "A" in equation (2) in the documentation volume. 


PUP SURVIVAL RATE. Set the proportion of pups that will 
survive from birth until weaning. Intrauterine mortality is 
not considered. 


ADULT MALE SURVIVAL RATE. Set the annual rate of survival of 


adult males against incidental risks. This) ds, ag v eid 
equation (1) in the documentation volume. Express aS a 
proportion. 


MODAL AGE OF MALE SENESCENCE. Set the age, in years, of the 
modal age of death due to old age for males. This is "T" in 
equation (10) in the documentation volume. 


% VARIATION IN ADULT SURVIVAL. Set the relative percent by 
which annual adult survival rates may vary. This variance is 
used only in the recovery phase of the simulation, and the 
distribution of annual survival rates is assumed to be 
uniform between the specified boundaries. This simulates the 
"environmental stochasticity" parameter, "p", described in 
the description of the structure of LESLIE in the 
documentation volume. 


% VARIATION IN PUP SURVIVAL. Set the relative percent by 
which annual pup survival rate may vary. This variance is 
used only in the recovery phase of the simulation, and the 
distribution of annual survival rates is assumed to be 
uniform between the specified boundaries. Percent variation 
in pup survival may differ from percent variation in adult 
survival. 


334 


Figure 5. OTPOP Screen #3 -- survival and reproductive 
parameters. 


Lines indicate location of parameter values edited by 
user. 


SET SURVIVAL & REPRODUCTIVE PARAMETERS: 


ADULT FEMALE SURVIVAL RATE 
MODEL FEMALE AGE OF SENESCENCE 
PRIME REPRODUCTIVE RATE 

PUP SURVIVAL RATE 

ADULT MALE SURVIVAL RATE 


MODEL MALE AGE OF SENESCENCE 


I+ 


PERCENT VARIATION IN ADULT SURVIVAL 


I+ 


PERCENT VARIATION IN PUP SURVIVAL 


Fl 
F10 


PROCEED 
PREVIOUS SCREEN 


335 


Screen 4: Movement parameters (Fig. 6): 


Classes of animals are listed down the left side of the 
screen, movement parameters along the top. AR is the 
autoregressive parameter, CE is the displacement parameter, 
SIGMA is the standard deviation of daily distance moved, and 
VMAX is the maximum possible daily distance moved. See the 
discussion of the structure of OTMOVE and equation (22) in 
the documentation volume. 


336 


Figure 6. OTPOP Screen #4 -- Movement parameters. 
Lines indicate location of parameter values edited by 
user. 


SET OTTER MOVEMENT PARAMETERS: 


AR CE SIGMA VMAX 
JUVENILE FEMALES 
ADULT FEMALES W/PUP 
ADULT FEMALE W/O PUP 
JUVENILE MALES 
ADULT NON-TERRITORIAL 
MALES 
ADULT TERRITORIAL MALES 
Fl PROCEED 


F10 PREVIOUS SCREEN 


337 


Screen 5: Male territorialit arameters (Fig. 7): 


AGE AT WHICH POTENTIALLY TERRITORIAL. Enter the age, in 
years, at which males may hold breeding territories. 


MAXIMUM % OF POTENTIALS THAT HOLD TERRITORIES. Enter the 
percent of potentially territorial males that will hold 
territories at the height of the breeding season. 


MINIMUM % OF POTENTIALS THAT HOLD TERRITORIES. Enter the 
minimum percent of potentially territorial males that will 
hold territories at any time throughout the year. 


MEAN TERRITORY LENGTH. Enter the average length of a male 
territory, measured along the 5-fathom line, in 1/2 km units. 


S. D. OF TERRITORY LENGTH. Enter the standard deviation 
around mean territory length, in 1/2 km units. 


MEAN ARRIVAL DATE. Enter the average date (month (1-12) / 
day (1-31)) of arrival on a territory. 


S. D. OF ARRIVAL DATE. Enter the standard deviation, in 
days, of average arrival date. 


MEAN DEPARTURE DATE. Enter the average date (month (1-12) / 
day (1-31)) of departure from a territory. 


S. D. OF DEPARTURE DATE. Enter the standard deviation, in 
days, of average departure date. 


90 Oo OOD 9 OO OO Oo 8 Oo O to oS 
Note: The seasonality of territorial behavior and migrations 
can be controlled through the standard deviations of arrival 
and departure dates. Large standard deviations lead to less 
pronounced seasonality. See discussion in the migratory 
movements by adult males section of the documentation volume. 


338 


Figure 7. OTPOP Screen #5 -- Male territoriality 
parameters. 


Lines indicate location of parameter values edited by 
user. 


SET MALE TERRITORIALITY PARAMETERS: 


AGE AT WHICH POTENTIALLY TERRITORIAL 
MAXIMUM % OF PCTENTIALS THAT HOLD TERRITORIES 
MINIMUM % OF POTENTIALS THAT HOLD TERRITORIES 
MEAN TERRITORY LENGTH 
S.D. OF TERRITORY LENGTH 

sory fives MEAN ARRIVAL DATE (MONTH/DAY) 

S.D. OF ARRIVAL DATE (IN DAYS) 

n/a MEAN DEPARTURE DATE (MONTH/DAY) 


S.D. OF DEPARTURE DATE (IN DAYS) 


Fl 
F10 


PROCEED 
PREVIOUS SCREEN 


339 


Screen 6: Oil spill response parameters (Fig. 8): 


Columns 1 and 5 show the day of the oil spill, columns 2 and 
6 list the probabilities of mortality after contact with the 
spill for each day, columns 3 and 7 list the probabilities of 
locally avoiding a spill on each day, and columns 4 and 8 
list the probabilities of avoiding a spill by shifting the 
location of the home range on each day of the spill. 
Parameters for days greater than the duration of spill set in 
screen 1 are ignored by the program. See discussion of the 
structure of OTMOVE in the documentation volume. 


340 


Figure 8. OTPOP Screen #6 -- Oil spill response 
parameters. Lines indicate location of parameter values 
edited by user. 


SET DAILY PROBABILITIES OF MORTALITY, AVOIDANCE, AND 
EMIGRATION DURING EXPOSURE TO OIL SPILL: 


DAY P(MORT) P(AVOID) P(EMIG) ) DAY P(MORT) P(AVOID) P(EMIG) 
al 16 
2 17 
3 18 
4 19 
5 20 
6 ial 
7 22 
8 23 
9 24 
10 25 
int 26 
17 27 
13 28 
14 29 
15 30 
Fl = PROCEED 
F10 = PREVIOUS SCREEN 


341 


Screen 7: RUNID (Fig. 9): 


Enter character string of up to 6 characters that will be 
used to identify the output from the current model run. This 
identification string will appear at the top of the .LOG file 
and on output from PROC. Different RUNID strings should be 
used for every production run of the model to ensure that 
output\ files from a particular run may be permanently 
associated with the appropriate .LOG file. 


342 


Figure 9. OTPOP Screen #7 -- Set run identification string. 
Lines indicate location of parameter values edited by user. 


ENTER SIX CHARACTER RUN IDENTIFICATION STRING: 


Fl 
F10 


PROCEED 
PREVIOUS SCREEN 


343 


Screen 8: Set seed for random number generator Fig. 10): 


Toggle back and forth between "Use constants" and "Use clock" 
with the cursor arrow keys. The same integers will be used 
as random number seeds on every run that is initiated by 
using constants -- this is supplied as a testing or debugging 
aid as it ensures that the same sequence of random numbers 
will be used in each run. The clock should be used to set 
the random number seeds for production runs. 


344 


Figure 10. OTPOP Screen #8 -- Set random number generator 
seed. 


SET SEED FOR PSUEDO-RANDOM NUMBER GENERATOR: 


= SELECT | 


F1l = PROCEED 


345 


After all screens have been examined and parameters set 
the model will begin execution. Parameter values used in the 
current run will be recorded in the OTPOP.LOG file once the 
model begins execution. To terminate the program prematurely 
at any time press <control> and <break> simultaneously (since 
the program will break only during input or output operations 
it may be several seconds to a few minutes before the 
program terminates after pressing <control><break>). 


346 


IV. REQUIRED FILES. 

Several files are required to reside on the same disk 
and subdirectory for the program to function correctly. They 
are: 


MAINPOP. EXE (OTPOP program execution module) 

PROC. EXE (Data processing execution module) 

SGX. EXE (Loads memory-resident portion of 
screen generator) 

MODE. COM (Restores correct mode after 
execution) 

OP.BAT (Batch file to execute above 
programs) 

PARASC (Screen generator library) 

PROFORT. ERR (Error messages for IBM Professional 
FORTRAN) 

ZSCORES . DAT 

DFLT.DAT 

CDIST.DAT 

SR.DAT 

SBST.DAT 


Files with the .DAT extension are data files that are read in 
during program execution. ZSCORES.DAT contains values of the 
standard normal distribution and should not be disturbed. 
The remainder of the .DAT files may be edited by the user 
using a standard word processor. If these files are edited, 
the edited files must be saved as ASCII text files under 
their original names. 


DFLT.DAT contains the default values of the parameters 
set at runtime. Parameter values are in list format, with 1 
space separating each value. The values are listed in the 
order they appear on the screens (Fig. 11). 


CDST.DAT contains the density functions for summer and 
winter used in the distribution algorithms of the model. 
They should be derived from the most recent USFWS/CDFG census 
data. The file consists of 5 columns in list format, with a 
space separating the columns (Fig. 12). The first column is 
an integer representing each 1/2 kilometer segment of the 5- 
fathom ordinate system. The second column is the proportion 
of the population that was observed in that segment of the 
range during the most recent spring/summer census. The third 
column indicates the method by which that section of the 
coast was counted ("0" = from ground, "1" = by air) for that 
census. The fourth column is the proportion of the 
population that was observed in that segment of the range 
during the most recent fall/winter census. The fifth column 
indicates the method by which that section of the coast was 
counted ("0" = from ground, "1" = by air) for that census. 


The digitizing program used by USFWS to enter the census 
data will produce files in the correct format that can be 


347 


Figure 11. Part of the DFLT.DAT file. 


aL 0 as es as B= 0K 0 as ea ie Bs 0 a Ee hoy=) 
1500 201 955 0.0900 0.0050 0.0000 1.0000 
0.9300 15.000 0.2500 0.5300 0.8700 9.0000 5.0 5.0 


0.36700 
-0.0250 
-0.0090 
-0.0450 
00.1050 
00.0420 


-0.1630 
-0.4060 
-0.7060 
-0.2900 
-0.8150 
-1.0440 


8.09000 
6.39000 
2.95000 
8.56000 
4.64000 
1.93000 


37.5000 
37.5000 
08.4000 
48.9000 
48.9000 
48.9000 


60 20 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 
1.0000 


8.1100 0.4400 5 23 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 
0.0000 0.0000 


2 LS 


OBDIHMNFPWNEF OV 


348 


Figure 12. Part of the CDST.DAT file. 


ro 
wo 

° 
ro 

AS 
or 
Pe 


PR 


1 0.00039 1 


(o) 
PRR. 
° 


N 
(o} 
0 
ooo 
oooooo0oo0oo0oo0o0o0o0o0°o°coeo°c0o0o090d 


(2) 
(o} 


Ha tan a WeaR ag Cag ay btn ea eae et eae Nea 
| ed 
° 
RP 


N 
Pp 
a 
(2) 


N 

rR 

NI 
foMo Mo Mo MoM Moone K-) 
PRPPPPPPPPPRP-. 


{o) 


N 
N 
Nu 
fo) 
Pr 
rh 
oO 
Pr 


IN) 
rN) 
as 
ooo 
PRP. 


349 


copied onto the appropriate disk after each census to keep 
the model input data updated. 


SR.DAT contains sex ratio data for the sea otter range, 
used in the distribution algorithms of the model. Sex ratios 


are recorded by CDFG carcass recovery area. The file 
consists of 3 columns, in list format, with spaces separating 
the columns (Fig. 13). The first column contains an integer 


representing the northern boundary of the segment in the 5- 
fathom line ordinate system, the second column contains the 
number of males per female in the segment during the 
spring/summer, the third column contains the number of males 
per female in the segment during fall/winter. 


SBST.DAT contains codes for the substrate along the 
coast. Substrate is either rock or sand. The file consists 
of 2 columns, in list format, single spaces separating the 
columns (Fig. 14). The first column is an _ integer 
representing a point along the 5-fathom line ordinate system 
where the substrate changes from one to the other, the second 
column represents the substrate of the area south of the 
point to the next point where the substrate changes. A "1" 
represents rock substrate, a "0" represents sand substrate. 
This file was coded from U.S. Geological Survey topographic 
maps, and should not have to be changed. 


350 


Figure 13. Part of the SR.DAT file. 


201 4.5 16.7 
216 4.5 16.7 
229 4.5 16.7 
256 4.5 16.7 
271 4.5 16.7 
321 al 3 ; 
366 0.5 0.88 
377 0.5 0.88 
334 0.5 0.88 
390 0.33 0.33 
400 0.33 0233 
411 0.43 0.43 
430 0.43 0.43 
449 0.5 0.44 
473 0.5 0.44 
499 0.5 0.44 
524 0.5 0.44 
599 0.5 0.44 
640 0.41 0.37 
668 0.41 0.37 
694 0.41 0.37 
730 0.33 0.30 
753 0.33 0.30 
787 0.82 0.61 
806 0.82 0.61 
825 0.82 0.61 
844 lO 51510) 
853 1.0 3.0 
891 4.5 16.7 
916 4.5 16'. 7; 
942 4.5 16.7 
989 4.5 16.7 
999 5. Lee 


351 


Figure 14. Part of the SBST.DAT file. 


199 
378 
421 
426 
466 
468 
489 
497 
680 
682 
722 
723 
730 
732 
740 
742 
746 
750 
815 
843 
894 
896 
914 
963 
966 
970 
977 
978 
979 
981 
989 
1007 


PoOorPOrPOrOrPOrPOrRPORrRPORPORPOrRPORORPOROROFO 


352 


V. RAW OUTPUT FILES. 


Seven files are generated as output from OTPOP. 
OTPOP.LOG is a log file, giving the date and time of 
execution, and the values of the parameters set at runtime 
(Fig. 15). RUNID.DAT will contain the run identification 
string. The other 5 files are output data files, containing 
the raw results of the simulations. Program PROC is used to 
process and summarize the raw output data. 


NTS.DAT contains the simulated population sizes for runs 
with an oil spill. There are 5 columns; the first gives the 
run number, the second gives the year (the oil spill always 
occurs in year 0, so that the years simulated before the 
spill are designated -3,-2,-1). Year O will be recorded 
twice for each run, once for just before the simulated spill, 
once for just after the spill. The third, fourth, and fifth 
columns give the number of independent females, number of 
independent males, and number of pups, respectively, for that 
year during the month of the spill (Fig. 16). 


NTC.DAT is exactly the same as NTS.DAT, but contains 
data for the control (without oil spill) runs, so year 0 wil 
be recorded only once for each run. 


DTS.DAT contains the total numbers of animals killed by 
the simulated spills. There are 13 columns; the first gives 
the run number, the next 6 pairs of columns give the number 
of animals and the proportion of the population killed for 
each of the 6 classes of animals (see discussion of the 
structure of OTMOVE in the documentation volume) respectively 
(Pilg 7) 


DDS.DAT contains the numbers of animals killed by the 
simulated spills on a daily basis. It has 8 columns; the 
first gives the run number, the second gives the day of the 
spill, and the third through eighth give the numbers of 
animals of classes 1-6 (see discussion of the structure of 
OTMOVE in the documentation volume) killed on that particular 
day (Fig. 18). 


RVS.DAT contains the population's reproductive value 
(see discussion of model output in documentation volume) 
before and after the simulated spill. It has 3 columns, the 
first giving the run number, the second giving the population 
reproductive value just before the spill, and the third 
giving the population reproductive value just after the spill 
(Fig. 19). 


RCS.DAT contains the recovery times for population. It 
has 3 columns. The first gives the run number, the second 
gives the population size just prior to the spill, and the 
third gives the number of years simulated before that 


353 


Figure 15. The OTPOP.LOG file. 


FOLLOWING ARE RUNTIME INPUTS FOR RUNIL: 
EXAMP1 


DATE -- 9/ 4/87 TIME -- 22:54 


DATE AND TIME USED TO GENERATE RANDOM NUMBER SEEDS 
10,"NUMBER OF YEARS PER RUN" 
10,"NUMBER OF RUNS WITH OIL SPILL" 
10,"NUMBER OF CONTROL RUNS" 
1000,"INITIAL POPULATION SIZE" 
1,"MONTH OF SPILL" 
1,"DAY OF SPILL" 
1,"DURATION OF SPILL 
345,"NORTH BOUNDARY OF SPILL" 
400,"SOUTH BOUNDARY OF SPILL" 
1000,"EQUILIBRIUM POPULATION SIZE" 
201,"NORTH BOUNDARY OF RANGE" 
955,"SOUTH BOUNDARY OF RANGE" 
0.090,"MAXIMUM PER CAPITA ANNUAL GROWTH RATE" 
0.005, "NON-LINEARITY OF DENSITY DEPENDENCE" 
0.000,"DENSITY INDEPENDENT MORTALITY RATE" 
1.000,"DEGREE OF COMPENSATION" 
0.930,"ADULT FEMALE SURVIVAL RATE" 
15.000,"MODAL AGE OF FEMALE SENESCENCE" 
0.250,"PRIME REPRODUCTIVE RATE" 
0.530,"PUP SURVIVAL RATE" 
0.870,"ADULT MALE SURVIVAL RATE" 
9.000,"MODAL MALE AGE OF SENESCENCE" 
5.000,"PERCENT VARIATION IN ADULT SURVIVAL" 
5.000,"PERCENT VARIATION IN PUP SURVIVAL" 
woe WARY "CE" "STGMA™, "VMAX" 
"JUVENILE FEMALES" 0.367-0.163 8.09037.500 
WADULT FEMALES W/PUP"=-0.025-0.406 6.39037.500 
"ADULT FEMALES W/O PUP"=-0.009-0.706 2.950 8.400 
"JUVENILE MALES"-0.045-0.290 8.56048.900 
“ADULT NON-TERRITORIAL MALES" 0.105-0.815 4.64048.900 
"ADULT TERRITORIAL MALES" 0.042=-1.044 1.93048.900 
6,"AGE AT WHICH POTENTIALLY TERRITORIAL" 
60.000,"MAXIMUM % OF POTENTIALS THAT HOLD TERRITORIES" 
20.000,"MINIMUM % OF POTENTIALS THAT HOLD TERRITORIES" 
8.110,"MEAN TERRITORY LENGTH" 
0.440,"S.D. OF TERRITORY LENGTH" 
5,"MEAN ARRIVAL DATE MONTH" 
23,"MEAN ARRIVAL DATE DAY" 
11,"S.D. OF ARRIVAL DATE IN DAYS" 
12,"MEAN DEPARTURE DATE MONTH" 
1,"MEAN DEPARTURE DATE DAY" 
15,"S.D. OF DEPARTURE DATE IN DAYS" 
"DAY", "P (MORTALITY) ","P(AVOIDANCE) ","P(EMIGRATION) " 
1 1.000 0.000 0.000 


354 


Figure 16. Part of the NTS.DAT file. 


1 1 593 361 130 
1 2 591 367 126 
1 3 590 368 127 
1 4 617 358 124 
1 5 641 376 138 
1 6 622 368 120 
1 7 599 345 122 
al 8 597 347 118 
1 9 578 335 121 
1 10 579 327 125 
2 -4 647 352 101 
2 =3 630 343 128 
2 -2 630 331 137 
2 -1 632 348 131 
2 0 644 337 117 
2 0) 597 324 108 
2 1 607 310 137 
2 2 623 344 130 
2 3 624 340 114 
2 4 621 322 137 
2 5 650 346 121 
2 6 673 331 142 
2 7 676 350 143 
2 8 666 366 132 
2 9 656 351 138 
2 10 664 355 140 
3 -4 647 352 105 
3 -3 659 355 130 
3 =2 666 366 129 
3 cal 658 371 121 
3 0) 662 350 118 
3 fo) 621 330 111 
3 1 635 332 119 
3 2 645 340 142 
3 3 623 337 132 
3 4 626 345 123 
3 5 624 322 132 
3 6 630 337 123 
3 7 645 321 121 
3 8 637 327 130 
3 9 654 325 114 
3 10 654 319 138 
4 -4 647 352 122 
4 -3 653 336 148 
4 -2 669 318 137 
4 al, 668 325 133 


355 


Figure 17. 


rR 


OUWUDNAHAUPWNER 


(oe 


h 
WWONNWE OO W 


0.020 
0.040 
0.043 
0.073 
0.019 
0.044 
0.012 
0.061 
0.017 
0.057 


Part of the DTS.DAT file. 


0.045 
0.037 
0.066 
0.064 
0.031 
0.044 
0.020 
0.050 
0.057 
0.052 


0.093 
0.071 
0.079 
0.070 
0.080 
0.083 
0.105 
0.101 
0.039 
0.068 


356 


0.083 
0.074 
0.055 
0.065 
0.083 
0.090 
0.070 
0.098 


0.071 


0.067 


0.122 
0.077 
0.059 
0.089 
0.065 
0.102 
0.060 
0.102 
0.083 
0.074 


0.073 
0.063 
0.060 
0.070 
0.062 
0.076 
0.057 
0.084 
0.057 
0.064 


Figure 18. Part of the DDS.DAT file. 


1 1 3 7 3 ab7/ 23 16 
2 1 6 5 2 12 26 9 
3 1 6 13 1 15 19 7 
4 1 11 10 1 13 21 11 
5 1 3 4 3 15 33 8 
6 1 7 8 1 15 31 12 
7 1 2 1 3 18 26 7 
8 1 10 9 2 16 33 11 
9 1 3 10 2 6 23 11 
10 1 9 9 2 13 22 10 


357 


Figure 19. Part of the RVS.DAT file. 


1 820.5 749.6 
2 819.9 761.1 
3 829.4 781.5 
4 861.8 807.8 
5 856.5 788.0 
6 829.5 753.2 
7 800.4 739.0 
8 762.9 691.9 
9 807.1 755.7 
10 859.6 802.3 


358 


population size was reached again (Fig. 20). For simulations 
in which the population never attained pre-spill size the 
time to recovery is recorded as a negative number. 


359 


Figure 20. Part of the RCS.DAT file. 


1 1158 -10 
2 1098 5 
3 1130 -10 
4 1121 -10 
5 1191 -10 
6 1139 8 
7 1131 -10 
8 1099 9 
9 1152 -10 
10 1176 -10 


360 


VI. PROCESSING RAW OUTPUT DATA 


Raw output data are summarized using the FORTRAN program 
PROC and LOTUS123. PROC is run automatically by the batch 
file OP.BAT after MAINPOP finishes execution. It produces a 
summary of the run (Fig. 21) in the file "OTPOP.RPT". At the 
top of OTPOP.RPT the RUNID identification string is given. 
Following is a summary of the control runs listing the year 
and the number of independent females, independent males, and 
dependent pups in the population. In parentheses after each 
of these are given the range and standard deviations of the 
population sizes in each year. 


Following the summary of the control runs is a summary 
of the oil spill runs, in the same format. Year "-0" is just 
prior to the spill, year "+0" is just after the spill. 


Following that is a summary of the total number of 
deaths from the oil spili, and a summary of the recovery 
after the spill. 


Besides OTPOP.RPT, PROC generates 3 files, 
"NO SPILL.DAT", "DEATHS.DAT", and "SPILL.DAT", that can be 
used by LOTUS123 to produce rough graphs of the model 
output. A LOTUS macro, "0OT123.WK1" is supplied. teas 
expected that you have a general understanding of the 
LOTUS123 package (including PrintGraph) in order to use the 
macro. 


Files "SPILL.PIC", "NO _SPILL.PIC", "COMPARE.PIC", and 
"DEATHS.PIC" are supplied with "0T123.WK1" and must be 
present in order for the macro to run. To run the macro, run 
Lotus 123 and then retrieve [/FR] the worksheet 0T123.WK1 in 
the usual manner (see your LOTUS123 manual for details). At 
this point you may invoke the macro by keying [alt]G. As the 
macro runs, 4 graphs will be displayed on the screen. The 
macro will pause while displaying each graph and you will be 
required to depress the space bar in order to proceed with 
the execution of the macro. 


After the macro has been executed successfully the 4 
above mentioned picture files (*.PIC) will be created. In 
addition to these newly created files, the graphs will be 
named in the current worksheet. If you wish to keep this 
information, save the current worksheet OT123.WK1 under a new 
name. From this point, you may wish to modify the graphs 
using the LOTUS123 /Graph menus and save modified graphs in 
new .PIC files. For information on how to create or modify 
graphs consult your LOTUS123 manual. 


The first graph (NO SPILL.PIC) produced by the macro 


traces the mean total independent population size through 
each year of the control simulations (Fig. 22). The lines on 


361 


OUTPUT FOR RUNID EARP! 


CONTROL RUNS: 


YEAR FEMALES 


HALES 


PUPS 


ee on ow ow on oo nn ono nn ws een nw coe wn come wn ce ewn ones ececese sc coen nese coeooe 


647.0( 647- 647, 
836.61 622- 447, 
640.44 624- 451, 
&44.81 622- 667, 
0 640.81 S97- 468, 
639.51 S99- 659, 
632.81 b01- d61, 
624.91 592- bbl, 
625.0( 602- 472, 
830.51 604- 447, 
637.3( 602- 676, 
837.41 602-63, 
640.91 608- 688, 
645.41 625- 478, 
648.81 631- 669, 


ee a 


OIL SPILL RUNS: 


0.0) 

6.3) 

8.3) 
13.4) 
21.8) 
19.0) 
20.9) 
19.5) 
16.1) 
16.4) 
21.2) 
18.8) 
22.8) 
15.3) 
13.9) 


352.0( 352- 352, 
349.5( 333- 341, 
352.1 331- 372, 
353.7( 345- 377, 
354.8( 317- 376, 
345.2 328- 385, 
363.41 327- 390, 20.2) 
358.7( 330- 383, 18.5) 
356.51 320- 393, 2101) 
348.94 32t- 387, 26.01 
341.3 315- 376, 18.0) 
345.20 31S- 371, 18.6) 
345.8 313- 376, 18.2) 
348.6¢ 304- 401, 25.5) 
352.71 31S 385, 19.0) 


0.0) 
8.2) 
14.0) 
9.5) 
18.4) 
16.8) 


$6.4) 
7.3) 
9.7) 
7.3) 
11.8) 
7.2) 
10.2) 
8.7) 
7.2) 
3.6) 
6.1) 
6.4) 
12.3) 
10.4) 
5.1) 


$20.1 
135.01 
131.01 
126.51 
127.3 
126.51 
126.44 
126.71 
130.9 
126.1 
127.16 
127.7 
130.3 
134.86 
135.41 


10b- 129, 
123- 152, 
115-149, 
117-139, 
102- 142, 
112-135, 
112-148, 
113- 141, 
119- 141, 
110-138, 
\N7= 138, 
116-139, 
IN1- 152, 
121- 153, 
123- 142, 


847.04 G47- 847, 
842.8 G21- 459, 
846.71 B10- 69, 
850.9 G1S- 668, 
852.21 608- 6&0, 
602.31 548- 629, 
605.7 SSI- 835, 
6°.5( ST7- 85l, 
o1b.8t Sal- 654, 
624.34 607-443, 
830.21 595- 450, 
&35.41 16-873, 
639.31 579- 476, 
842.5 S97- 873, 
b49.1( 578- 678, 
$50.9 579- 681, 


owm vo Ue WA 


18.2) 
17.8) 
21.4) 
24.9) 
28.0) 
26.2) 


352.0( 352 352, 0.0) 
348,1( 335- 364, 8.6) 
354.8( 316- 380, 18.8) 
340.8( 325- 382, 17.7) 
364.9 323- 387, %.5) 
349.30 Pre- 377, 22-91 
347.6( 310- 379, 22.4) 
348.5 324- 369, 14.9) 
345.41 329- 372, 15.5) 
338.6 322- 362, 13.91 
343.2 305- 387, 23.3) 
338.0( 299- 388, 19.21 
334.8 314- 352, 13.8) 
330.7( 298 366, 21.2) 
330.8( 305- 351, 14.3). 
333.5( 297- 355, 18.9) 


WUMBER OF DEATHS FROM DIL SPILL: 


10t- 128, 
125- 134, 
M7- 137. 
{7- *s 
108- 
q- 
104- 


1272.4 
112.26 
118.46 
121.0 
120.71 
126.81 
126.4 
123.7 
127.5 
127.4 
129.26 
134.71 


113-137, 
12i- 138, 5.2) 
120- 142, 6.3) 
S18- 143, 
118- 141, 
118-141, 
123- 147, 


PERCENT OF POPULATION 


CLASS WEAN S.D. RANGE REAR S.D.. RANGE 
JUVENILE HALES 6.0 63.1 2- it 3.86 1.99 1.20 7.30 
ADULT MALES 9.6 2.8 4- 14 4.65 1.37 2.00 6.60 
JUVENILE FEMALES 14.0 3.2 6- 18 7.89 1.80 3.70 10.50 
ADULT FEMALES 35.9 5.4 26 - 44 7.56 1.22 5.50 9.80 
PUPS 10.2 2.6 T- 16 8.33 1.96 5.90 12.20 
TOTAL ANIMALS 13.7 9.0 b4- 92 6.66 0.84 5.70 8.40 
RECOVERY AFTER OIL SPILL: 

MEAN $.D. RANGE 
REPRODUCTIVE VALUE BEFORE SPILL 824.8 29.0 762.9- 9861.8 
REPRODUCTIVE YALUE AFTER SPILL 763.0 32.4 671.9- 907.8 
REPRODUCTIVE VALUE REDUCTION (2) 7.506 1.19617 S.775- 9.307 
YEARS TO RECOVERY 7.333 1.89967 3.000- 9.000 


$8 ON 67 OF 10 RUNS ( 70.02) THE POPULATION 


DID MOT RECOVER TO PRE-SPILL SIZE 


TIME TO RECOVERY CALCULATED ONLY FOR RUNS THAT DID RECOVER. 
Figure 21. The OTPOP.RPT file generated by PROC. 


362 


Figure 22. Population sizes through each year of control 
simulations. 


NO SPILL @ YEAR ZERO 


EXAMP14 
1.03 
1.02 
4.01 
o 1 
nn 
\ 
+@ 
-3 0.99 
us 
oO 
23 c.98 
FE 
= yw 
rf 0.97 
z 
0.96 
0.95 
0.94 7 
ag = oO 2 4 6 8 10 
YEAR 


363 


either side of the mean trace the range of population sizes 
that occurred during the simulations. The second graph 
(SPILL.PIC) traces the oil spill simulations in the same 
format as NO SPILL.PIC (Fig. 23). The third graph 
(COMPARE.PIC) plots the means from both the control and the 
oil spill runs on the same graph, allowing visual comparison 
of the population trajectories (Fig. 24). The last graph 
(DEATH.PIC) traces the mean cumulative number of deaths due 
to oil spill on each day of the spill (Fig. 25). 


If you wish to modify the macro or just want to see the 
macro depress the [End] then the [Home] key and you will move 
to the far end of the worksheet, where the macro is located. 
If you wish to modify the macro we suggest that you first 
make a copy of it in case you need to refer to the original 
while editing. 


364 


Figure 23. 


MEAN g INGEP. +—S.D. 
(CThousande) 


Population sizes during oil spill simulations. 


SPIRE TG YEAR ZER© 


SAMP 1 


365 


MEAN # W/SPILL & W/O SPILL 
(Thousands) 


Figure 24. Plot of control population and oil 
population for simulations. 


COMPARE 


EXAMP1 
1,02 7 


1.01 


099 
0.98 > 
0.97 
0.96 


0.98 


366 


10 


CUMM. LOSS INOEP. & PUPS 


Figure 25. Cumulative deaths due to oil spill. 


DEATHS 


EXAMP 4 


DAY 


VII. ERRORS. 


Errors in model runs can occur at 3 levels. Screen 
generator errors occur when values out of range or of the 
wrong type are input. The program will warn you of the error 
and it can be corrected by simply typing in an acceptable 
value. A screen generator error will also occur if the SGX 
interface has not been loaded prior to running the model, or 
if the screen library PARASC is not available. 


IBM Professional FORTRAN runtime errors can occur when 
the program is unable to execute a program statement. For 
instance, if one of the input .DAT files is missing, 
incomplete, or in the wrong format. The error messages are 
generated from the PROFORT.ERR file, and will be noted on the 
screen. These messages may often be cryptic. If a PROFORT 
message indicates a unit or input error, check that all input 
files are present and in the correct format. Other 
situations that may lead to PROFORT runtime errors occur when 
impossible mathematical operations are‘ attempted, such as 
division by 0 or taking the log of a negative number. 
Usually ‘this will occur when an unrealistic combination of 
parameter values have been input. The error may or may not 
be fatal to program execution, but even if the program 
continues to run the output will be suspect. 


Program error trapping is the third level of possible 
error. Because there are so many possible combinations of 
parameter values that can be set, it has been impossible to 
build in complete logical error-trapping, but the program 
does check for many impossible combinations, such as setting 
survival rates too low to achieve specified growth rates, or 
setting pup survival rate too low to achieve the specified 
reproductive rate. Should an error occur on this level the 
program will terminate and you should correct the parameter 
values in the next run. 


The program is extremely complex, and it is very likely 
that errors not described herein may occur. No program of 
this size can ever be guaranteed to be bug-free. Should you 
not be able to correct an error, be sure to save the log file 
and all input files, and any output files generated during 
the run. Notify Allan Brody or Don Siniff and send copies of 
the saved files. 


368 


SMITHSONIAN INSTITUTION LIBRARIES 


wi 


10 2245) put il 


e7/93 canst 


DAN 
S> 

4 

Aas 

‘NOSE 


NVINOSHLINS S31Y¥V 


a 

e [sj 

ie sy 
fon 0c Y 


ON, 


1st 


NSTITUTION 


\e 
“ 
& NY 


ny, 


INSTITL 


NOILNLILSNI_ NVINOSHLIWS 


SMITHSONIAN 
NVINOSHLINS S3IYV 


3J1YVYGIT LIBRARIES SMITHSONIAN 


iS) 


NOILNLILSNI 


NSTITUTION NOILNLILSNI NVINOSHLINS 


SMITHSONIAN 


INSTITUTION NOILNLILSNI 


INSTITUTION 


LIBRARIES 


NVINOSHLINS S3IYVYUSITLIBRARIE 


SMITHSONIAN 


STITUTION NOILAL NVINOSHLIWS 


=  F} 
e ¢ 
x 

YY, 
K S 


eC; 
on ne. Y 


S3Iuvuaii_LiB 


INSTITUTION 


Saluvudi7 LIBRARIES 


INSTITUTION 


saruvuagi 


INSTITU 


SAIYVYSIT LIBRARIES SMITHSONIAN 


NVINOSHLIWS 


LIBRARIES SMITHSONIAN 


X 
nase 


RARIES 


NOILNLILSNI 


S SA1uv 


NVINOSHLIW 


INSTITUTION NOILNLILSNI 


SMITHSONIAN 


SMITHSONIAN 


< 207, 


() 
A, 
SINOSA 


NVINOSHLINS SJ1av: 


nis 


INSTITUTION 
WS 
AN 


s31uvugit 


INSTITU 


NOILNLILSNI 


Gy 


LIBRARIES 


N 


SMITHSONIA 
SMITHSONIAN 


NVINOSHLINS s 


Ss 


vine in 


LIBRARIES 
NOILNLILSNI 


N 


SMITHSONIAN _INSTITUTION NOILNLILSNI N 


o 


S 


SAIYVUSITLIBRARIE 


Ssalyvagit 


INSTITUTION 


NVINOSHLINS Sit 


uvugi7_LIBRARIES 


NVINOSHLIWS 
SMITHSONIAN 
NVINOSHLINS 


INSTITUTION NOILNL N 


=a 
< 
za 
ro) 
wo 
ae 
E 
a 
wo 2 Mi Zz 4 i 2 @ aq 
is) re) Ww , Lu ‘ 
z a E 4 = Oy, % = 4 
ss e yx = < Often < aa 
= = a S oc 5 = =i 
3 2 z 2 = 2 = 2 
— =) od } 
JIYVUGIT LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSNI NVINOSHLIWS Salevia LIBRARIES § 
& Zz - z c z = 
= S = S Str CE = S 
g = 2 = SON 2 = 
> eis > = > MG 2 > 3 
= n = n = eo ie m ca 
m z mn z i z a Z 
on = wn = n Ses wn =. 
NSTITUTIO NOILNLILSNI NVINOSHLINS S3IYVYalT_LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSNI_N 
<8 Z g 2 ee 2 gE z 
XK < _< VY .- = =< NURS => J 
Wo = is = = KS SG 5 =) 
Wee Sr fe) 35 DG KK = SD Wes = oe 
WE 2 : g 3 ER 3 ZUR 2 a 
\ \. = = z E XY Ze = SS - 
> = 5 = > = = 
‘. 7a (2) i Zz w : a ?)) 
31uvagia SMITHSONIAN _ INSTITUTION, NOILNLILSNI_NVINOSHLIWS’ S31¥VHEI1_ LIBRARIES. S 
= wy 2 ws 2 Ww a wl 
= =. ih fp 3 z = ez = z 
si i Yip = < = < = <} 
5 a “fora = A 4 a =| a 
= mS = st = 
re) = p 3° = ro) = fe = 
2 ; 2 =| z = z = 
CST UTONR NOLLNMLILSNI~NVINOSHLIWS _ S3 fuvugd iy tl BRARI ESe PH INSTITUTIONS NOILALILSNI_N 
~ & 
ie) S ° = ° = ° o 
= x S 2 E 2 = 2 
= > = > = > = = 
= ie = 2 a 12 E a 
n 2) n* m m 
z Chee ue o z o = pee 
ZLIBRARIES, SMITHSONIAN INSTITUTION NOILMLILSNI_ NVINOSHLINS, S31uvua17 LIBRARIES, 
=." = < = wb, = = ff, 5 = 
z a = g = z 
oO # Go (eo) a= 
2 NOs 5 MWe ig z 8 3 
= NS 9 = ENS Q = 2 E zZ 
Sh aN = SS NES = > = SS > 
(7) “aS > ose > w = w P=! 
NSTITUTION p NOMLALILSNI_NVINOSHLINS ” S31 uUVYsIT LIBRARI ES SMITHSONIAN_INSTITUTION on 
a 2 z - z Ww <Gul) 
JAN us STITUS > uw <p> Sa us y, a ONS uw VON n 


2 Sa = Ray 2 SQ = Gap 2 apy = 2 Rea) = yyy) 2 
m Quoc Z Cents m CR assy m Sonne 2 m None ” CoE 5 
pT IEVUAITULIBRARIES SMITHSONIAN — yy NOLLNLILSNI_ NVINOSHLINS, Sa1uvu a (ou BRARIES, 
= < Va = < = <= = = NS z 
4 z \\ = 2 S z Sj Wis Xs = 
<= DD Qe Sz ro) aw {o} Bea 0 RYN 5 
g ZN 3 2 g : B 2 Ne : 
2 i eli = 2 E = 2 \. : 
: Br ae EN Ral tena = fe en Se 
_INSTITUTION NOILNLILSNI NVINOSHLINS S3IYVYSIT_LIBRARIES SMITHSONIAN _ NOILALILSNI 
z “ S <8 2) = 7) S “” = 
B = a es a Uh ery tee w é 
AM %, = 
=] iB = Ss = Oe, wy = e : 
c c = Ee <= YY c < = 
= o = z = : “Vg = 
ro) = re) a fe) a TG fe) = C 
z a z Set a a =a =) > 
_SalaVu¥di7 LIBRARIES SMITHSONIAN_INSTITUTION NOILNLILSNI_NVINOSHLINS S3IYVYSIT LIBRARIES 
a Sie = 2 Eas : e 
= o = o = o \ = ive) - 
5 Ee 5 a = rs NN oS D t 
= > = > = > \WSE > : 
= 2 =. a iE 2 Xs E 2 : 
| INSTITUTION NOILMLILSNI NVINOSHLINS S31NVYSIT_LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSNI 
Zs 2) z a) 5 Z 2 et n g 
< = < = x = < = : 
za ar z 4 z a =z = ate 
8 en AB o a WX. F o ~NN & cf 
aT : Oo = ° 3e SS fo} ac SENSO) : 
Ee z S Z, = AS = E Wr 2 : 
= Se 3 z Flue ee = ane : 
> SSIMYVUSIT LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSNI_NVINOSHLINS S3IYVYGIT LIBRARIES 
a w =, g (ap) a n , art wn 
a z= i Gps. =a fa el fea =I ; 
< = < Gils A < 2a YS < 3] : 
o S 02 Gt S co = WL oc 5 P 
a 5 a. 3 es Soe S 3 
a z a z =) z a) a 
| INSTITUTION NOILNLILSNI NVINOSHLINS S31YVYGIT LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSNI 
a z Ur $6 Z & S = = : 
wo 2 wo Ne ° o 2 se) 2 C 
2 = a WS 2 5 a 5 : 
> = > MG 2 > rE ; > E : 
r 5 fe = = 7 = 
ae = a = ae re a ie 
E : 
e Z B be o = @ z | 
> SSJYVYGIT LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSN! SAlYVYEIT_ LIBRARIES 
on z ; 7) z= : 7p) z a) z an < 
= eS = < Ww = = = i 4 is 
4 z 4 z GX =) z Sl ag “ : 
aoe o = fo) Ve SM Is 9 5 YLY Ors Se 
ny n 7) rp) Na S\ wo n  » GY 2) ( 
2 : : ESQ : ag! | 
5 : = > = > = = a Sage 
Za 7) ae a . S o Z 7) 
1 INSTITUTION NOILQLILSNI SAIMVHYGIT LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSNI 
s a) = “n 5 on z a 
Oo us 7) a on iS OB = 
| = = Set See Asa =H 
= ee = ce = ed) a s 
e = E s E a Uity § e 
3 » z © 2 a “Gf 3 s 
z x z i zE a ¢ 72 - 
3_SSIYVYGIT LIBRARIES SMITHSONIAN_INSTITUTION  NOILOLILSNI_NVINOSHLINS_Saluvualt 
3 = ee 3 Se : 
— [ov] — ow — oO ay ee) 
= 0 i= a) 5 2 5 a 
E D> re i > ps > = > 
= Bd =. ae = x ie 2 
Zz ny B° nm Z o = o | 
\_ INSTITUTION NOILALILSNI_ NVINOSHLINS SaluvugI7_LIBRARIES SMITHSONIAN INSTITUTION NOILNLILSNI 
a “e w =z w p = 2) Zz x 2) ) 
< = < = < = < = | 
BO = 2 = Zz = 53. = | 
8 NN: - 5 g 3 s WA: 
= MR AS GO x= fe) <= = =ZE Q'S | 
E WW 2 S = = = EF WM’ z } 
Sa) = = > = = ihe tes : 
S$ ,94!8VNGIT_LIBRARIES SMITHSONIAN INSTITUTION SEE ees eee Tiew Tt ae 
" LGR 2 GER" Gin,7 aR» * e = O gore | 
ake a) 5 (AM) «= ly, | (Ma) = S E S | 
oe we] Clee se) < (fpr e Be) < eS ao \eK 45) S | 


1 * if oe 
a t x ‘Penee ib steyaste 


farbea emia Dy DAG Si ben hy 
DIMLaN Ha Woh 
SUCH ak hast ehube ata ¢ 
peepee 8, 
deh ‘an Udit 


Loree tees ' : ayiaale 
RLS AUA ENT Ai i PCa Cee ee ane 
ter cory Bee te wr 


hae iec ks 8 teh 
See bins readin Sen! ie 
PR ROCHE TES Thhe a Myre ie alata 


ebwlih 
Dyk aes DD pra 


Sieh tee Ue alstgain 


nurde onan 
x isk 3 Saat 

ns nets TEP 
achueteiardnate ceestaree sey? faye ssbuit a. (abl 
i Digan nn ashe} 


teiree 
vacate fader 


ie yee E 


Ta yrubed pub es bre, 
ednderd ent 


BETTS ce yte fe 


Ey 


fat one eet si 
Peake: 


Predhangeasay 


prec 


iG 
eee Tuy Ty ath M 
Saernriete’ 


nese art 
Wgearesyaus 
eae kl gy 
yearn? 
arery 
SPIN 
Rhruaes Fippeaeieeu ne 
SiNpA ae atyouee sts ee yan: 
be vatept ela tary ans Mb e ere ¥ 


Bebra 


Seaside 
concn wlye dao r 
payee is 


" pee ta wasias 
Hee eT (Choo EEE Sn 
Viviyre Ne 


syrat 


eh 


Ee 
ith