31
Annu. Rev. Physiol. 2006. 68:223–51 doi: 10.1146/annurev.physiol.68.040104.105739 Copyright c 2006 by Annual Reviews. All rights reserved First published online as a Review in Advance on October 19, 2005 THE COMPARATIVE PHYSIOLOGY OF F OOD DEPRIVATION: From Feast to Famine Tobias Wang, 1 Carrie C.Y. Hung, 2 and David J. Randall 2 1 Department of Zoophysiology, Aarhus University, 8000 Aarhus C, Denmark; email: [email protected] 2 Department of Biology and Chemistry, City University of Hong Kong, Kowloon Tong, Hong Kong PRC; email: [email protected], [email protected] Key Words feeding, fasting, starvation, metabolism, atrophy, digestion, gastrointestinal organs, specific dynamic action, phenotypic plasticity Abstract The ability of animals to survive food deprivation is clearly of consider- able survival value. Unsurprisingly, therefore, all animals exhibit adaptive biochemical and physiological responses to the lack of food. Many animals inhabit environments in which food availability fluctuates or encounters with appropriate food items are rare and unpredictable; these species offer interesting opportunities to study physiological adaptations to fasting and starvation. When deprived of food, animals employ vari- ous behavioral, physiological, and structural responses to reduce metabolism, which prolongs the period in which energy reserves can cover metabolism. Such behavioral responses can include a reduction in spontaneous activity and a lowering in body tem- perature, although in later stages of food deprivation in which starvation commences, activity may increase as food-searching is activated. In most animals, the gastrointesti- nal tract undergoes marked atrophy when digestive processes are curtailed; this struc- tural response and others seem particularly pronounced in species that normally feed at intermittent intervals. Such animals, however, must be able to restore digestive func- tions soon after feeding, and these transitions appear to occur at low metabolic costs. INTRODUCTION All animals supply the energy required for basal metabolism, physical activity, growth, and reproduction from their food. When food is not available, animals must use internal energy stores to fuel these activities. Starvation resistance re- flects an animal’s ability to store energy and control its allocation during extreme resource limitation. Many animals live in environments in which food abundance and quality vary drastically over time, and periods of starvation are common. Bacteria can lead a feast-or-famine existence (1, 2), many planktonic species live in a resource-limited world (3), and many fish overwinter with little or no food. Sommer (4) suggested organisms can be grouped according to their responses to variations in food supplies. “Velocity specialists” are those species with high, 0066-4278/06/0315-0223$20.00 223 Annu. Rev. Physiol. 2006.68:223-251. Downloaded from arjournals.annualreviews.org by UNIVERSITY OF ILLINOIS - CHICAGO on 09/09/06. For personal use only.

Adaptación Ciclo Ayuno Alimentación

  • Upload
    roger

  • View
    10

  • Download
    0

Embed Size (px)

DESCRIPTION

All animals supply the energy required for basal metabolism, physical activity,growth, and reproduction from their food. When food is not available, animals must use internal energy stores to fuel these activities.

Citation preview

Page 1: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

10.1146/annurev.physiol.68.040104.105739

Annu. Rev. Physiol. 2006. 68:223–51doi: 10.1146/annurev.physiol.68.040104.105739

Copyright c© 2006 by Annual Reviews. All rights reservedFirst published online as a Review in Advance on October 19, 2005

THE COMPARATIVE PHYSIOLOGY OF FOOD

DEPRIVATION: From Feast to Famine

Tobias Wang,1 Carrie C.Y. Hung,2 and David J. Randall21Department of Zoophysiology, Aarhus University, 8000 Aarhus C, Denmark;email: [email protected] of Biology and Chemistry, City University of Hong Kong, Kowloon Tong,Hong Kong PRC; email: [email protected], [email protected]

Key Words feeding, fasting, starvation, metabolism, atrophy, digestion,gastrointestinal organs, specific dynamic action, phenotypic plasticity

■ Abstract The ability of animals to survive food deprivation is clearly of consider-able survival value. Unsurprisingly, therefore, all animals exhibit adaptive biochemicaland physiological responses to the lack of food. Many animals inhabit environmentsin which food availability fluctuates or encounters with appropriate food items are rareand unpredictable; these species offer interesting opportunities to study physiologicaladaptations to fasting and starvation. When deprived of food, animals employ vari-ous behavioral, physiological, and structural responses to reduce metabolism, whichprolongs the period in which energy reserves can cover metabolism. Such behavioralresponses can include a reduction in spontaneous activity and a lowering in body tem-perature, although in later stages of food deprivation in which starvation commences,activity may increase as food-searching is activated. In most animals, the gastrointesti-nal tract undergoes marked atrophy when digestive processes are curtailed; this struc-tural response and others seem particularly pronounced in species that normally feedat intermittent intervals. Such animals, however, must be able to restore digestive func-tions soon after feeding, and these transitions appear to occur at low metabolic costs.

INTRODUCTION

All animals supply the energy required for basal metabolism, physical activity,growth, and reproduction from their food. When food is not available, animalsmust use internal energy stores to fuel these activities. Starvation resistance re-flects an animal’s ability to store energy and control its allocation during extremeresource limitation. Many animals live in environments in which food abundanceand quality vary drastically over time, and periods of starvation are common.Bacteria can lead a feast-or-famine existence (1, 2), many planktonic species livein a resource-limited world (3), and many fish overwinter with little or no food.Sommer (4) suggested organisms can be grouped according to their responsesto variations in food supplies. “Velocity specialists” are those species with high,

0066-4278/06/0315-0223$20.00 223

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 2: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

224 WANG � HUNG � RANDALL

maximum population-growth rates that respond to sudden increases in resourceabundance with rapid increases in population density; many bacteria fall into thiscategory. “Affinity specialists” can maintain growth at low-food-intake levels; thismay describe many tropical fish. Lastly, “storage specialists” such as hibernatingmammals create large internal stores that increase the chances of survival whenresource abundance is extremely low.

Many of the animals that inhabit environments with fluctuating food availabilityare adapted to consume very large meals when prey is available. Classic examplesare sit-and-wait predators, such as snakes; such species feed only once or a fewtimes a year. Very large meals followed by low rates of energy expenditure, typicalof crocodiles as well as a number of other reptiles and various fish, allow fordays or weeks between feeding. When the lack of food is due to seasonal changesin temperature or water availability, animals may enter into dormancy, duringwhich digestive processes are curtailed. With the exception of birds and somelarge mammals, which migrate to more desirable areas, dormancy is common invertebrates living in temperate or arctic environments that hibernate during coldperiods of the year. In tropical areas, many vertebrates, with the notable exception ofbirds, enter into a dormant estivating state during dry, and typically warm, periods.Fasting may also occur when animals engage in activities that compete with oreven preclude feeding or the search for food. Such activities include migration,moulting, or the care of eggs or young (e.g., 5–7). Examples of voluntary anorexiaexist in all major groups of vertebrates and may be very prolonged. Some speciesof penguins, for example, do not eat for several months when tending to eggs (8,9); Pacific salmon do not feed during their upstream migration, which can be morethan 1000 km in length (10); and eels do not feed during their migration across theAtlantic to spawn.

The ability of animals to survive food deprivation is clearly of considerablesurvival value. Unsurprisingly, therefore, all animals exhibit adaptive biochemicaland physiological responses to the lack of food. These responses prolong survivalwhen food is not available. Equally important, however, these responses also helpanimals to (a) preserve physiological functions so that behaviors, such as physicalactivity to avoid predators or to seek food, can be maintained and (b) ensure thatthe animals can resume digestive and metabolic processes when food becomesavailable again.

OVERALL FASTING TOLERANCE IS DETERMINED BYENERGY STORES RELATIVE TO USAGE

When animals experience food deprivation, they must derive the energetic costsfor basal metabolism, physical activity, growth, and reproduction from the inter-nal energy available at the onset of fasting. Large energy stores at the onset offasting, therefore, obviously aid in prolonging starvation tolerance; animals com-monly gain weight before dormancy. A reduction in metabolism also prolongsstarvation tolerance, and various biochemical and physiological responses to food

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 3: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 225

deprivation contribute to the efficient use of resources. Animals reduce metabolismin diverse ways. Many animals inhibit reproduction and reduce both activity andbody temperature. In fact, many animals breed only when food supplies are read-ily available, for example, in the springtime in temperate regions, and they do notbreed during the winter, when food supplies are limited. The response to starvationis integrated at all levels of organization and is directed toward the survival of thespecies.

Starvation has been more extensively studied in birds and mammals than ininvertebrates and other vertebrates. Birds and mammals are special among ver-tebrates because they normally must eat at regular intervals owing to their highmetabolic rates relative to their body stores. Most definitions and ideas of fastingcome from the human literature. The extent to which our knowledge of fasting andstarvation in birds and mammals can be transferred to other animals is not clear,as we discuss below.

The Phases of Fasting and Starvation in Mammals and Birds

The responses to absolute food deprivation in birds and mammals proceed instages, culminating in death. The initial period involves fasting, and the laterstages starvation. The demarcation between these two states is rarely appreciated,perhaps owing to lack of definition. In humans, fasting often refers to abstinencefrom food, whereas starvation is used for a state of extreme hunger resulting froma prolonged lack of essential nutrients. In other words, starving is a state in whichan animal, having depleted energy stores, normally would feed to continue normalphysiological processes. The metabolic transitions of food deprivation in birdsand mammals have been divided into three phases. Most investigators probablywould agree that the transition from fasting to starvation occurs by the end ofphase II or the start of phase III. Alternatively, and not in conflict with this view,others have argued that the transition between fasting and starvation may occurwhenever animals opt to abort voluntary anorexia (sensu Reference 5). In ourview, fasting should denote voluntary anorexia. Thus, salmon moving upstreamand eels crossing the Atlantic to spawn are fasting rather than starving. The stagesof starvation, if they exist in fish, amphibians, and reptiles, span a much moreextended time frame, and the distinction between fasting and starvation becomessomewhat esoteric. In any event, the severity of the food deprivation and timerequired before an animal enters actual starvation vary among species, and thetransition to starvation will depend on individual responses and nutritional status aswell as a number of environmental conditions. Thus, whereas a small endothermicanimal may be starving within a day of lacking food, it would take much longer forlarge ectothermic animals to undergo the same transition. Future studies shouldaddress questions such as whether large predatory ectotherms, such as sharks, thathave not fed for a few days, are starving or fasting while waiting for the next meal.Similarly, is a hibernating ground squirrel starving or fasting?

The different phases of response to food deprivation, which are defined accord-ing to the progressive metabolic changes that occur, were initially characterized

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 4: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

226 WANG � HUNG � RANDALL

for humans and other mammals (e.g., 11–14). These stages have been successfullyapplied to birds (8, 9, 15), and although not thoroughly investigated in ectother-mic vertebrates, the overall progression in metabolic adaptation appears similar formost vertebrates. As a major difference, however, resting and maximal metabolismare much lower in ectothermic vertebrates than in endotherms (e.g., 16). For a givenbody composition and amount of energy stores, ectothermic vertebrates thereforecan maintain normal metabolic functions for much longer than can endother-mic animals, thereby deferring the detrimental consequences of food deprivation.Thus many ectothermic vertebrates can tolerate more lengthy starvation than canendotherms. Eels, for example, can migrate many thousands of kilometers overalmost a year without feeding and may survive lack of food for many years (17,18), whereas a similar-sized mammal would die from starvation within a few daysor a week. Likewise, small animals are much more susceptible to food deprivationthan are larger animals.

In mammals, the three metabolic phases during food deprivation are character-ized as follows on the bases of the primary fuel available for use and the associatedchanges in overall body mass:

Phase I. The postabsorptive phase is the initial phase of fasting immediatelyafter the last meal has been absorbed from the gastrointestinal tract. During thisperiod, which normally lasts for hours, metabolism is largely fueled by glycogenol-ysis, or glycogen depletion of liver stores, which maintain constant blood sugarlevels. In addition, fatty acids are liberated from adipose depots, and the availabil-ity of plasma fatty acids allows for some tissues, such as skeletal muscle, to sparethe overall use of glucose.

Phase II. When liver glycogen stores are depleted, gluconeogenesis becomesnecessary to supply the requirements of glucose-requiring organs such as the brain.In humans, the initial fuel for gluconeogenesis is amino acids from proteolysisof muscle protein, but this contribution falls markedly as increased amounts ofglycerol, another substrate for gluconeogenesis, is liberated from adipose tissues.Increased oxidation of fatty acids leads to an elevated production of ketone bodies,which can be used as an oxidative fuel in many tissues including the brain. As phaseII progresses, protein degradation is rather slow, and degradation of adipose tissuefuels most bodily metabolism. In owls, lipid contributes more than 90% of the en-ergy consumption in phase II, and approximately 2.5% of the energy consumptionis derived from protein (19). In humans, this state can be maintained for severalweeks and has often been referred to as a period of adapted starvation. Because ofthe high energy content of lipid, weight loss is rather slow during this state.

Phase III. If starvation continues until the adipose stores are depleted, muscleis rapidly degraded for gluconeogenesis. The rapid loss in muscle mass cannot besustained for long and eventually kills the animal.

Figure 1 presents these three phases in rats. It shows how, as rats enter phaseIII, the rate of body mass loss along with nitrogenous waste production and itsexcretion increase as a result of protein degradation.

Although similar metabolic changes and the transitions between fasting andstarvation remain to be studied in detail for ectothermic vertebrates, numerous

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 5: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 227

Figure 1 Changes in body mass, daily loss of body mass, and excretion of nitrogenous

waste in rats during the three phases of starvation (modified from Reference 48).

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 6: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

228 WANG � HUNG � RANDALL

studies have reported on the gradual but slow decrease in body weight and somaticindices as food is withheld. Most species studied utilize fat before protein is de-graded (e.g., 10, 20–24). Other reports suggest that in ectotherms, as in mammals,glycogen is utilized even before lipid or protein (25, 26). Table 1 shows expressionprofiles of the main energy-generating pathways that are related to the three phasesof the mammalian response to starvation. Mammals have to utilize energy reservesmuch earlier than fish in response to starvation. Genes that encode protein productsin lipolysis and protein turnover were induced after 24 and 48 h of starvation inmice and rats. Figure 2 shows that starvation did not trigger significant changes ingene expression in carp until after at least 16 days of food deprivation. Lipolysisgenes, such as β-oxidation, remained unchanged in carp liver throughout the sixweeks of food deprivation, reflecting the fact that, unlike in mice and rats, hep-atic lipid utilization was not enhanced in carp. Carp hepatic ubiquitin-proteasomegenes were upregulated by approximately 1.3-fold after 28 days of starvation butdid not trigger a significant decline in total hepatic protein content. In fact, hepaticprotein appears to be well conserved during starvation in carp (27–30). Of course,before mobilizing hepatic reserves, carp use other lipid sources such as viscerallipid, as do rainbow trout (31). Carp contained a large amount of visceral lipids,although they were not quantified in the experiment. Carp hepatic glycogen, onthe other hand, was mobilized during the first four days of starvation and then de-clined again after six weeks, which coincided with an increase in glycolytic geneexpression. Hepatic glycogen was not exhausted completely in carp after 100 daysof starvation (32). Early mobilization of hepatic glycogen in carp may be relatedto glucagon release during the initial phase of starvation; this has been observed inteleosts and may be a response to stress rather than starvation (24, 33). Migratingsalmon utilize lipid and spare protein until later in the migration phase, when lipidstores are almost completely depleted (10).

REDUCTIONS IN ENERGY EXPENDITURE DURINGFOOD DEPRIVATION

Responses to starvation occur at the behavioral, physiological, biochemical, andmolecular levels. In general, the time to reach starvation-induced death increaseswith body mass (34–37), reflecting larger animals’ greater abilities to lower specificmetabolic rate and increase stores of energy. Reductions in energy expenditurecan also occur via a reduction in body temperature, which reduces metabolicrate (the Q10 or Arrhenius effect). Animals also can decrease energy expenditureduring starvation by reducing locomotor activity as well as other behavioral andphysiological functions such as reproduction and care for young. Many of theseactivities, although interconnected, are often studied in isolation. Protein synthesisis decreased, and expression of many metabolic genes is also reduced. Whethergene expression is reduced in response to decreased energy expenditure or viceversa is not clear.

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 7: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 229T

AB

LE

1G

ene

exp

ress

ion

of

mam

mal

san

dca

rpli

ver

inre

spo

nse

tost

arva

tio

n.U

pre

gu

lati

on

of

gen

esin

volv

edin

lip

oly

sis

and

pro

teo

lysi

s

iso

bse

rved

on

lyin

star

ved

mam

mal

saf

ter

1to

2d

ays

of

star

vati

on

,bu

tit

isn

ot

ob

serv

edin

com

mo

nca

rpev

enaf

ter

wee

ks

of

foo

dd

epri

vati

on

.

Fo

rd

etai

lso

fg

ene

list

s,p

leas

ere

fer

too

rig

inal

pu

bli

cati

on

s.

Ani

mal

mod

elM

etho

dsD

urat

ion

ofst

arva

tion

Upr

egul

ated

path

way

sD

ownr

egul

ated

path

way

sP

roba

ble

phas

eR

efer

ence

Mal

e1

29

/sv

cDN

Am

icro

arra

y2

4h

Lip

oly

sis:

β-o

xid

atio

nL

ipo

gen

esis

II1

48

mic

e(8

–1

5

wee

ks

old

)

Ure

acy

cle

gen

esC

ho

lest

ero

lsy

nth

esis

and

DH

EA

met

abo

lism

Am

ino

met

abo

lism

S-ad

eno

sylm

eth

ion

ine

(SA

M)

cycl

e

48

hS

tro

ng

erin

du

ctio

no

f

gen

esin

volv

edin

the

above

pat

hw

ays

Str

on

ger

sup

pre

ssio

no

f

gen

esin

volv

edin

the

above

pat

hw

ays

II,

III

tran

siti

on

Mal

eS

par

gu

e-

Daw

ley

rats

(14

5–

15

5g

)

Su

pp

ress

ion

sub

trac

tio

n

hybri

diz

atio

n

(SS

H)∗

48

hL

ipo

lysi

s:β

-ox

idat

ion

Glu

con

eog

enes

is

Pro

tein

turn

over

II1

49

Cyp

rinu

sca

rpio

cDN

Am

icro

arra

y4

day

sV

itel

log

enin

26

(20

0–

30

0g

)8

day

s

16

day

s1

6to

42

day

s:

28

day

sG

lyco

lysi

s(2

gen

es)

Lip

idb

iosy

nth

esis

?

Tri

carb

ox

yli

c(T

CA

)

cycl

e

Glu

con

eog

enes

is

Ub

iqu

itin

-pro

teas

om

e

pat

hw

ay(1

.3-f

old

)

Gly

coly

sis

42

day

sS

ust

ain

edex

pre

ssio

no

f

the

above

gen

es

Var

ied

exp

ress

ion

so

f

glu

con

eog

enic

gen

es

?

∗ SS

Hte

chniq

ue

iden

tifi

esonly

induce

dbu

tnot

suppre

ssed

gen

es.

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 8: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

230 WANG � HUNG � RANDALL

Fig

ure

2E

xp

ress

ion

pro

file

so

fst

atis

tica

lly

sig

nifi

can

tg

enes

of

carp

liver

ind

icat

eth

atg

ene

exp

ress

ion

do

esn

ot

chan

ge

un

til

afte

r1

6d

ays

of

star

vati

on

.G

enes

that

hav

esi

mil

arex

pre

ssio

np

rofi

les

are

clu

ster

edto

get

her

.

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 9: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 231

Basal and Resting Metabolic Rates

Humans and other mammals decrease resting metabolism during fasting and star-vation (38). A very pronounced example is that of the golden spiny mouse, Acomysrussatus, which inhabits dry deserts in the Middle East. Within one day, this smallrodent apparently reduces oxygen uptake to half of the normal value and main-tains this low metabolism when kept on restricted food availability for two weeks(39). This reduction takes places without changes in body temperature and mayrepresent sympathetic control of energy-requiring processes (39). However, a re-cent study on the same species showed that food restriction elicits a reduction inboth body temperature and metabolism that resembles torpor in other mammalianspecies (40). Nevertheless, although most animals do seem to lower body tempera-ture when food is limited, substantial reductions in basal metabolic rate may occur.In salmon, for example, oxygen uptake decreased gradually over approximatelytwo months when food was withheld (22).

A reduction in basal metabolic rate, with no attendant decline in body tempera-ture, requires that some energy-requiring processes be reduced at the cellular level.The changes in cellular metabolism responsible for a reduction in basal metabolismin fasting or food-restricted animals have not been studied, but the response mayinvolve some of the same mechanisms as those occurring during the metabolicreduction observed during hypoxia. Many vertebrates respond to lack of oxygenby lowering protein synthesis, and a lower membrane permeability decreases thedemand for active ion transport. Cell cycle may arrest, and cell proliferation maydecrease, leading to the observed reduction in growth (41). Certainly food restric-tion and therefore the lower rate of intestinal nutrient uptake should also decreaseprotein synthesis, cell proliferation, and growth. It is difficult to envision, however,that changes in membrane properties would occur without an associated loss offunction.

Measuring basal metabolic rate while controlling for changes in spontaneousactivity or alertness and sleep is difficult. Also, as described in more detail below,digestion also involves specific dynamic action and a rise in metabolism of animalsseemingly at basal conditions. Fasting or starvation may therefore be associatedwith an apparent decline in basal metabolism that actually should be ascribed to agradual cessation of digestive processes. The mammalian digestive tract displaysgreat morphological and functional changes in response to starvation. Epithelialcell renewal and cell migration from crypt to villi tips were both reduced in starvedmice (42) and rats (43). Total intestinal and jejunal mucosal mass was decreasedby a factor of two in fasting rats, accompanied by reductions in jejunal crypt sizeand villi size and numbers as well as increases in villi tip cell apoptosis (44). An-other recent study on rats, however, reports only minor changes in intestinal villusapoptosis (45). In any event, these studies concur that during phase III, apoptosisdecreases and cell proliferation and migration increase (44, 45). Altogether, thisupregulation of cellular events and presumably of absorptive capacity may reflectpreparation for future feeding (45). The restoration of the intestine during phase III

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 10: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

232 WANG � HUNG � RANDALL

of starvation may be related to a “refeeding signal,” which has been described inpenguins (46), and/or to behavioral changes in food hunting that mammals displayduring this phase (47). Intestinal restoration may also be related to increased expres-sion of genes encoding orexigenic, hypothalamic peptides such as neuropeptide Y,agouti-related protein, and pro-opiomelanocortin (48). Rapid restoration of intesti-nal structures was observed during refeeding in mammals (49). This restorationtook place as quickly as 30 minutes after refeeding following phase II starvation(43). By contrast, starvation had no significant effect on the intestinal tract of thecommon carp. After 42 days of starvation, intestinal mucosal thickness of carpwas not affected, and cellular events (e.g., apoptosis and cell proliferation) of thegut remained active. The expression of many digestive genes—including thosefor chemotrypsin A and B, elastases, trypsins, carboxypeptidase A and B, propro-teinase E, and amylase 3—were suppressed greatly during starvation in carp, withmost of the downregulation occurring after 16 days of starvation.

The acquisition and processing of food are expensive processes; their cessationis manifested as a reduction in basal metabolic rate. The large reduction in gutmucosal surface area during starvation probably results from a large reduction inenergy expenditure in maintaining the gut. This reduction in gut energy expenditureconstitutes a part of the overall bodily reduction in energy expenditure duringstarvation. Several studies reporting on metabolic depression of fasting animalsdid not report body temperature, and some of the decline in metabolism may stemfrom hypothermia. Thus, part of the alleged reduction in standard metabolic rateduring food deprivation may be ascribed to factors that lead to a reduction incellular metabolism.

Body Temperature

In both birds and mammals, fasting and the associated depletion of energy reservesare important physiological cues to initiate torpor, which is a reduction in bodytemperature during inactive parts of the diurnal cycle (50). Thus, in some species,torpor occurs only when energy stores have reached a certain threshold and can beprevented by artificial administration of nutrients such as glucose (e.g., 36, 40, 51–53). The gradual depletion of energy stores may also explain why the hypothermicresponse is enhanced as fasting is prolonged (e.g., 54). Torpor reduces energy us-age by the direct effect of temperature on metabolism and because the metabolismof activity is negligible. Hypothermia is more pronounced in small as comparedto large mammals and birds (e.g., 55–57). Some hummingbirds, for example, maydecrease body temperature by as much as 30◦C. Small animals may benefit fromthis body-size effect because of their higher mass-specific metabolism and greaterease for heat transfer due to their large surface area relative to body mass. However,body mass alone does not explain the occurrence and patterns of torpor. Many smallbirds, such as passerines, rarely reduce body temperature by more than 3–5◦C,whereas some larger birds can undergo much larger changes (e.g., 57–59). Tor-por appears to be more pronounced in animals that inhabit areas in which large

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 11: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 233

fluctuations in temperature and food availability are common, such as deserts.However, torpor also occurs in laboratory rats (e.g., 60) and may therefore be arather common and widespread response.

An entrance into torpor in response to food deprivation has been described invarious animals. Nocturnal reductions in peripheral temperatures, associated withlower heart rate and presumably reduced metabolism, occur in large mammalssuch as red deer and reindeer (e.g., 61, 62). Torpor also occurs in primates; forexample, torpor in response to food deprivation has recently been documented inthe gray mouse lemur, Microcebus murinu (63). Also, barnacle geese and PuertoRican todies undergoing long-term migration without feeding reduce body temper-ature by several degrees (64, 65). However, as Schleucher (57) points out, neitherfood supply nor energetic stress per se appear to be the ultimate factor determiningthe hypothermic response in these species, as the response is more pronounced infatter premigratory birds. Torpor, therefore, may be a strategy to reduce energyexpenditure during accumulation of fat stores. Thus, in addition to fasting and en-ergy status of the individual, diverse ecological, morphological, and physiologicalvariables, breeding, or migration periods, as well as physical parameters such asweather and annual cycles, are likely to influence the extent to which differentendotherms utilize torpor (e.g., 57).

Ectothermic animals rely on external heat sources and appropriate behaviorto regulate body temperature, and when provided with these opportunities, theymaintain remarkably constant and well-regulated body temperatures. The effectsof food deprivation have been studied in a few species freely selecting body temper-ature in laboratory settings. Several of these studies on fish and lizards have shownreductions in the preferred body temperature by a few degrees, which developsprogressively as food is withheld (e.g., 66, 67).

Physical Activity

Decreasing physical activity and allowing body temperature to decline are likely tocontribute more to energy sparing, and thereby to tolerance of starvation, than doreductions in basal metabolic rate, which are comparatively small. When food is notavailable, however, animals may search more actively for food, at the expense ofincreased energy usage, or decrease activity so as to reduce energy expenditure. Ananimal’s use of these alternatives depends on its foraging mode, the causes of fooddeprivation, and many other aspects of the animal’s natural history. In general, sit-and-wait predators are likely to reduce activity when food is not available, whereasactive hunters and grazers are more likely to increase activity as they search forfood. Furthermore, although many animals reduce physical activity during theinitial phases of fasting, many other animals exhibit a marked stimulation of activityduring the later and more critical phases of starvation.

In captive rats, food deprivation leads to reduced physical activity during theinitial phases of food deprivation (68), followed by a marked hyperactivity whenthe animals enter phase III of starvation (47; see also Reference 69). Similar events

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 12: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

234 WANG � HUNG � RANDALL

ensue in captive emperor penguins, in which the transition from phase II to III andthe associated depletion of fat stores coincide with increased activity and escapebehaviors. Teleonomically, these responses appear beneficial, as the transition tophase III of starvation signifies that existing resources are limited and that needfor food is acute.

Some [but not all (23, 66)] fasting fish and amphibians reduce activity. Mendez& Wieser (21) proposed that the behavioral response of fish to starvation consistsof three phases, which has some resemblance to the biochemical changes outlinedabove. The first phase is short lasting (approximately 24 h) and involves the in-creased activity of food searching. A transition phase, in which the fish graduallyreduce swimming activity and thereby lower energy expenditure, then follows. Thethird and final phase, adaptation, is characterized by low activity and metabolism,which persist until the fish are presented with the possibility of food. Van Dijket al. (66) did not observe the stress phase in fasting roach (Rutilus rutilus), andit is quite likely that the specific responses will vary among fish with differentbehaviors and with the experimental setting.

Reproduction

Reproduction is energy expensive and requires either large energy stores in themother or a ready source of food for both parents and offspring. Reproduction inmany animals coincides with a high probability of food. Starvation is an inhibitorof reproduction in vertebrates; for example, most anorexic human females do notmenstruate and cannot conceive (38). Female hamsters generally fail to ovulate andshow little interest in sex if deprived of food for one or two estrous cycles. Ovulatoryfailure in these animals is related to an absence of an ovulatory gonadotropinsurge and a set of immature follicles (70). Starvation for three days suppressessexual receptivity in female rats, and this is associated with a reduction in theestrogenic response in the ventromedial nuclei of the hyopothalamus, critical forsome reproductive behaviors (71). Sexually mature zebrafish spawn daily, but whenthey are starved, the number of eggs they produce per day drops rapidly (Figure 3)(26). The prompt decline in egg production is associated with decreased expressionof CYP19a, an enzyme that converts testosterone into estrogen in female zebrafish(Figure 4).

Not all animals exhibit inhibition of reproduction during starvation. Tessieret al. (72) suggested that reproduction during starvation may be advantageous forshort-lived species. Some rotifer species maintain or even increase egg productionduring starvation (73), whereas other rotifer species reduce reproduction and sur-vive starvation for longer than the reproducing rotifer species. Those rotifers thatincrease reproduction when energy supplies are limited invest resources in theiroffspring, which presumably have a better chance of surviving starvation. Thisis at the expense of the parent’s survival, possibly because of the accompanyingbenefits of predator avoidance, reduced energy requirements of the young, and/orincreased chance of moving to a resource-rich environment.

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 13: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 235

Figure 3 Daily egg production of zebrafish when fed continuously for 10 days and after

11 days of food deprivation (fasting). A parallel experiment on continuously fed fish, which

served as a control group, is shown on the right side of the figure. Each group represents results

from 26 pairs of zebrafish, sex ratio 1:1, and presented as mean ± S.E. Food-deprived fish

produced significantly lower numbers of eggs produced than did fed fish (modified from

Reference 26).

Some vertebrates starve during reproduction; the Pacific salmon and eel areclassic examples. Yellow eels feed and grow in freshwater, but they stop feedingwhen they become silver and start their migration across the Atlantic to spawn inthe Sargasso Sea. Eels that have been starved in both seawater (74) and freshwater(75) for up to three to four years have survived. During this time, these eels lost

Figure 4 RT-PCR analysis of CYP19a mRNA expression in female gonad of control and

zebrafish starved for 11 days.

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 14: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

236 WANG � HUNG � RANDALL

between 70 and 80% of their body weight. Interestingly, starvation was associatedwith an increase in growth hormone (GH) brain-cell hypertrophy and plasma GHlevels (75). GH binds predominately to liver cell membrane receptors, but theincrease in plasma GH during starvation is not associated with an increase in livergrowth, as these liver GH receptors are downregulated.

Atlantic salmon and trout species spawn repeatedly, whereas Pacific salmon dieafter their first spawning. The lifestyles of Pacific salmon vary between species.Sockeye salmon, Oncorhynchus nerka, are born in freshwater and enter the oceanas one- or two-year olds, weighing between 4 and 15 g. They return to their natalstream to spawn after two or three years at sea, weighing between 1.6 and 3 kg (76).These animals do not feed once they have entered the river. The upstream migrationmay be more than 1000 km, depending on the river. Not only do these animalsswim such large distances but they also produce numerous eggs. The gonad of thefemale reaches 14% of the pre-spawning total body mass. The starving migratingPacific salmon use fat to fuel both their upstream migration and egg production;at death, both sexes will have expended more than 95% of their fat reserves (10).The fat reserves when the fish enters the river are, in general, proportional tothe distance to be traveled. Female pink salmon spend less energy on migrationthan do males but more on gonad production such that, in the end, total energyexpenditure is approximately the same for both sexes. When fat reserves start toexhaust, first protein (from white muscle, but not heart or red muscle) and thencarbohydrates are utilized (77). When the fish reaches the spawning ground, thecalorific content of the fish is reduced to less than half of that which existed whenthe fish entered the river (10). There is increased interrenal activity and cortisolproduction, presumably directing some of the metabolic changes. The fish spawnand are usually in a moribund condition associated with energy depletion. All ofthe fish die, but they are not all moribund, and the cause of death is not clear,although the depletion of energy reserves must be an important component. Othersalmonids, such as trout and Atlantic salmon, are repeat spawners.

Immediately after reproduction, survival of the parent rather than the offspringis usually favored; this is particularly true for vertebrates. In penguins that faston the ice during incubation of their eggs, the transition to phase III also leads toincreased activity, and parents will abandon their eggs to secure their own survivalat the expense of successful reproduction (46). There are, however, exceptions.Some species of octopus continually ventilate their developing embryos; feedingbehavior is inhibited during this time, and these octopi can starve to death in theprocess. Adult cuttlefish, Sepia officinalis, migrate toward coastal waters to spawnand then die a month later. They also stop feeding and age rapidly during thisperiod, with a marked deterioration in long-term memory. The cessation of feedingis associated with defects in visuomotor coordination as a result of degenerativechanges in the central nervous system (78). There may be selection for genesthat cause the rapid death of the postreproductive, or even just older and larger-sized, individuals within the population. This may be the case in spawning salmonand lampreys. Nothing is known of such death genes, but if they exist, they may

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 15: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 237

be functionally similar to genes observed in Sepia (78), in which degenerativechanges in the central nervous system lead to a loss in prey capture ability and, asa consequence, death by starvation.

PHENOTYPIC PLASTICITY OF THE VISCERAL ORGANSIN RESPONSE TO DIGESTIVE STATE

The size and functional capacity of most visceral organs and muscle change inresponse to the physiological demands that are placed upon them (e.g., 79, 80). Thisphenotypic plasticity is pronounced for the gastrointestinal organs, which undergoa marked structural and functional reduction during fasting. The gastrointestinalorgans are very metabolically active and have been estimated to account for asmuch as 40% of basal metabolic rate (e.g., 81). Thus, a reduction in organ sizeduring fasting may confer a significant energetic savings, which may contribute toa marked reduction in the basal metabolic rate of fasting animals.

Within nonmammalian vertebrates, the effects of food deprivation on gastroin-testinal organs have predominantly been studied in snakes. This group of rep-tiles has attracted particular interest because they tolerate very prolonged fastingperiods—in some cases up to several years—and because they can ingest verylarge meals. Thus, under natural conditions, some snakes may eat only a few timesa year, but when they do eat, they can consume prey items of 50% of their ownbody mass or more (e.g., 82, 83). When snakes such as pythons or rattlesnakeseat these large meals after fasting for a few weeks or longer, the mass of the smallintestine increases drastically within the first 12–24 h after ingestion (84–86). Thegut wall of reptiles, like that of other vertebrates, consists of an outer muscularcoat and an inner mucosal layer with an epithelial lining toward the gut lumen(80, 87, 88). The mucosa in particular increases in mass upon feeding (Figure 5)(85, 89), and is attended by a many-fold increase in the transport capacities forvarious amino acids and glucose (85). This rise in nutrient transport capacity likelyreflects that the length of the intestinal microvilli increases almost fivefold within24 h (Figure 6) (90).

In ectothermic vertebrates, whether the nutrient transport proteins are beingsynthesized de novo as the enterocytes expand and microvilli lengthen, or whetherthe increased nutrient transport capacity merely reflects the increased surface areaand that more transport proteins are exposed to the lumen, is unknown. Whenexpressed relative to the mass of the intestine, nutrient transport capacity actuallyincreases in mammals during hibernation (91–93). Thus, although the mucosaatrophies, the intestinal transport proteins are well preserved, and the mRNA levelsof the transporter protein SGLT1 does not change during hibernation in groundsquirrels (94). This is also the case for the activity and mRNA levels of Na+,K+ ATPase (94). The membrane potential of the enterocytes actually increasesslightly during hibernation, which enhances the Na+ gradient that drives many ofthe intestinal nutrient transporters (92). It is not clear how this hyperpolarization

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 16: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

238 WANG � HUNG � RANDALL

Figure 5 Effects of feeding on the mass of the small intestine of the snake Python molurusbefore and after ingestion of a meal equalling 25% of the snake’s body mass. Intestinal mass

is shown for fasting snakes (time 0). Each bar represents the total mass of the intestine and is

divided into three parts representing the three parts of the intestine: proximal (black), middle

(gray), and distal (white). The data are modified from Reference 90.

of the enterocytes affects cellular metabolism, and it certainly would be of interestto perform similar studies in ectothermic vertebrates in which cellular functionscould be compared at similar temperatures in fasting and digesting animals.

In contrast to that of the mucosa, the thickness of the intestinal muscle layerappears unchanged (85, 89). Furthermore, digestive status does not affect the totalnumber of gut neurons in the intestinal muscle layer, spontaneous activity of themuscle layer in vitro, or the motility responses of isolated intestinal preparationswhen exposed to various excitatory neuropeptides (87). The mass and nutrienttransport capacity of the gastrointestinal system undergo progressive reductionsduring subsequent food deprivation. Although other species of nonmammalian ver-tebrates have received much less attention than have snakes, phenotypic plasticity,both in terms of mass of the organs and their functional correlates, is seeminglyuniversal, albeit less pronounced, in species with a more continuous feeding pat-tern, where prolonged periods of food deprivation are less common (e.g., 95).Thus, a progressive reduction in the intestinal epithelium during fasting, whichis rapidly reversed after feeding, occurs in all other major groups of ectothermicvertebrates (e.g., 95–98, 150, 151).

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 17: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 239

Figure 6 Electron micrographs of proximal intestinal microvilli of Python molurus during

fasting (a) and at 0.25 (b), 0.5 (c), 1 (d ), 3 (e), 6 ( f ), and 14 (g) days after ingestion of a meal

equalling 25% of the snake’s body mass. Note the immediate lengthening of the microvilli

and the subsequent regression (bar = 1 μm). Modified from Reference 90.

Organ growth can occur by increased cell size (hypertrophy) or cell proliferation(hyperplasia). A two- or threefold increase in mucosal mass through hyperplasiawould require extremely high rates of cell division and mitotic activities and pre-sumably would be energetically expensive. Several recent studies on snakes andother ectothermic vertebrates have shown that the feeding-induced rise in intestinalmass is due predominantly to increased size of the individual enterocytes (84, 89,90, 99–101), suggesting that hypertrophy is the major mechanism. Thus, althoughcell proliferation may start early in the digestive phase, cell division reaches itsmaximal rate rather late in the digestive process. Therefore, the cells that have been“worn down” during digestion may be replaced. In this manner, the fully func-tional gut can be rapidly restored when food becomes available again (89). In allectothermic vertebrates, the epithelium—which in fasting animals is pseudostrat-ified, with folded cell membranes of neighboring cells—may be rapidly unfoldedafter feeding and converted to a single layer of cells with stretched membranesas the enterocytes expand (reviewed in Reference 80). Each enterocyte appears toswell owing to a very rapid incorporation of lipid droplets (89, 90, 100), and thereis also a small increase in fluid content as relative wet mass of the intestine in-creases (87). Although evidence is still inconclusive, the lipid droplets likely comefrom the ingested food; alternatively, some of the lipids may stem from fat bodiesin the body of the predator. Although incorporation of the lipid droplets certainlymust account for a major part of enterocytic expansion, Starck & Beese (89) alsohave suggested that increased lymphatic pressure contributes. The increased watercontent of the enterocytes, however, cannot be caused by lymph pressure per sebut would require movements of osmolytes such that osmotically obliged water isdragged along. Clearly, this aspect needs further experimental clarification.

The structure and function of the intestines of birds and mammals are alsoflexible. However, mammals and birds normally feed on a much more regularbasis than do ectothermic vertebrates, and the former generally have a constantrenewal of the gut epithelium. Structural and functional changes occur rapidly

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 18: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

240 WANG � HUNG � RANDALL

after food deprivation in small mammals, whose high metabolism places extrapremium on energy-saving mechanisms. For frequent feeders that are not adjustedto long periods of fasting, prolonged food deprivation or actual starvation maybe more destructive to the gut, as compared to animals that normally experiencelong periods of fasting. The reduction in intestinal mass is due to atrophy, and therestoration of the gut upon subsequent feeding is accomplished by hyperplasia,although hypertrophy also contributes (e.g., 49, 102–105). This is also the case inhibernating mammals, although lower body temperature and metabolism greatlyextend starvation tolerance (91–93, 106). The evolution of endothermy, whichoccurred independently in birds and mammals and has required much higher ratesof nutrient uptake across the gut because of the high metabolism requirements(107), seemingly has led to a structure for which gastrointestinal and digestiveplasticity is energetically more expensive.

The signals that elicit the growth of the gastrointestinal organs are not wellunderstood in nonmammalian vertebrates. In mammals, gastrointestinal growthcan be elicited through luminal, hormonal, neural, and secretory pathways (e.g.,108). Although these regulatory pathways appear phylogenetically old and con-served (e.g., 88), few studies have experimentally investigated the respective rolesof the individual mechanisms. Secor et al. (109) performed systematic infusionsof nutrients into the intestine of fasting animals. Infusion of amino acids or proteindirectly into the intestine increased intestinal mass and transport capacity, whereasinfusion of glucose, lipid, or bile had no effect (109). However, only infusion ofhomogenized rats caused a structural and functional response equivalent to thatelicited by a normal meal (109). Cephalic responses, investigated by allowing thesnake to constrict a prey item, followed by its removal, did not affect the intestine(109). Luminal signals predominantly from protein, therefore, appear sufficient forintestinal expansion and rise in transport capacity during digestion. However, themucosal mass and transport capacity of surgically isolated portions of the intestine(Thiry-Vella loops) increase in voluntarily eating snakes (101), so hormonal and/orneural pathways also seem to contribute to gut expansion. Circulating levels of anumber of regulatory peptides, some released from various gastrointestinal organs,increase dramatically during digestion (110); some of these peptides may serve astrophic factors.

THE METABOLIC RESPONSE TO DIGESTION:SPECIFIC DYNAMIC ACTION

Lavoisier first showed that metabolism increases in response to digestion, andthis metabolic response, documented in all animals investigated, now representsa general phenomenon. The postprandial rise in metabolism, normally referred toas the specific dynamic action of food (SDA), includes the energetic costs associ-ated with the ingestion, digestion, absorption, and assimilation of the food. Thus,the physiological mechanisms that underlie the SDA response may vary among

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 19: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 241

different animals depending on feeding habits, food composition, temperature, andother factors. In the past ten years, SDA in carnivorous reptiles has received muchattention, owing both to its magnitude and its potential to elucidate the large struc-tural and functional changes in the gut. Also, in the animals that exhibit a largeSDA response, the costs of digestion may account for a large proportion of thetotal energy budget, and the metabolic response to feeding becomes ecologicallyrelevant (e.g., 111, 112).

The SDA response is normally characterized as the factorial rise in oxygenuptake and by its duration. Another useful manner of expressing the response isvia the SDA coefficient, which is the integrated metabolic response, calculated incaloric equivalents, relative to the energy consumed. Although some have criticizedthis parameter (113), it provides information on the energetic costs of digestionand allows therefore for a quantitative comparison between digestive responsesunder different environmental parameters or among different types or amountsof food ingested. There does not appear to be an anaerobic contribution to theSDA response, and the entire response therefore is reflected in the rate of oxygenuptake (114–117). The causes and determinants of the SDA response in vertebratesis beyond the scope of the present review, and this area has been summarizedelsewhere (95, 118–120). Also, the respiratory and cardiovascular correlates ofthe high metabolic rate during digestion have been reviewed recently (121, 122).

The effects of fasting duration on the SDA response can be a particularly in-sightful example of the phenotypic plasticity of the gastrointestinal organs. Thus,Overgaard et al. (123) studied the effects of the previous fasting duration on SDAresponse. Upon feeding, animals exhibit elevated intestinal mass and function formany days, suggesting that if the expansion of the gut is energetically expensive,then a second meal, ingested while intestinal function is still elevated, should elicita SDA response smaller than the first response. Overgaard et al. (123) showed thatthe SDA coefficient does not change with a fasting duration between 3–60 days(Figure 7) and that intestinal growth does not constitute a major contributor toSDA response. Fasting duration does not affect the SDA coefficient in skinks orrattlesnakes either (124, 125). A small contribution of intestinal growth was alsosuggested for turtles (126). Collectively, these findings are consistent with the pro-posal that intestinal expansion is structurally simple and energetically cheap (89).Secor (127) subsequently estimated that gastrointestinal upregulation contributesonly 5% of the SDA response in pythons. A recent study of frogs, nevertheless,shows that the rate of digestion of the first meal following three months of esti-vation is slower than for subsequent meals and that reconstitution of the gut mayaccount for this delay (98). The efficiencies of accumulation of various nutrients,however, were not affected (98).

A recent study on snakes has implied that the stomach and the secretion of acidand digestive enzymes are the main contributors to the SDA response (97). In thisstudy, the SDA response to a meal of 25% of the snake’s body weight was reducedby more than half when digesting a liquid meal. The study further showed thatthe response was a third of its normal value when the liquid meal was infused

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 20: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

242 WANG � HUNG � RANDALL

Figure 7 The SDA coefficient in the snake Python molurus following fasting periods of

various duration. The SDA coefficient does not change with the duration of the previous

fasting duration, indicating that structural and functional upregulation of the intestine occurs

at a low energetic cost (modified from Reference 123).

directly into the small intestine. It was estimated therefore that gastric functionscontribute 55% of the SDA response and that the stomach operates on a “pay-before-pumping” principle, in which the snakes must spend endogenous energysources to initiate acid production and other digestive processes before ingestednutrients can be absorbed and used for metabolic pathways. To investigate thispossibility further, we recently used another strategy: tying off of the pylorus,which is the anatomical connection between the stomach and the intestine, sothat the chyme was unable to enter the intestine from the stomach. In the thus-operated animals, the SDA response was completely abolished, whereas sham-operated animals had a normal response (Figure 8). Visual inspection of the preyitems clearly indicated that gastric functions had started digestion, and these datatherefore suggest that secretion of acid and digestive enzymes can proceed at arelatively low energetic cost. That gastric acid secretion has a low energetic costis further supported by the observation that treatment with omeprazole, a specificinhibitor of the H+, K+ ATPase that drives gastric acid secretion, does not affectthe SDA response in another snake species, Boa constrictor (128).

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 21: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 243

Figure 8 The metabolic response to digestion in snakes (Python molurus) in which the

pylorus has been ligated to prevent chime from entering the intestine from the stomach (M.

Andersen, H. Cueto, & T. Wang, unpublished data).

In most animals studied, the SDA response elicited by a given food type in-creases proportionally with meal size, and both the maximal oxygen uptake duringdigestion and the duration of the response increase as meal size increases (e.g.,129–133). In most cases, the SDA coefficient is unaffected by meal size, indicat-ing that the costs of digestion are proportional to the amount of food ingested.Although these data are often interpreted to reflect that it is merely the caloriccontent of the food that determines the SDA, numerous studies have documentedthat protein-rich meals elicit larger metabolic changes than do diets composed offat or carbohydrates. Thus, force-feeding reptiles with fat or carbohydrates elicitsalmost no metabolic response (e.g., 134–138). This would indicate that stimulationof protein synthesis in response to high circulating levels of amino acids (139) isa major contributor to the SDA response (140). The role of protein metabolismin the SDA response (140, 141) is pivotal in fasting catfish, toads, alligators, andpythons in which either systemic infusion of amino acids, or infusion of proteinor amino acids directly into the stomach, leads to a rise in metabolism that iscomparable to that observed during normal feeding (138, 141–144). In catfish andpythons, inhibition of protein synthesis with cyclohexamide completely abolishesthe SDA response (138, 142, 143). If increased protein synthesis is indeed the

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 22: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

244 WANG � HUNG � RANDALL

major contributor to the SDA response, metabolism of all organs should increaseduring the postprandial period, a reasonable suggestion in light of the very highgrowth efficiency of snakes in which some 40–60% of ingested food is directedto growth (123, 145, 146). Obviously, the resulting rates of growth must requireprotein synthesis in all organs.

SUMMARY AND FUTURE DIRECTIONS

Digestive status affects virtually all physiological and behavioral responses, andselective pressure to enhance feeding strategies and digestive processes must besignificant. The ectothermic vertebrates, with their lower metabolic rates, can en-dure prolonged periods of fasting, and many of these species exhibit much morepronounced changes in gastrointestinal organs than are normal in healthy mammals(see also Reference 147). The extreme structural and functional changes in theirdynamic guts make ectothermic vertebrates useful models to explore largely unre-solved issues regarding the interaction and prioritization of physiological functionsamong organ systems. These issues are of basic physiological importance. Suchstudies may contribute to our understanding of the mechanisms that enable organsto adapt to physiological demands. They also may help us to understand the factorsthat in humans can promote intestinal repair following either intestinal resectionsor diseases such as colitis and Crohn’s disease in which there is inflammatorydestruction.

ACKNOWLEDGMENTS

The authors are supported by the Danish Research Council as well as the ResearchGrants Council of Hong Kong Special Administrative Region, People’s Republicof China (project number: CityU RGC1224/02M).

The Annual Review of Physiology is online athttp://physiol.annualreviews.org

LITERATURE CITED

1. Koch AL. 1971. The adaptive responses

of Escherichia coli to a feast or famine

existence. Adv. Microb. Physiol. 6:147–

217

2. Morita RY. 1993. Bioavailability of en-

ergy and the starvation state. In Starvationin Bacteria, ed. S Kjelleberg, pp. 1–53.

New York: Plenum

3. McCauley E, Murdoch WW, Nisbet RM.

1990. Growth, reproduction, and mortal-

ity of daphnia-pulex leydig—life at low

food. Funct. Ecol. 4:505–14

4. Sommer U. 1984. The paradox of the

plankton: Fluctuations of phosphorus

availability maintain diversity of phyto-

plankton in flow-through cultures. Limn.Oceanogr. 29:633–36

5. Mrosovsky N, Sherry DF. 1980. Animal

anorexias. Science 207:837–42

6. Robin JP, Frain M, Sardet C, Groscolas R,

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 23: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 245

Le Maho Y. 1988. Protein and lipid uti-

lization during long-term fasting in em-

peror penguins. Am. J. Physiol. 254:R61–

68

7. Cherel Y, Robin J-P, Le Maho Y. 1988.

Physiology and biochemistry of long-

term fasting birds. Can. J. Zool. 66:159–

66

8. Le Maho Y. 1984. Metabolic adaptations

to prolonged fasting in birds. J. Physiol.79:113–19

9. Groscolas R, Robin J-P. 2001. Long-term

fasting and re-feeding in penguins. Comp.Biochem. Physiol. 128A:645–55

10. Brett JR. 1995. Energetics. In Physiolog-ical Ecology of Pacific Salmon, ed. C

Groot, L Margolis, WC Clarke, pp. 1–68.

Vancouver: UBC Press

11. Cahill GF Jr, Herrera MG, Morgan AP,

Soeldner JS, Steinke J, et al. 1966.

Hormone-fuel interrelationships during

fasting. J. Clin. Invest. 45:1751–69

12. Cahill GF Jr. 1976. Starvation in man.

Clin. Endocrinol. Metab. 5:397–415

13. Henry CJK. 1990. Body mass index and

the limits of human survival. Eur. J. Clin.Nutr. 44:329–35

14. Owen OE, Tappy L, Mozzoli MA, Smal-

ley KJ. 1990. Acute starvation. In TheMetabolic and Molecular Basis of Ac-quired Disease, ed. RD Cohen, B Lewis,

KGMM Alberti, AM Denman, pp. 550–

70. London: Bailliere Tinall

15. Cherel Y, Groscolas R. 1999. Relation-

ships between nutrient storage and nutri-

ent utilisation in long-term fasting birds

and mammals. In Proc. 22nd Int. Or-nithol. Congr., Durban, eds. NJ Adams,

RH Slotow, pp. 17–34. Johnannesburg:

BirdLife South Africa

16. Nagy KA, Girard IA, Brown TK. 1999.

Energetics of free-ranging mammals, rep-

tiles, and birds. Annu. Rev. Nutr. 19:247–

77

17. Schmidt J. 1923. Breeding places and mi-

gration of the eel. Nature 111:51–54

18. Van Ginneken VJT, Antonissen E, Muller

UK, Booms R, Eding E, et al. 2005.

Eel migration to the Sargasso: remarkably

high swimming efficiency and low energy

costs. J. Exp. Biol. 208:1329–35

19. Thouzeau C, Robin J-P, Le Maho Y, Han-

drich Y. 1999. Body reserve dynamics and

energetics of barn owls during fasting in

the cold. J. Comp. Physiol. 169B:612–

20

20. Jobling M. 1980. Effects of starvation on

proximate chemical composition and en-

ergy utilization of plaice, Pleuronectesplatessa L. J. Fish Biol. 17:325–34

21. Mendez G, Wieser W. 1993. Metabolic re-

sponses to food deprivation and refeeding

in juveniles of Rutilus rutilus (Teleostei:

Cyprinidae). Environ. Biol. Fishes 36:73–

81

22. Cook JT, Sutterlin AM, McNiven MA.

2000. Effect of food deprivation on oxy-

gen consumption and body composition

of growth-enhanced transgenic Atlantic

salmon Salmo salar. Aquaculture 188:47–

63

23. Hervant F, Mathieu J, Durand J. 2001.

Behavioural, physiological and metabolic

responses to long-term starvation and

refeeding in a blind cave-dwelling (Pro-teus anguinus) and a surface-dwelling

(Euproctus asper) salamander. J. Exp.Biol. 204:269–81

24. Figueiredo-Garutti M, Navarro LI,

Capilla E, Souza RH, Moraes G, et al.

2002. Metabolic changes in Bryconcephalus (Teleostei, Characidae) during

post-feeding and fasting. Comp. Biochem.Physiol 132A:467–76

25. Navarro I, Gutierrez J. 1995. Fasting and

starvation. In Biochemistry and Molec-ular Biology of Fishes, Vol. 4, ed. PW

Hochachka, TP Mommsen, pp. 393–434.

Amsterdam: Elsevier Science B.V.

26. Hung CY. 2005. Survival strategies ofcommon carp, Cyprinus carpio, to pro-longed starvation and hypoxia. PhD the-

sis. City Univ. Hong Kong. 254 pp.

27. Takeuchi T, Watanabe T. 1982. The ef-

fects of starvation and environmental tem-

perature on proximate and fatty acid

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 24: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

246 WANG � HUNG � RANDALL

compositions of carp and rainbow trout.

Bull. Jap. Soc. Sci. Fish 48:1307–16

28. Shimeno S, Kheyyali D, Takeda M. 1990.

Metabolic adaptation to prolonged star-

vation in carp. Nippon Suisan Gakkaishi56:35–41

29. Shimeno S, Saida Y, Tabata T. 1996.

Response of hepatic NAD- and NADP-

isocitrate dehydrogenase activities to sev-

eral dietary conditions in fishes. NipponSuisan Gakkaishi 62:642–46

30. Shimeno S, Shikata T. 1993. Effects of ac-

climation temperature and feeding rate on

carbohydrate-metabolizing enzyme activ-

ity and lipid content of common carp. Nip-pon Suisan Gakkaishi 59:661–66

31. Jezierska B, Hazel JR, Gerking SD. 1982.

Lipid mobilization during starvation in

the rainbow trout, Salmo gairdneri R.,

with attention to fatty acids. J. Fish Biol.21:681–92

32. Nagai M, Ikeda S. 1971. Carbohydrate

metabolism in fish—I. Effects of starva-

tion and dietary composition on the blood

glucose level and the hepatopancreatic

glycogen and lipid content in carp. Bull.Jap. Soc. Scient. Fish 37:404–409

33. Moon TW, Foster GD. 1995. Tissue

carbohydrate metabolism, gluconeogene-

sis and hormonal and environmental in-

fluences. In Biochemistry and Molecu-lar Biology of Fishes, Vol. 4, ed. PW

Hochachka, TP Mommsen, pp. 254–96.

Amsterdam: Elsevier Science B.V.

34. Threlkeld ST. 1976. Starvation and the

size structure of zooplankton communi-

ties. Freshw. Biol. 6:489–96

35. Peters RH. 1983. The Ecological Impli-cations of Body Size. Cambridge, MA:

Cambridge Univ. Press

36. Calder WA. 1994. When do humming-

birds use torpor in nature? Physiol. Zool.67:1051–76

37. Øritsland NA. 1990. Starvation survival

and body composition in mammals with

particular reference to Homo sapiens.

Bull. Math. Biol. 52:643–55

38. Keys A, Brozek J, Hennschel A,

Michelsen O, Taylor HL. 1950. The Bi-ology of Human Starvation. Minneapolis:

Univ. Minn. Press

39. Merkt JR, Taylor CR 1994. “Metabolic

switch” for desert survival. Proc. Natl.Acad. Sci. USA 91:12313–16

40. Ehrhardt N, Heldmaier G, Exner C. 2005.

Adaptive mechanisms during food restric-

tion in Acomys russatus: the use of torpor

for desert survival. J. Comp. Physiol. B175: 193–200

41. Poon WL. 2005. In vivo changes in com-mon carp (Cyprinus carpio L.) liver dur-ing hypoxia at the molecular and cellularlevels. PhD thesis. City Univ. Hong Kong.

182 pp.

42. Brown HO, Levine ML, Lipkin M. 1963.

Inhibition of intestinal epithelial cell re-

newal and migration induced by starva-

tion. Am. J. Physiol. 205:868–872

43. Habold C, Chevalier C, Dunel-Erb S,

Foltzer-Jourdainne C, Le Maho Y, Lignot

J-H. 2004. Effects of fasting and refeeding

on jejunal morphology and cellular activ-

ity in rats in relation to depletion of body

stores. Scand. J. Gastroenterol. 39:531–

39

44. Iwakiri R, Gotoh Y, Noda T, Sugihara H,

Fujimoto K, et al. 2001. Programmed cell

death in rat intestine: effect of feeding and

fasting. Scand. J. Gastroenterol. 36:39–

47

45. Habold C, Foltzer-Jourdainne C, Le Maho

Y, Lignot J-H. 2005. Intestinal apop-

totic changes linked to metabolic status

in fasted and refed rats. Eur. J. Physiol.In press

46. Robin J-P, Boucentet L, Chillet P, Grisco-

las R. 1998. Behavioral changes in fasting

emperor penguins: evidence for a “refeed-

ing signal” linked to metabolic shifts. Am.J. Physiol. 43:R746–53

47. Koubi HE, Robin JP, Dewasmes G, Le

Maho Y, Frutoso J, Minaire Y. 1991.

Fasting-induced rise in locomotor activ-

ity in rats coincides with increased protein

utilization. Physiol. Behav. 50:337–43

48. Bertile F, Oudart H, Criscuolo F, Le Maho

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 25: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 247

Y, Raclot T. 2003. Hypothalamic gene

expression in long-term fasted rats: rela-

tionship with body fat. Biochem. Biophys.Res. Commun. 303:1106–13

49. Dunel-Erb S, Chevalier C, Laurent P,

Bach A, Decrock F, Le Maho Y. 2001.

Restoration of the jejunal mucosa in rats

refed after prolonged fasting. Comp. Bio-chem. Physiol. 129A:933–47

50. Geiser F. 2004. Metabolic rate and body

temperature reduction during hibernation

and daily torpor. Annu. Rev. Physiol.66:239–74

51. Lovegrove BG, Raman J, Perrin MR.

2001. Daily torpor in elephant shrews

(Macroscelidea: Elephantulus spp.) in

response to food deprivation. J. Comp.Physiol. 171B:11–21

52. Bech C, Abe AS, Steffensen JF, Berger

M, Bicudo JEPW. 1997. Torpor in three

species of Brazilian hummingbirds under

semi-natural conditions. Condor 99:780–

88

53. Powers DR, Brown AR, Van Hook JA.

2003. Influence of normal daytime fat de-

position on laboratory measurements of

torpor use in territorial versus nonterri-

torial hummingbirds. Physiol. Biochem.Zool. 76:389–97

54. Prinzinger R, Schleucher E, Preßmar A.

1992. Long-term telemetry of body tem-

perature with synchronous measurement

of metabolic rate in torpid and non-

torpid blue naped mousebirds (Urocoliusmacrourus). J. Ornithol. 133:446–50

55. Geiser F, Ruf T. 1995. Hibernation ver-

sus daily torpor in mammals and birds:

physiological variables and classification

of torpor patterns. Physiol. Zool. 68:935–

66

56. Geiser F. 1998. Evolution of daily torpor

and hibernation in birds and mammals:

importance of body size. Clin. Exp. Phar-macol. Physiol. 25:736–40

57. Schleucher E. 2004. Torpor in birds: tax-

onomy, energetics, and ecology. Physiol.Biochem. Zool. 77:942–49

58. Graf R, Krishna S, Heller HC. 1989. Reg-

ulated nocturnal hypothermia induced in

pigeons by food deprivation. Am. J. Phys-iol. 256:R733–38

59. McKechnie AE, Lovegrove BG. 2002.

Avian facultative hypothermic responses:

a review. Condor 104:704–24

60. Severinsen T, Munch IC. 1999. Body core

temperature during food restriction in rats.

Acta. Physiol. Scand. 165:299–305

61. Mesteig K, Tyler NJ, Blix AS. 2000. Sea-

sonal changes in heart rate and food intake

in reindeer (Rangifer tarandus tarandus).

Acta. Physiol. Scand. 170:145–51

62. Arnold W, Ruf T, Reimoser S, Tataruch F,

Onderscheka K, Schober F. 2004. Noctur-

nal hypometabolism as an overwintering

strategy of red deer (Cervus elaphus). Am.J. Physiol. 286:R174–81

63. Genin F, Perret M. 2003. Daily hypother-

mia in captive grey mouse lemurs (Mi-crocebus murinus): effects of photope-

riod and food restriction. Comp. Biochem.Physiol. 136B:71–81

64. Merola-Zwartjes M, Ligon JD. 2000. Eco-

logical energetics of the Puerto Rican

tody: heterothermy, torpor and intraisland

variation. Ecology 81:990–1002

65. Butler PJ, Woakes AJ. 2001. Seasonal hy-

pothermia in a large migrating bird: sav-

ing energy or fat deposition? J. Exp. Biol.204:1361–67

66. Van Dijk PLM, Staaks G, Hardewig I.

2002. The effect of fasting and refeeding

on temperature preference, activity and

growth of roach, Rutilus rutilus. Oecolo-gia 130:496–504

67. Brown RP, Griffin S. 2005. Lower se-

lected body temperatures after food de-

privation in the lizard Anolis carolinensis.

J. Thermal Biol. 30:79–83

68. Sclafani A, Rendel A. 1978. Food

deprivation-induced activity in dietary

obese, dietary lean and normal-weight

rats. Behav. Biol. 24:220–28

69. Pierre PJ, Skjoldager P, Bennett AJ, Ren-

ner MJ. 2001. A behavioral characteriza-

tion of the effects of food deprivation on

food and nonfood object interaction: an

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 26: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

248 WANG � HUNG � RANDALL

investigation of the information-gathering

functions of exploratory behavior. Phys-iol. Behav. 72:189–97

70. Morin LP. 1986. Environment and ham-

ster reproduction: responses to phase-

specific starvation during estrous cycle.

Am. J. Physiol. 251:R663–69

71. Jones JE, Wade GN. 2002. Acute fasting

decreases sexual receptivity and neural es-

trogen receptor-α in female rats. Physiol.Behav. 77:19–25

72. Tessier AJ, Henry LL, Goulden CE, Du-

rand MW. 1983. Starvation in daphnia:

energy reserves and reproductive alloca-

tion. Limnol. Oceanogr. 28:667–76

73. Kirk KL. 1997. Life-history responses to

variable environments: starvation and re-

production in planktonic rotifers. Ecology78:434–41

74. Boetius I, Boetius J. 1985. Lipid and

protein content in Anguilla anguilla dur-

ing growth and starvation. Dana 4:1–

17

75. Olivereau M, Olivereau JM. 1997. Long-

term starvation in the European eel: gen-

eral effects and responses of pituitary

growth hormone-(GH) and somatolactin-

(SL) secreting cells. Fish Physiol. Bio-chem. 17:261–69

76. Weatherley AH, Gill HS. 1985. Dynam-

ics of increase in muscle fibers in fishes in

relation to size and growth. Experientia41:353–54

77. French CJ, Hochachka PW, Mommsen

TP. 1983. Metabolic organization of liver

during spawning migration of sockeye

salmon. Am. J. Physiol. 245:R827–30

78. Chichery MP, Chichery R. 1992. Be-

havioural and neurohistological changes

in aging Sepia. Brain Res. 574:77–

84

79. Piersma T, Lindstrom A. 1997. Rapid re-

versible changes in organ size as a compo-

nent of adaptive behaviour. Trends Ecol.Evol. 12:134–38

80. Starck JM. 2005. Structural flexibility

of the digestive system of tetrapods—

patterns and processes at the cellular and

tissue levels. In Physiological and Eco-logical Adaptations to Feeding in Verte-brates, ed. JM Starck, T Wang, pp. 175–

200. New Delhi: Sci. Publ. Inc.

81. Cant JP, McBide BW, Croom WJ. 1996.

The regulation of intestinal metabolism

and its impact on whole animal energetics.

J. Anim. Sci. 74:2541–53

82. Greene HW. 1983. Dietary correlates of

the origin and radiation of snakes. Am.Zool. 23:431–41

83. Shine R, Harlow PS, Keogh JS, Boeadi.

1998. The influence of sex and body size

on food habits of a giant tropical snake,

Python reticulatus. Func. Ecol. 12:248–

58

84. Secor SM, Stein ED, Diamond J. 1994.

Rapid upregulation of snake intestine

in response to feeding: a new model

of intestinal adaptation. Am. J. Physiol.29:G695–705

85. Secor SM, Diamond J. 1995. Adaptive

responses to feeding in Burmese pythons:

pay before pumping. J. Exp. Biol. 198:

1313–25

86. Secor SM, Diamond J. 1998. A vertebrate

model of extreme physiological regula-

tion. Nature 395:659–62

87. Holmberg A, Joanna K, Persson A, Jensen

J, Wang T, Holmgren S. 2002. Effects of

digestive status on the reptilian gut. Comp.Biochem. Physiol. 133A:499–18

88. Holmgren S, Holmberg A. 2005. Control

of gut motility and secretion in fasting and

fed non-mammalian vertebrates. In Phys-iological and Ecological Adaptations toFeeding in Vertebrates, ed. JM Starck, T

Wang, pp. 325–62. New Delhi: Sci. Publ.

Inc.

89. Starck JM, Beese K. 2001. Structural

flexibility of the intestine of Burmese

python in response to feeding. J. Exp. Biol.204:325–35

90. Lignot J-H, Helmstetter C, Secor SM.

2005. Postprandial morphological re-

sponse of the intestinal epithelium of

the Burmese python (Python molurus).

Comp. Biochem. Physiol. 141A:280–91

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 27: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 249

91. Carey HV. 1990. Seasonal changes

in mucosal structure and function in

ground squirrel intestine. Am. J. Physiol.259:R385–92

92. Carey HV. 1995. Gut feelings about hi-

bernation. News Physiol. Sci. 10:55–61

93. Carey HV. 2005. Gastrointestinal re-

sponses to fasting in mammals: lessons

from hibernators. In Physiological andEcological Adaptations to Feeding in Ver-tebrates, ed. JM Starck, T Wang, p. 229–

254. New Delhi: Sci. Publ. Inc.

94. Carey HV, Martin SL. 1996. Preservation

of intestinal gene expression during hiber-

nation. Am. J. Physiol. 271:804–13

95. Secor SM. 2001. Regulation of di-

gestive performance: a proposed adap-

tive response. Comp. Biochem. Physiol.128A:563–75

96. McLeese JM, Moon TW. 1989. Seasonal

changes in the intestinal mucosa of the

winter flounder, Pseudopleuronectesamericanus (Wallbaum) from Passam-

quoddy Bay, New Brunswick. J. FishBiol. 35:381–93

97. Secor SM. 2005.Physiological responses

to feeding, fasting and estivation for an-

urans. J. Exp. Biol. 208:2595–609

98. Cramp RL, Franklin CE. 2003. Is re-

feeding efficiency compromised by pro-

longed starvation during aestivation in the

green striped burrowing frog, Cycloranaalboguttata? J. Exp. Zool. 300:126–32

99. Jackson K, Perry G. 2000. Changes in in-

testinal morphology following feeding in

the brown treesnake, Boiga irregularis. J.Herpetol. 34:459–62

100. Starck JM, Beese K. 2002. Structural flex-

ibility of the small intestine and liver of

Garter snakes in response to feeding and

fasting. J. Exp. Biol. 205:1377–88

101. Secor SM, Whang EE, Lane JS, Ashley

SW, Diamond J. 2000. Luminal and sys-

temic signals trigger intestinal adaptation

in the juvenile python. Am. J. Physiol.279:G1177–87

102. Altmann GG. 1972. Influence of starva-

tion and refeeding on mucosal size and

epithelial renewal in the rat small intes-

tine. Am. J. Anat. 133:391–400

103. Starck JM, Kloss E. 1995. Structural re-

sponses of Japanese quail intestine to dif-

ferent diets. Dtsch. Tierarztl. Wochenschr.

102:146–50

104. Hume ID, Biebach H. 1996. Digestive

tract function in the long distance mi-

gratory garden warbler, Sylvia borin. J.Comp. Physiol. B 166:388–95

105. Ferrari RP, Carey HV. 2000. Intestinal

transport during fasting and malnutrition.

Annu. Rev. Nutr. 20:195–19

106. Hume ID, Beiglbock C, Ruf T, Frey-

Ross F, Bruns U, Arnold W. 2002. Sea-

sonal changes in morphology and function

of the gastrointestinal tract of free-living

alpine marmots (Marmota marmota). J.Comp. Physiol. B 172:197–207

107. Karasov WH, Diamond JM. 1985. Diges-

tive adaptations for fueling the cost of en-

dothermy. Science 228:202–4

108. Johnson LR. 1997. Gastrointestinal Phys-iology. St. Louis, MO: Mosby. 1023 pp.

109. Secor SM, Lane JS, Whang EE, Ashley

SW, Diamond J. 2002. Luminal nutri-

ent signals for intestinal adaptation in py-

thons. Am. J. Physiol. 283:G1298–1309

110. Secor SM, Fehsenfeld D, Diamond J,

Adrian TE. 2001. Responses of python

gastrointestinal regulatory peptides to

feeding. Proc. Natl. Acad. Sci. USA 98:

13637–42

111. Peterson CC, Walton BM, Bennett AF.

1999. Metabolic costs of growth in free-

living garter snakes and the energy bud-

gets of ectotherms. Func. Ecol. 13:500–

507

112. McCue MD, Lillywhite HB. 2002. Oxy-

gen consumption and the energetics

of island-dwelling Florida cottonmouth

snakes. Physiol. Biochem. Zool. 75:165–

78

113. Beaupre SJ. 2005. Ratio representations

of specific dynamic action (mass-specific

SDA and SDA coefficient) do not stan-

dardize for body mass and meal size.

Physiol. Biochem. Zool. 78:126–31

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 28: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

250 WANG � HUNG � RANDALL

114. Andersen JB, Wang T. 2003. Cardiores-

piratory effects of forced activity and

digestion in toads. Physiol. Biochem.Zool. 76:459–70

115. Overgaard J, Busk M, Hicks JW, Jensen

FB, Wang T. 1999. Respiratory conse-

quences of feeding in the snake Pythonmolorus. Comp. Biochem. Physiol. 124A:

361–67

116. Busk M, Jensen FB, Wang T. 2000. The

effects of feeding on blood gases, acid-

base parameters and selected metabolites

in the bullfrog Rana catesbeiana. Am. J.Physiol. 278:R185–95

117. Busk M, Overgaard J, Hicks JW, Bennett

AF, Wang T. 2000. Effects of feeding on

arterial blood gases in the American alli-

gator, Alligator mississippiensis. J. Exp.Biol. 203:3117–24

118. Jobling M. 1981. The influences of feed-

ing on the metabolic rate of fishes: a short

review. J. Fish Biol. 18:385–400

119. Wang T, Zaar M, Arvedsen S, Vedel C,

Overgaard J. 2002. Effects of temperature

on the metabolic response to feeding in

Python molurus. Comp. Biochem. Phys-iol. 133A:519–27

120. Andrade DV, Abe AS, Cruz-Neto AP,

Wang T. 2005. Specific dynamic action

in ectothermic vertebrates: a general re-

view on the determinants of the metabolic

responses to digestion in fish, amphibians

and reptiles. In Adaptations in Physiolog-ical and Ecological Adaptations to Feed-ing in Vertebrates, ed. JM Starck, T Wang,

pp. 305–24. New Delhi: Sci. Publ. Inc.

121. Wang T, Busk M, Overgaard J. 2001. The

respiratory consequences of feeding in

amphibians and reptiles. Comp. Biochem.Physiol. 128A:533–47

122. Wang T, Andersen J, Hicks JW. 2005. Ef-

fects of digestion on the respiratory and

cardiovascular physiology of amphibians

and reptiles. In Adaptations in Physiolog-ical and Ecological Adaptations to Feed-ing in Vertebrates, ed. JM Starck, T Wang,

pp. 279–303. New Delhi: Sci. Publ. Inc.

123. Overgaard J, Andersen JB, Wang T.

2002. The effects of fasting duration

on the metabolic response to feeding in

Python molurus: an evaluation of the

energetic costs associated with gastroin-

testinal growth and upregulation. Physiol.Biochem. Zool. 75:360–68

124. Iglesias S, Thompson MB, Seebacher F.

2003. Energetic cost of a meal in a fre-

quent feeding lizard. Comp. Biochem.Physiol. 135A:377–82

125. Zaidan F, Beaupre SJ. 2003. Effects of

body mass, meal size, fast length, and

temperature on specific dynamic action

in the timber rattlesnake (Crotalus hor-ridus). Physiol. Biochem. Zool. 76:447–

58

126. Hailey A. 1998. The specific dynamic ac-

tion of the omnivorous tortoise Kinixysspekii in relation to diet, feeding pattern,

and gut passage. Physiol. Zool. 71:57–

66

127. Secor SM. 2003. Gastric function and its

contribution to the postprandial metabolic

response of the Burmese python Pythonmolurus. J. Exp. Biol. 206:1621–30

128. Andrade DV, Toledo LP, Abe AS, Wang

T. 2004. Ventilatory compensation of the

alkaline tide during digestion in the snake

Boa constrictor. J. Exp. Biol. 207:1379–

85

129. Soofiani NM, Hawkins AD. 1982. Ener-

getic costs at different levels of feeding

in juvenile cod, Gadus morhua L. J. FishBiol. 21:577–92

130. Andrade DV, Cruz-Neto AP, Abe AS.

1997. Meal size and specific dynamic

action in the rattlesnake Crotalus duris-sus (Serpentes: Viperidae). Herpetologica53:485–93

131. Boyce SJ, Clarke A. 1997. Effect of body

size and ration on specific dynamic action

in the Antarctic plunderfish, Harpagiferantarcticus Nybelin 1947. Physiol. Zool.70:679–90

132. Secor SM, Diamond J. 1997. Effects of

meal size on postprandial responses in ju-

venile Burmese pythons (Python molu-rus). Am. J. Physiol. 272:R902–12

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 29: Adaptación Ciclo Ayuno Alimentación

18 Jan 2006 14:21 AR ANRV265-PH68-09.tex XMLPublishSM(2004/02/24)P1: OKZ /OKZ P2:OJO

PHYSIOLOGY OF FOOD DEPRIVATION 251

133. Secor SM, Faulkner AC. 2002. Effects of

meal size, meal type, body temperature,

and body size on the specific dynamic

action of the marine toad, Bufo marinus.

Physiol. Biochem. Zool. 75:557–71

134. Benedict FG. 1932. The Physiology ofLarge Reptiles with Special Reference tothe Heat Production of Snakes, Tortoises,Lizards, and Alligators. Washington, DC:

Carnegie Inst. Publ. 254 pp.

135. Jobling M, Davies PS. 1980. Effects of

feeding on metabolic rate, and the specific

dynamic action in plaice, Pleuronectesplatessa L. J. Fish Biol. 16:629–38

136. Coulson RA, Hernandez T. 1983. Alliga-

tor metabolism: studies on chemical re-

actions in vivo. Comp. Biochem. Physiol.74A:1–182

137. Somanath B, Palavesam A, Lazarus S,

Ayyapan M. 2000. Influence of nutrient

source on specific dynamic action of pearl

spot, Etroplus suratensis (Bloch). Naga23:15–17

138. McCue MD, Bennett AF, Hicks JW. 2005.

The effect of meal composition on spe-

cific dynamic action in burmese pythons

(Python molurus). Physiol. Biochem.Zool. 78:182–92

139. Houlihan DF. 1991. Protein turnover in

ectotherms and its relationships to en-

ergetics. Adv. Comp. Physiol. Biochem.7:1–43

140. Ashworth A. 1969. Metabolic rates dur-

ing recovery from protein-calorie malnu-

trition: the need for a new concept of spe-

cific dynamic action. Nature 223:407–409

141. Coulson RA, Hernadez T. 1979. Increase

in metabolic rate of the alligator fed pro-

teins or amino acids. J. Nutr. 109:538–

50

142. Brown CR, Cameron JN. 1991. The rela-

tionship between specific dynamic action

(SDA) and protein synthesis rates in the

channel catfish. Physiol. Zool. 64:298–

309

143. Brown CR, Cameron JN. 1991. The in-

duction of specific dynamic action in

channel catfish by infusion of essential

amino acids. Physiol. Zool. 64:276–97

144. Wang T, Burggren WW, Nobrega E. 1995.

Metabolic, ventilatory, and acid-base re-

sponses associated with specific dynamic

action in the toad Bufo marinus. Physiol.Zool. 68:192–205

145. Vinegar A, Hutchison VH, Dowling HG.

1970. Metabolism, energetics and ther-

moregulation during brooding of snakes

of the genus Python (Reptilia, Boidae).

Zoologica 55:19–48

146. Secor SM, Diamond J. 1997. Determi-

nants of the postfeeding metabolic re-

sponse of burmese pythons, Python molu-rus. Physiol. Zool. 70:202–12

147. Pennisi E. 2005. The dynamic gut. Science307:1896–99

148. Bauer M, Hamm AC, Bonaus M, Jacob A,

Jaekel J, et al. 2004. Starvation response in

mouse liver shows strong correlation with

life-span-prolonging processes. Physiol.Genomics 17:230–44

149. Zhang J, Underwood LE, D’Ercole AJ.

2001. Hepatic mRNAs up-regulated by

starvation: an expression profile deter-

mined by suppression subtractive hy-

bridization. FASEB J. 15:1261–63

150. Rios FS, Kalinin AL, Fernandes MN,

Rantin FT. 2004. Changes in gut gross

morphology of traira, Hoplias malabari-cus (Teleostei, Erythrinidae) during long-

term starvation and after refeeding. Braz.J. Biol. 64:683–89

151. Van Dijk PLM, Hardewig I, Holker F.

2005. Energy reserves during food depri-

vation and compensatory growth in juve-

nile roach: the importance of season and

temperature. J. Fish Biol. 66:167–181

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 30: Adaptación Ciclo Ayuno Alimentación

P1: JRX

January 18, 2006 10:52 Annual Reviews AR265-FM

Annual Review of PhysiologyVolume 68, 2006

CONTENTS

Frontispiece—Watt W. Webb xiv

PERSPECTIVES, David L. Garbers, Editor

Commentary on the Pleasures of Solving Impossible Problems

of Experimental Physiology, Watt W. Webb 1

CARDIOVASCULAR PHYSIOLOGY, Jeffrey Robbins, Section Editor

Cardiac Regeneration: Repopulating the Heart, Michael Rubartand Loren J. Field 29

Endothelial-Cardiomyocyte Interactions in Cardiac Development

and Repair, Patrick C.H. Hsieh, Michael E. Davis,Laura K. Lisowski, and Richard T. Lee 51

Protecting the Pump: Controlling Myocardial Inflammatory Responses,

Viviany R. Taqueti, Richard N. Mitchell, and Andrew H. Lichtman 67

Transcription Factors and Congenital Heart Defects, Krista L. Clark,Katherine E. Yutzey, and D. Woodrow Benson 97

CELL PHYSIOLOGY, David L. Garbers, Section Editor

From Mice to Men: Insights into the Insulin Resistance Syndromes,

Sudha B. Biddinger and C. Ronald Kahn 123

LXRs and FXR: The Yin and Yang of Cholesterol and Fat Metabolism,

Nada Y. Kalaany and David J. Mangelsdorf 159

ECOLOGICAL, EVOLUTIONARY, AND COMPARATIVE PHYSIOLOGY,Martin E. Feder, Section Editor

Design and Function of Superfast Muscles: New Insights into the

Physiology of Skeletal Muscle, Lawrence C. Rome 193

The Comparative Physiology of Food Deprivation: From Feast to Famine,

Tobias Wang, Carrie C.Y. Hung, and David J. Randall 223

Oxidative Stress in Marine Environments: Biochemistry and

Physiological Ecology, Michael P. Lesser 253

GASTROINTESTINAL PHYSIOLOGY, John Williams, Section Editor

Brainstem Circuits Regulating Gastric Function, R. Alberto Travagli,Gerlinda E. Hermann, Kirsteen N. Browning, and Richard C. Rogers 279

vii

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.

Page 31: Adaptación Ciclo Ayuno Alimentación

P1: JRX

January 18, 2006 10:52 Annual Reviews AR265-FM

viii CONTENTS

Interstitial Cells of Cajal as Pacemakers in the Gastrointestinal Tract,

Kenton M. Sanders, Sang Don Koh, and Sean M. Ward 307

Signaling for Contraction and Relaxation in Smooth Muscle of the Gut,

Karnam S. Murthy 345

NEUROPHYSIOLOGY, Richard Aldrich, Section Editor

CNG and HCN Channels: Two Peas, One Pod, Kimberley B. Cravenand William N. Zagotta 375

RENAL AND ELECTROLYTE PHYSIOLOGY, Gerhard H. Giebisch, Section Editor

Claudins and Epithelial Paracellular Transport, Christina M. Van Itallieand James M. Anderson 403

Role of FXYD Proteins in Ion Transport, Haim Gartyand Steven J.D. Karlish 431

Sgk Kinases and Their Role in Epithelial Transport, Johannes Loffing,Sandra Y. Flores, and Olivier Staub 461

The Association of NHERF Adaptor Proteins with G Protein–Coupled

Receptors and Receptor Tyrosine Kinases, Edward J. Weinman,Randy A. Hall, Peter A. Friedman, Lee-Yuan Liu-Chen,and Shirish Shenolikar 491

RESPIRATORY PHYSIOLOGY, Richard C. Boucher, Jr., Section Editor

Stress Transmission in the Lung: Pathways from Organ to Molecule,

Jeffrey J. Fredberg and Roger D. Kamm 507

Regulation of Normal and Cystic Fibrosis Airway Surface Liquid Volume

by Phasic Shear Stress, Robert Tarran, Brian Button,and Richard C. Boucher 543

Chronic Effects of Mechanical Force on Airways, Daniel J. Tschumperlinand Jeffrey M. Drazen 563

The Contribution of Biophysical Lung Injury to the Development

of Biotrauma, Claudia C. dos Santos and Arthur S. Slutsky 585

SPECIAL TOPIC, TRP CHANNELS, David E. Clapham, Special Topic Editor

An Introduction to TRP Channels, I. Scott Ramsey, Markus Delling,and David E. Clapham 619

Insights on TRP Channels from In Vivo Studies in Drosophila,

Baruch Minke and Moshe Parnas 649

Permeation and Selectivity of TRP Channels, Grzegorz Owsianik,Karel Talavera, Thomas Voets, and Bernd Nilius 685

TRP Channels in C. elegans, Amanda H. Kahn-Kirbyand Cornelia I. Bargmann 719

Ann

u. R

ev. P

hysi

ol. 2

006.

68:2

23-2

51. D

ownl

oade

d fr

om a

rjou

rnal

s.an

nual

revi

ews.

org

by U

NIV

ER

SIT

Y O

F IL

LIN

OIS

- C

HIC

AG

O o

n 09

/09/

06. F

or p

erso

nal u

se o

nly.