Download Developing Standardized Behavioral Tests for

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Limbic system wikipedia , lookup

Aging brain wikipedia , lookup

Metastability in the brain wikipedia , lookup

Donald O. Hebb wikipedia , lookup

Neuroethology wikipedia , lookup

Eyeblink conditioning wikipedia , lookup

State-dependent memory wikipedia , lookup

Theory of reasoned action wikipedia , lookup

Psychological behaviorism wikipedia , lookup

Behaviorism wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Optogenetics wikipedia , lookup

Theory of planned behavior wikipedia , lookup

Behavior analysis of child development wikipedia , lookup

Neuroeconomics wikipedia , lookup

Environmental enrichment wikipedia , lookup

Transcript
Developing Standardized Behavioral Tests for Knockout and Mutant Mice
Richard E. Brown, Lianne Stanford, and Heather M. Schellinck
Introduction
T
he ability to produce transgenic, knockout, and mutant
mice provides new animal models for the genetic
analysis of behavior and its underlying neural mechanisms (Chen and Tonegawa 1997; Wehner et al. 1996).
Knockout and transgenic mice have been developed as
animal models for the study of human diseases such as
Alzheimer's disease (Flood and Morley 1998; Nalbantoglu
et al. 1997), Down's syndrome (Schuchmann et al. 1998;
Tremml et al. 1998), Huntington's disease (Bates et al. 1997),
schizophrenia (Mohn et al. 1999), and other neurobiologic
disorders (Nelson and Young 1998). Transgenic mice can
also be used to study the behavioral effects of altering specific
neurochemical receptors such as for N-methyl-D-aspartate
and a-amino-3-hydroxy-5-methyl-4-isoxasole-propionate
(Sprengel and Single 1999), metabotropic glutamate (Galani
et al. 1997), serotonin (Parks et al. 1998), or dopamine
(Steiner et al. 1997).
In many genetically altered mice, the most noticeable
difference between the transgenic mouse and its background
strain is a change in behavior. As indicated by Nelson and
Young (1998, p. 453), however, "correlations among behavioral assessments of knockouts are difficult to make because
no standardized behavioral tests are available." Crawley and
Pay lor (1997) and Crawley (1999) list a number of behavioral test paradigms that can be used to test genetically altered
mice, and Nelson and Young (1998) list a number of genetically altered mice that can be tested in these paradigms.
Although early studies of transgenic, mutant, and knockout
mice included very few behavioral tests, it is becoming more
common to use a test battery approach when examining the
behavioral phenotypes of these mice (Brunner et al. 1999).
In this article, we review test batteries for examining sensory
and motor behavior, species-typical behavior, learning and
memory, and development and aging in transgenic mice.
Before these test batteries are used, however, it is important
to consider nongenetic variables such as animal housing and
experience and procedural variables, which may interact with
genetic differences to alter behavior.
Behavioral genetic studies on mice have been conducted
for more than 40 yr (Sprott and Staats 1975,1981). We have
Richard E. Brown, Ph.D., is a Professor, Lianne Stanford, M. A., is a doctoral
candidate, and Heather M. Schellinck, Ph.D., is a Senior Instructor in the
Department of Psychology and Neuroscience Institute, Dalhousie University, Nova Scotia, Canada.
Volume 4 1 , Number 3
2000
not attempted a thorough review of all studies but have
selected some recent representative studies to illustrate our
points.
Questions That Must Be Answered before
Using a Behavioral Battery
To conclude that differences in behavior between mutant and
wild-type mice are due to genetic differences, one must control for the effects of confounding variables. The apparent
effects of genetic manipulation on behavior may depend on
factors other than those associated with the genetic differences in the subjects. Often there is not one best way to deal
with these variables, thus we recommend that researchers
ensure the inclusion of detailed information in the Methods
sections of published papers. In particular, before testing
mice in a standardized behavioral paradigm, a number of
questions must be answered about the background of the
animals and the conditions under which they are maintained
(Table 1).
First, what is the source of the animals? Animals from
different breeding facilities may have been reared under different conditions and procedures, and investigators who
purchase adult mice from different sources may not obtain
consistent results due to such differences.
Second, what is the health status of the animals? Strains
of mice developed in facilities that are not specific pathogen
free may contain parasites, bacteria, and viruses that can
alter the health and behavior of the mice. For example, persistent viral infections disrupt learning, memory, and other
behaviors in mice (Brot et al. 1997; Hotchin and Seegal 1977;
Yayou et al. 1993), and these behavioral disruptions may
continue for several months after antiviral therapy (Brot et
al. 1997). Neonatal endotoxin exposure also influences the
development of social behavior (Granger et al. 1996).
Third, what are the physical housing conditions of the
animals? Cage size, number of animals per cage, light:dark
(LiD1) cycle, room temperature, and humidity all influence
behavior. For example, cage size can alter activity levels
(Poon et al. 1997), and social housing conditions may cause
hormonal changes that influence behavior and physiology
(Grimm et al. 1996). Because mice are nocturnal and thus
more active in the dark, attention should be paid to the L:D
'Abbreviations used in this article: DNMTS, delayed nonmatching to
sample; L:D, light:dark.
163
Table 1 Background questions that must be
answered before using a behavioral test battery
1. What is the source of the animals?
2. What is the health status of the animals?
3. What are the physical housing conditions?
• Cage size?
• Number of animals per cage?
• Light:dark cycle?
• Temperature?
• Humidity?
4. What type of food, water, and medications are provided?
• Brand and amount of food?
• Tap, distilled, or specially treated water?
• Medications?
5. What are the litter size and sex composition?
• To cull or not to cull?
6. What are the effects of maternal care?
• Cross-fostering?
• Age at weaning?
7. What are the social experiences from weaning to adulthood?
• Isolation vs. social rearing?
8. Should animals receive environmental enrichment?
cycle under which animals are housed and whether this L:D
cycle is a natural or reversed day:night cycle. Routine cleaning of animal rooms and cage changes can also disrupt
behavior (Milligan et al. 1993; Saibaba et al. 1996).
Fourth, what type of food, water, and medications are
given to the animals? The type and amount of food provided
to the animals may alter their behavior; for example, dietary
restriction alters motor behavior (Duan and Mattson 1999)
and levels of constituents such as lineolate in the diet influence activity, emotionality, and cognitive performance
(Umezawa et al. 1999). Are the animals given tap, distilled,
or specially treated water? Are medications added to the
water or food? The routine administration of medications
such as tetracycline to animals may influence social behavior
in adults (Barnett and Sandford 1982) and behavioral development (Seo et al. 1993).
Fifth, what litter size and sex composition are used? Litter
size and sex composition affect development (Laviola and
Terranova 1998; Tanaka 1998). Should litter size be adjusted
if one strain produces significantly more pups per litter than
another strain? Should the number of pups in litters be
increased or decreased to match for litter size, and what
effects does manipulation of litter size have on behavior?
Sexual differentiation in utero can be modulated by the sex
ratio of the litter, resulting in changes in sexually dimorphic
behaviors (Vandenbergh and Huggett 1995; Vom Saal 1981).
Sixth, what effects does maternal care have on behavior?
Maternal care influences neural and behavioral development
(Caldji et al. 1998; Liu et al. 1997), and there are strain
differences in maternal care patterns (Brown et al. 1999b;
164
Carlieretal. 1982). Should pups be cross-fostered to "good"
mothers if their own mothers fail to care for them? How does
one determine whether strain differences are due to the
effects of maternal environment (pre- and postnatal) or to
genetic differences between strains (Carlier et al. 1983)?
Comparison of the behavior of pups reared by mothers of
their own strain and pups cross-fostered to mothers of other
strains provides a mechanism for dissociating the effects of
postnatal maternal care from the putative effects of genetics
on behavior (van Abeelen 1980), but this paradigm does not
control for prenatal maternal influences on behavior (Carlier
et al. 1999). The age at which animals are weaned can also
influence behavior. Animals weaned 3 days earlier than
normal or weaned at lower than the normal weight are more
likely to develop behavioral stereotypy (Wuerbel and
Stauffacher 1997).
Seventh, what are the social experiences from weaning
to adulthood that can affect behavior? Housing in single-sex
littermate groups, single-sex mixed litter groups, or mixed
sex groups can affect social behavior (Hahn and Schanz
1996; Harshfield and Grim 1997; Koolhaas et al. 1997;
Young et al. 1997). Comparison of inbred strains of mice
reared under different social conditions is an important procedure for dissociating genetic and social experiences in
determining behavior (Hoffmann et al. 1993). Social isolation induces physiologic and behavioral abnormalities in
mice (Brain 1975; Hall 1998; Morse et al. 1993), whereas
housing mice in all male groups leads to the formation of
dominance hierarchies. Male mice housed in groups may
develop dominance hierarchies, resulting in increased corticosteroid and opioid release and decreased LH and folliclestimulating hormone levels in subordinate males (Bronson
and Eleftheriou 1965; Bronson et al. 1973; Miczek et al.
1982). Dominant and subordinate animals differ in locomotor
activity, swimming, and anxiety measures (Ferrari et al.
1998; Hilakivi-Clarke and Lister 1992), and these differences may have significant effects on performance in tests of
learning and memory (Dubrovina et al. 1997). When animals are group housed, individuals are often identified using
ear tags, ear punching, or toe clipping. If a stressful procedure such as toe clipping is to be used, the effect of this
procedure on future performance should be considered
(Schellinck and Anand 1999).
Eighth, should animals receive environmental enrichment? Environmental enrichment facilitates neural and cognitive development (Boehm et al. 1996; Soffie et al. 1999)
and reduces anxiety (Chapillon et al. 1999). The effects of
environmental enrichment may be age and strain specific
(Soffie et al. 1999; van de Weerd et al. 1994). Environmental enrichment may enable the same strains of mice to
perform better in tests of learning and memory, have greater
dendritic arborization, and have longer axons (Greenough
and Volkmar 1973). Current North American animal care
standards advocate the use of shelters, chewing blocks, and
nesting material to reduce animal stress (NRC 1996;
Wolfensohn and Lloyd 1998); and although the use of nesting material as environmental enrichment does not appear to
ILAR Journal
alter behavior in mice (van de Weerd et al. 1997), the use of
such materials for environmental enrichment should be
noted.
Procedural Questions for Designing
Behavioral Studies
In addition to the background questions listed in Table 1, a
number of procedural questions must be answered when designing behavioral experiments with transgenic and mutant
mice (Table 2). First, when a genetically altered mouse is
tested, what control strains should be used? Genetically
engineered mice may differ from their background strains in
physiologic, endocrine, and neural functions and in the development of these functions (Gassmann and Hennett 1998;
Nelson 1997; Shastry 1995). Of the variety of different background strains used in the development of knockout and
mutant mice, some (e.g., the 129Sv strain) have substrains
that show behavioral differences (Simpson et al. 1997;
Balogh et al. 1999). For example, the 129/SvJ strain has agerelated visual pathology (Hengemihle et al. 1999).
Every attempt should be made to use the appropriate
control strain whenever the animals are bred as homozygous
knockouts. If a single control strain is not available, then all
parent strains must be used because there are large differences in behavior, physiology, and anatomy between inbred
strains. When heterozygotic mice are bred to achieve knockout (-/-), homozygous (+/+), and heterozygous (+/-) strains,
littermates are appropriate controls provided they are from
an appropriate hybrid cross (Banbury Conference 1997;
Frankel 1998; Lathe 1996). The use of littermate controls is
often recommended (Banbury Conference 1997; Frankel
1998; Zorrilla 1997). Some information on appropriate control strains is given by Fox and Witham (1997) and on the
Jackson Laboratory website (WWW.JAX.ORG).
Second, at what age should mice be tested: during
infancy, adolescence, adulthood or old age? Mice are weaned
at about 21 days of age and can be mated as early as 28 days
of age, but there are strain and sex differences in reproductive performance and longevity (Fox and Witham 1997).
Table 2 Procedural questions for designing
behavioral studies
1.
2.
3.
4.
5.
6.
7.
8.
9.
What control strains should be used?
At what age should mice be tested?
Should both males and females be tested?
What time of the day:night cycle should mice be tested?
How many tests should be given to each animal, and
how many subjects per group should be tested?
How many different test paradigms should be used?
How should mice be handled before and during testing?
What apparatus should be used for each test?
What is the testing room environment?
Volume 4 1 , Number 3
2000
Relatively small differences in the age at testing can result in
different behavioral results (Hascoet et al. 1999; Strohle et
al. 1998) because neural development and sexual differentiation of the brain continues until puberty (Andersen et al.
1997). With knockout mice, the longevity of the strain is
greatly affected by the targeted gene. The lack of some
genes can be lethal at birth, cause animals to have a shortened life span, decrease reproductive viability, or increase
the time course of development (Miyamoto 1997; Nishii et
al. 1998; Yagi et al. 1998). In determining the age at which
the animals should be tested, one must consider the rationale
behind the test. Animals may perform differently at different
ages, so it may also be necessary to test both young and old
animals (Miyamoto et al. 1986).
Third, should both sexes be tested? Sex differences
should be considered as males and females may be differentially affected by gene mutations. For example, male and
female ddY mice differ in the performance of spatial learning tasks but not in avoidance learning (Mishima et al. 1986),
and delta-168 transgenic mice show sex-linked changes in
locomotor activity and learning (Douhet et al. 1997).
Genotype-by-sex interactions also have been found in tests
of anxiety in 5-HT1A receptor knockout mice (Ramboz et al.
1998).
Fourth, what time of the day:night cycle should mice be
tested? Behavior may be influenced by the time of testing
during the L:D cycle (Miyamoto et al. 1986; Rosenwasser et
al. 1979; Sandman et al. 1971; Stephens et al. 1967). For
example, Kriegsfeld et al. (1999) found that nitric oxide synthase knockout mice had deficits in motor coordination when
tested in the dark phase of the L:D cycle but not when tested
in the light phase. They state, "Even though rodents are
nocturnal animals, most behavioral studies, including learning, are conducted during the light period. Our findings
emphasize the importance of examining the diurnal variations, especially in gene knockout research" (p. 314).
Fifth, how many tests should be given to each animal and
how many subjects per group should be tested? Because
transgenic and knockout mice are expensive and difficult to
maintain, only a few may be available. Before testing each
mouse multiple times, one should consider whether repeated
testing affects the animal and whether there is a test order
interaction. How many mice should there be in each group
so that adequate statistical power is reached? As more
advanced statistical and multivariate analyses are being used
in animal behavior studies, one must provide sample sizes
adequate to attain the power level determined for the effect
sizes observed (Belknap 1998). In general, mice should be
tested with groups in excess of 10, but this recommendation
is dependent on the effect size estimated and the desired
power value (Keppel 1991). The problem of experimental
design, sample size, and power in genetic experiments is
discussed by Wahlsten (1999).
Sixth, how many different test paradigms should be used?
The number of tests to use with each strain and which tests to
use have been discussed by Crawley and Pay lor (1997) and
is discussed below (see Tests for Sensory and Motor Deficits
165
and Species-typical Behavior: The Mouse Ethogram).
Because many behavioral tests are designed for rats, one
must consider species differences in size and in behavior
when adapting these tests for use with mice. The size and
sensitivity of the apparatus and even the behavioral measures
scored may have to be altered to adapt behavioral tests
designed for rats for use with mice (Crawley and Paylor
1997).
Seventh, how should mice be handled during testing?
How the animals are handled before and during testing is
important. Test behavior may be influenced by habituation
to handling and the method of retrieving the subject (Barry
1957). For example, whether mice are picked up by the tail,
carried in a container, or allowed to move themselves into
transport container may influence their behavior.
Eighth, what test apparatus should be used? Its size,
material, and method of cleaning may all influence the test
results. Using a rat apparatus to test mice is often not appropriate inasmuch as the activity levels of mice make it difficult for them to perform tasks that are easy for rats (Crawley
and Paylor 1997). Additionally, cleaning the apparatus in a
manner that will eliminate odor trails is especially important
because odor may provide cues that alter the behavior of the
subjects (Roullet et al. 1993).
Ninth, what is the testing room environment? Animals
tested in an open laboratory are subject to distractions from
visual, auditory, and olfactory stimuli. Extraneous stimuli
such as odors from perfume or chemicals and noise from
equipment, both audible and ultrasonic, may interfere with
the behavior of the animal (Sales et al. 1988). Testing in a
purpose-built behavioral test room shields animals from these
distracting stimuli; however, even under controlled conditions, there are effects of the laboratory environment on
behavior (Crabbe et al. 1999).
Tests for Sensory and Motor Deficits
Many strains of congenic, inbred, and mutant mice suffer
from sensory and/or motor deficits that alter their behavior.
Mice may suffer from retinal degeneration and perform
poorly in tests requiring visual acuity (Hengemihle et al.
1999; Provencio et al. 1994). Some strains of mice such as
DBA/2 and C57BL/6J (B6) become deaf at an early age and
therefore perform poorly in tests requiring auditory discrimination (Erway et al. 1993; Steel 1995). Other strains may
have disrupted olfactory sensitivity (Feron and Baudoin
1998), somatosensory deficits (Wilkinson et al. 1996), or
abnormal pain thresholds (Mogil and Belknap 1997; Onaivi
et al. 1995). In each of these cases, poor performance in
behavioral tests may be due to sensory rather than cognitive
deficits. An example of a test battery for measuring sensory
and motor abilities is given in Table 3.
Strains of mice such as staggerer, lurcher, weaver, dystonia
musculorium mutants, and dopamine 1A receptor-deficient
mice suffer from locomotor abnormalities that may affect
movement and grooming (Bolivar et al. 1996a,b; Coscia and
166
Table 3 Example of a sensory-motor test battery
1. Visual cliff (Zhong et al. 1996)
2. Auditory startle response (Golub et al. 1994; Koch and
Schnitzler1997)
3. Odor orientation/discrimination test (Sundbergetal. 1982)
4. Locomotion in the open field test (Markowska et al.
1998; Moechars et al. 1998)
5. Motor learning and balance in the roto-rod test (Lalonde
et al. 1996; Le Marec and Lalonde 1997)
6. Pain sensitivity in the hot plate or tail-flick test
(Rubinstein et al. 1996)
Fentress 1993; Cromwell et al. 1998; Lalonde et al. 1994,
1996; Strazielle and Lalonde 1998). These motor deficits
may affect performance in behavioral tasks, particularly
when time is measured as the dependent variable. For these
reasons, tests of sensory and motor performance should be
conducted on transgenic and knockout mice to ensure that
abnormal responses, particularly in learning and memory
tasks, are not due to noncognitive sensory-motor abnormalities.
Species-typical Behavior: The Mouse
Ethogram
Genetic manipulation may alter species-typical behaviors in
mice, and test batteries should be able to detect these changes.
Ethologic descriptions of species-typical behavior use an
"ethogram," which is a complete inventory of the behavior
repertoire of a species (Tinbergen 1951, p. 7) in terms of
qualitatively and quantitatively measured "behavioural
units." Each behavioral study thus begins with a description
of the units of behavior to be recorded, which are often
grouped into functional categories such as locomotor, reproductive, and aggressive (Grant and Mackintosh 1963).
Eisenberg (1968) has provided an ethogram for mouse
behavior that describes the behavior observed in adult mice
in isolation and in social encounters (Table 4). Ethologic
analyses have been used to study locomotor behavior (Carter
et al. 1999; Crabbe 1986; Paulus et al. 1999), grooming
(Drago et al. 1999), exploration (Elias et al. 1975), fear
(Brown et al. 1999c; Trullas and Skolnick 1993), anxiety and
defensive behavior (Griebel et al. 1996), sexual behavior
(McGill and Tucker 1964), aggression (Schneider et al.
1992), and parental behavior (Ward 1980) in mice.
Strain differences in fearfulness and anxiety should be
studied because emotional responses interfere with the
expression of other behaviors and can affect performance in
learning and memory tasks (Markowska et al. 1998; Parks et
al. 1998; Smith etal. 1998; Strohle et al. 1998). For example,
serotonin 5-HT1A receptor knockout mice show increased
anxiety (Ramboz et al. 1998), and dopamine D3 receptor
knockout mice show reduced anxiety levels (Steiner et al.
IIAR Journal
Table 4 Behavioral categories in the mouse
ethogram*
Individual behavior patterns
1. Sleep and resting
2. Locomotion
3. Grooming: care of the body surface and comfort
movements
4. Ingestion of food
5. Gathering food and caching
6. Digging
7. Nest building
8. Exploration and foraging
9. Fear, anxiety, and defensive behavior
Social behavior patterns
1. Initial contact-social investigation
2. Contact promoting and sexual behavior
3. Approach, flight, attack, and other agonistic patterns
4. Miscellaneous behaviors seen in a social context
5. Parental behavior
a
Adapted from Eisenberg JF. 1968. Behavior patterns. In: King JA,
editor. Biology of Peromyscus (Rodentia). Washington DC: American Society of Mammologists. p 451-494.
1997). For these reasons, tests of fearfulness or anxiety
should be conducted on transgenic and knockout mice to
ensure that abnormal behavioral scores are not due to noncognitive anxiety factors. The emotional/defensive behavior
test battery (Table 5) is an example of an ethologic test
battery.
Learning and Memory Test Battery
Many tests have been devised to test learning and memory in
mice (Crawley and Paylor 1997; Wehner et al. 1996). For
example, the Hebb-Williams maze has been used as a general test of animal intelligence (Brown and Stanford 1997;
Meunier et al. 1986), and the habituation test has been used
as a nonassociative test of learning (Platel and Porsolt 1982).
Table 5 Example of a species-typical emotional/
defensive behavior test battery
1.
2.
3.
4.
5.
6.
7.
8.
Locomotion/exploration in the open field (Crawley 1985)
Elevated plus maze (Lister 1987)
Elevated zero maze (Brown et al. 1999a)
Light:dark box (Crawley 1985)
Holeboard test (File and Wardill 1975)
Social interaction test (Olivier and van Dalen 1982)
Social conflict test (Vogel et al. 1971)
Exploratory/defensive behavior in the visible burrow
system (Blanchard et al. 1990)
Volume 4 1 , Number 3
2000
One advantage of the Hebb-Williams test is that it has a
series of graded problems so that the same animals can be
compared on easy and difficult problems in the same apparatus (Stanford et al. 1998, 1999).
Although the neurobiologic basis for learning and
memory is far from understood, at least five different neural
systems have been proposed to underlie these different
memory systems. Damage to one of these structures disrupts
some memory systems but leaves others intact (Squire 1987).
The five neural systems most widely studied include the hippocampus, amygdala, dorsal striatum, rhinal cortex, and cerebellum. None of these neural systems is a unitary entity.
The hippocampus is a complex structure involving many
components (Shepherd 1994). The amygdala is composed
of at least four independent units (Swanson and Petrovich
1998), and the dorsal striatum is part of the basal ganglia
complex (White 1997). The rhinal cortex is part of the
temporal lobe (Aggleton and Saunders 1997), and the cerebellum has a number of components and connections to other
brain areas (Lavond et al. 1993).
Many behavioral paradigms have been used to examine
the contribution of each of the five neural systems implicated
in learning and memory function in rodents. Although primarily used with rats, these paradigms are being adapted for
use with transgenic and knockout mice (Bach et al. 1995;
Silva et al. 1992). One example of a learning and memory
test battery that is based on these different memory systems
is presented in Table 6. We are in the process of refining this
test battery (Brown et al. 1998b).
Hippocampal functioning has typically been examined
with tests of spatial learning such as the Morris water maze
(Upchurch and Wehner 1988) and the radial arm maze
(McDonald and White 1993). Learning the water maze, however, may depend on the amygdala as well as the hippocampus inasmuch as swimming in an opaque pool may
arouse anxiety (McNaughton 1991). The win-shift paradigm
in the radial arm maze can be used as a paradigm to test
hippocampally based spatial learning in mice (McDonald
and White 1993).
Table 6 Example of a learning and memory test
battery based on multiple memory systems
1.
2.
3.
4.
5.
6.
7.
8.
9.
Habituation (Platel and Porsolt 1982)
Hebb-Williams maze (Meunier et al. 1986)
Morris water maze (Upchurch and Wehner 1988)
Win-shift task on the 8-arm radial maze (McDonald and
White 1993)
Conditioned fear response (Fanselow 1994)
Step down passive avoidance (Mondadori and
Weiskrantz 1993)
Cued discrimination task of the 8-arm radial maze
(Ammassari-Teule et al. 1993; McDonald and White
1993)
Object exploration and memory (Messier 1997)
Delayed nonmatching to sample (Zhu et al. 1995)
167
The role of the amygdala in learning and memory is
generally studied using tests of conditioned fear responses.
In the context of conditioning, animals shocked in a distinct
environment display conditioned fear behaviors (e.g., freezing) when they remember the association between the context
and the shock (Fanselow 1990, 1994). In place avoidance
conditioning, animals avoid the place associated with the
shock (Siegfried and Frischknecht 1989). In active avoidance conditioning, mice learn to avoid the location/place
associated with fear; and in passive avoidance conditioning,
the mouse must learn to stay in the nonshock compartment
and avoid entering the shock compartment (Delprato and
Rusiniak 1991; Riekkinen et al. 1993; Schulteis and Koob
1993).
The dorsal striatum (caudate/putamen) memory system
is activated in operant conditioning tasks in which a stimulus
associated with a specific motor response is reinforced. The
cued win-stay radial maze task has been effective in dissociating deficiencies in the caudate/putamen from those in the
amygdala and hippocampus. In this task, a light is used as
the cue; and during each trial, the animal learns to enter the
arm of the maze associated with the light/tone (Packard et al.
1989).
The rhinal cortex appears to be involved in object recognition (Zhu et al. 1995). In rats, recognition memory in the
delayed nonmatching to sample (DNMTS1) object recognition task is impaired by lesions of the rhinal cortex but not by
lesions of the amygdala or hippocampus (Mumby and Pinel
1994; Mumby et al. 1992). This impairment difference suggests that the DNMTS object recognition task can be used to
dissociate the role of the rhinal cortex, hippocampus, and
amygdala in learning. When testing object recognition, it is
necessary to ensure that there is an absence of all spatial and
odor cues and that the objects have no biologic significance
so that the spatial and emotional learning and memory systems are not activated. Mumby and Pinel (1994) have used
an object recognition paradigm, and Winters et al. (2000)
have developed an olfactory DNMTS task.
The cerebellum controls coordinated motor learning tasks
such as the conditioned eye-blink (Thompson 1986) and the
vestibuloocular reflex (Ito 1984). Chen et al. (1996) have
studied the conditioned eye-blink response in mutant mice
with Purkinje cell degeneration.
Numerous test batteries exist for analyzing the development of behavioral reflexes (Fox 1965; Kodama 1993;
Rogers et al. 1997), sensory-motor development (Heuze et
al. 1997; Noel 1989; Roubertoux et al. 1985, 1992), locomotor behavior (Brown et al. 1999a), grooming (Bolivar et
al. 1996a), swimming (Bolivar et al. 1996b), exploration
(Tremml et al. 1998), anxiety and fear (Ramboz et al. 1998),
ultrasonic vocalizations (Bolivar and Brown 1994; Hahn et
al. 1987), and social behavior (Williams and Scott 1954).
Test batteries used for studying behavioral development
include the Collaborative Behavioral Teratology Study
Battery, the Cincinnati Psychoteratogenicity Screening Test
Battery, and the Functional Observational Battery. The Collaborative Behavioral Teratology Study Battery (Vorhees
1983,1985) was created to assess the reliability and sensitivity of several behavioral tests when conducted under identical
protocols in different laboratories or in the same laboratory
over time. This battery employs tests of developmental landmarks (e.g., eye opening, incisor eruption, vaginal opening,
testes descent) and six behavioral tests (negative geotaxis,
olfactory discrimination, auditory startle habituation, figure8 maze activity, visual discrimination learning, and activity).
The Cincinnati Psychoteratogenicity Screening Test Battery
(Vorhees 1985) was designed to be a comprehensive, reliable, and practical screening battery. It includes tests of
physical landmarks (e.g., upper and lower incisor eruption,
eye opening, vaginal patency), surface righting, negative geotaxis, swimming ontogeny, pivoting locomotion, olfactory
orientation, auditory startle, activity, spontaneous alternation, maze learning, and active and passive avoidance. The
Functional Observational Battery (Moser 1989,1990; Tilson
and Moser 1992) uses 31 behavioral tests to assess behavioral development. Although these test batteries were developed for toxicologic studies they can be adapted to study
behavioral development in transgenic mice.
Both physical and behavioral development should be
assessed in a developmental test battery. In Table 7, an
example of a developmental test battery is shown, which
includes tests that examine the sensory and motor development of mice and their locomotion, exploration, and fear
responses (Bolivar and Brown 1994; Brown et al. 1998a,
1999a). These tests begin at birth, and animals are tested
daily until they reach each developmental milestone, as
described by Fox (1965) and Bolivar and Brown (1994).
Behavioral Development
The genetic basis of neural and behavioral development is
important for understanding the interaction between genes
and environment in determining behavior (Bateson 1987,
1998; Gottlieb 1998). Likewise, the study of behavioral
development is an important tool for genetic analysis in mice
(Roubertoux et al. 1992). Analysis of behavioral development in mutant mice is also important for understanding the
ontogeny of neurodevelopmental disorders (Bolivar and
Brown 1994; Coscia and Fentress 1993; Greenough 1986;
Lyons and Wahlsten 1988; van Ableen and Kalkhoven 1970).
168
Aging Mice
Test batteries are important for measuring the behavioral
changes associated with aging (Jucker and Ingram 1997).
Sensory-motor, species-typical, and learning and memory
test batteries can be used with aging animals, or composite
test batteries can be developed that use subsets of the other
test batteries (Giuliani et al. 1994; Markowska et al. 1998;
Miyamoto 1997). Genetically defined mouse models are
important for the study of aging (Sprott 1997), and an
ILAR Journal
Table 7 Example of a developmental test battery
Physical development (Fox 1965)
1. Growth rate (body weight)
2. Age at eye opening
3. Age at pinna unfolding
4. Age at incisor eruption
5. Age at vaginal opening
Behavioral development
1. Reflexes: righting, grasp, locomotion, rooting, and
orienting (Fox 1965)
2. Ultrasonic vocalizations (Bolivar and Brown 1994)
3. Acoustic startle test (Pryor et al. 1983)
4. Negative geotaxis (Pryor et al. 1983)
5. Visual cliff (Tees 1974)
6. Olfactory orientation (Brown and Willner 1983)
7. Forelimb grip strength (Fox 1965; Pryor et al. 1983)
8. Locomotion and exploratory behavior in an open field
(Brown etal. 1999a)
9. Motor coordination
• Locomotion in the open field test (Brown et al. 1999a)
• Swimming (Bolivar et al. 1996b)
• Self-grooming (Coscia and Fentress 1993)
10. Exploration in the hole-board (File and Wardill 1975)
11. Elevated plus maze (Lister 1987)
example of a test battery for examining behavioral changes
during aging in mice is shown in Table 8.
Summary and Conclusions
In this article, we have examined some of the variables that
must be considered in the development of behavioral test
batteries and in the maintenance of animals being tested.
Sample test batteries for sensory-motor abilities, speciestypical behaviors, learning and memory, behavioral development, and aging are presented. One of the problems,
however, is that the development of such test batteries is
idiosyncratic, and no standardized test batteries exist for
mice. Future research projects should focus on the procedures for developing such a set of standardized test batteries,
the psychonomic properties of these test batteries, and their
practical application. The development of the SHIRPA test
battery is a step in this direction (Rogers et al. 1999)
(SHIRPA derives from SmithKline Beecham Pharmaceuticals of the Harwell MRC Mouse Genome Centre and Mammalian Genetics Unit of Imperial College School of Medicine
at St. Mary's Royal London Hospital, St. Bartholomew's,
and the Royal London School of Medicine). Tests batteries
developed for neurotoxicology (Tilson 1987) might be
adapted for use with transgenic mice, but there should be
some theoretical basis for the inclusion of items in a test
battery. We are developing test batteries for learning in
memory in mice based on the theory of multiple memory
Volume 4 1 , Number 3
2000
Table 8 Example of a test battery for the study of
aginga
1. Sensory tasks
• Visual cliff
• Auditory startle
• Olfactory discrimination
2. Motor tasks
• Roto rod
• Climbing
• Activity in a running wheel
3. Exploration in an open field
4. Emotionality in an elevated plus maze
5. Spatial memory
• Morris water maze
• Radial arm maze
6. Fear conditioning
7. Object recognition memory
8. Social memory
a
Adapted from Giuliani A, Ghirardi O, Caprioli A, di Serio S,
Ramacci MT, Angelucci L. 1994. Multivariate analysis of behavioral
aging highlights some unexpected features of complex systems
organization. Behav Neural Biol 61:110-122; Markowska Al,
Spangler EL, Ingram DK. 1998. Behavioral assessment of the
senescence-accelerated mouse (SAM P8 and R1). Physiol Behav
37:909-913; Miyamoto M. 1997. Characteristics of age-related
behavioral changes in senescence-accelerated mouse SAMP8 and
SAMP10. Exp Gerontol 32:139-148.
systems (Brown et al. 1998a,b) and a developmental test
battery (Brown et al. 1998a, 1999a) based on psychobiologic
and neurodevelopmental theories of development (Cicchetti
and Cannon 1999).
Acknowledgments
We thank Melanie McFadyen for her help in preparing this
paper. R.E.B. and H.M.S. were supported by research grants
from the Natural Science and Engineering Research Council
of Canada, and L.S. was supported by the K.M. Hunter
Medical Research Council of Canada Doctoral Research
Award.
References
Aggleton JP, Saunders RC. 1997. The relationships between temporal lobe
and diencephalic structures implicated in anterograde amnesia. Memory
5:49-71.
Ammassari-Teule M, Hoffmann HJ, Rossi-Arnaud C. 1993. Learning in
inbred mice: Strain-specific abilities across three radial maze problems.
Behav Genet 23:405-412.
Andersen SL, Rutstein M, Benxo JM, Hostetter JC, Teicher MH. 1997. Sex
differences in dopamine receptor overproduction and elimination.
Neuroreport 8:1495-1498.
Bach ME, Hawkins RD, Osman M, Kandel ER, Mayford M. 1995. Impairment of spatial but not contextual memory in CaMKII mutant mice with
169
a selective loss of hippocampal LTP in the range of the theta frequency.
Cell 81:905-915.
Balogh SA, McDowell CS, Stavnezer AM, Denenberg VH. 1999. A behavioral assessment of an inbred substrain of 129 mice with behavioral
comparisons to C57BL/6J mice. Brain Res 836:38-48.
Banbury Conference [Banbury Conference on Genetic Background in Mice].
1997. Mutant mice and neuroscience: Recommendations concerning
genetic background. Neuron 19:755-759.
Barnett SA, Sandford MH. 1982. Decrement in "social stress" among wild
Rattus rattus treated with antibiotic. Physiol Behav 28:483-487.
Barry in H. 1957. Habituation to handling as a factor in retention of maze
performance in rats. J Comp Physiol Psych 50:366-367.
Bates GP, Mangiarini L, Mahal A, Davies SW. 1997. Transgenic mouse
models of Huntington's disease. Hum Mol Genet 6:1633-1637.
Bateson P. 1987. Biological approaches to the study of behavioral development. Int J Behav Devel 10:1-22.
Bateson P. 1998. Genes, environment and the development of behavior.
Novartis Found Symp 213:160-170.
Belknap JK. 1998. Effect of within-strain sample size on QTL detections
and mapping using recombinant inbred mouse strains. Behav Genet
28:29-38.
Blanchard DC, Blanchard RJ, Tom P, Rodgers RJ. 1990. Diazepam changes
risk assessment in anxiety/defensive test battery. Psychopharm 101:511518.
Boehm GW, Sherman GF, Hoplight BJ II, Hyde LA, Waters NS, Bradway
DM, Galaburda AM, Denenberg VH. 1996. Learning and memory in
autoimmune BXSB mouse: Effects of neocortical ectopias and environmental enrichment. Brain Res 726:11-22.
Bolivar VJ, Brown RE. 1994. The ontogeny of ultrasonic vocalizations and
other behaviours in male Jimpy (jp/Y) mice and their normal littermates.
Dev Psychobiol 27:101-110.
Bolivar VJ, Danilchuck W, Fentress JC. 1996a. Separation of activation and
pattern in grooming development of weaver mice. Behav Brain Res
75:49-58.
Bolivar VJ, Manley K, Fentress JC. 1996b. The development of swimming
behavior in the neurological mutant weaver mouse. Dev Psychobiol
29:123-137.
Brain P. 1975. What does individual housing mean to a mouse? Life Sci
16:187-200.
Bronson FH, Eleftheriou BE. 1965. Adrenal response to fighting in mice:
Separation of physical and psychological causes. Science 147:627-628.
Bronson FH, Stetson MH, Stiff ME. 1973. Serum FSH and LH in male mice
following aggressive and nonaggressive interaction. Physiol Behav
10:369-372.
Brot MD, Rail GF, Oldstone MB, Koob GF, Gold LH. 1997. Deficits in
discriminated learning remain despite clearance of long-term persistent
viral infection in mice. J Neurovirol 3:265-273.
Brown RE, Dastur F, Stanford L, Schellinck HM. 1998a. Mouse IQ: Development of standardized test batteries for knockout and mutant mice
(Abstract 13). Dev Psychobiol 33:368.
Brown RE, Stanford L, Schellinck HM. 1998b. Mouse IQ: A systematic set
of learning and memory tests for knockout and mutant mice (Abstract
P3-3). Brain Res 809:A24.
Brown RE, Corey SC, Moore AK. 1999a. Differences in measures of exploration and fear in MHC-Congenic C57BL/6J and B6-H-2K mice. Behav
Genet 29:263-271.
Brown RE, Mathieson WB, Stapleton J, Neumann PE. 1999b. Maternal
behavior in female C57BL/6J and DBA/2J inbred mice. Physiol Behav
67:599-605.
Brown RE, Schellinck HM, Jagosh J. 1999c. Behavioural sudies of MHCcongenic mice. Genetica 104:249-257.
Brown RE, and Stanford L. 1997. The Hebb-Williams Maze: 50 years of
research (1946-1996). (Abstract 110.14). Society for Neuroscience
Abstracts 23:278.
Brown RE, Willner JA. 1983. Establishing an "affective scale" for odor
preferences of infant rats. Behav Neural Biol 39:251-260.
Brunner D, Buhot MC, Hen R, Hofer M. 1999. Anxiety, motor activation,
170
and maternal-infant interactions in 5-HT1B knockout mice. Behav
Neurosci 113:587-601.
Caldji C, Tannenbaum B, Sharma S, Francis D, Plotsky PM, Meaney MJ.
1998. Maternal care during infancy regulates the development of neural
systems mediating the expression of fearfulness in the rat. Proc Natl
Acad Sci U S A 95:5335-53340.
Carlier M, Roubertoux P, Cohen-Salmon C. 1983. Early development in
mice: I. Genotype and post-natal maternal effects. Physiol Behav 30:837844.
Carlier M, Roubertoux P, Cohen-Salmon C. 1982. Differences in patterns of
pup care in Mus musculus domesticus 1-Comparison between eleven
inbred strains. Behav Neural Biol 35:205-210.
Carlier, M, Roubertoux P, Wahlsten, D. 1999. Maternal effects in behavior
genetic analysis. In: Jones B, Mormede P. editors. Neurobehavioral
Genetics Methods and Applications. NewYork: CRC Press, p 187-197.
Carter RJ, Lione LA, Humby T, Mangiarini L, Mahal A, Bates GP, Dunnett
SB, Morton AJ. 1999. Characterization of progressive motor deficits in
mice transgenic for the human Huntington' s disease mutation. J Neurosci
19:3248-3257.
Chapillon P, Manneche C, Belzung C, Caston J. 1999. Rearing environmental enrichment in two inbred strains of mice: 1. Effects on emotional
reactivity. Behav Genet 29: 41-46.
Chen C, Tonegawa S. 1997. Molecular genetic analysis of synaptic plasticity, activity-dependent neural development, learning, and memory in
the mammalian brain. Annu Rev Neurosci 20:157-184.
Chen L, Bao S, Lockard JM, Kim JK, Thompson RF. 1996. Impaired classical eyeblink conditioning in cerebellar-lesioned and Purkinje cell
degeneration (ped) mutant mice. J Neurosci 16:2829-2838.
Cicchetti D, Cannon TD 1999. Neurodevelopmental processes in the ontogenesis and epigenesis of psychopathology. Dev Psychopath 11:375393.
Coscia EM, Fentress JC. 1993. Neurological dysfunction expressed in the
grooming behavior of developing weaver mutant mice. Behav Genet
23:533-541.
Crabbe JC. 1986. Genetic differences in locomotor activation in mice.
Pharmacol Biochem Behav 25:289-292.
Crabbe JC, Walhsten D, Dudek BC. 1999. Genetics of mouse behavior:
Interactions with laboratory environment. Science 284:1670-1672.
Crawley JN. 1985. Exploratory behavior models of anxiety in mice.
Neurosci Biobehav Rev 9:37-44.
Crawley JN. 1999. Behavioral phenotyping of transgenic and knockout
mice: Experimental design and evaluation of general health, sensory
functions, motor abilities, and specific behavioral tests. Brain Res
835:18-26.
Crawley JN, Paylor R. 1997. A proposed test battery and constellations of
specific behavioral paradigms to investigate the behavioral phenotypes
of transgenic and knockout mice. Horm Behav 31:197-211.
Cromwell HC, Berridge KC, Drago J, Levine MS. 1998. Action sequencing
is impaired in DIA-deficient mutant mice. Eur J Neurosci 10:24262432.
Delprato DJ, Rusiniak KW. 1991. Response patterns in shock avoidance
and illness aversion. In: Denny MR, editor. Fear, Avoidance, and Phobias: A Fundamental Analysis. Hillsdale NJ: Lawrence Erlbaum Associates, Inc. p 455.
Douhet P, Bertaina V, Durkin T, Calas A, Destrade C. 1997. Sex-linked
behavioural differences in mice expressing a human insulin transgene in
the medial habenula. Behav Brain Res 89:259-266.
Drago F, Contarino A, Busa L. 1999. The expression of neuropeptideinduced excessive grooming behavior in dopamine Dl and D2 receptordeficient mice. Eur J Pharmacol 365:125-131.
Duan W, Mattson MP. 1999. Dietary restriction and 2-deoxyglucose administration improve behavioral outcome and reduce degeneration of
dopaminergic neurons in models of Parkinson's disease. J Neurosci Res
57:195-206.
Dubrovina NI, Popova EV, Il'iuchenok RIu. 1997. [The retrieval of a
memory trace in mice with aggressive and submissive types of behavior].
Zhurnal Vysshei Nervnoi Deiatelnosti Imeni I. P. Pavlova 47:762-765.
ILAR Journal
Eisenberg JF. 1968. Behavior patterns. In: King JA, editor. Biology of
Peromyscus (Rodentia). Washington DC: American Society of
Mammologists. p 451-495.
Elias PK, Elias MF, Eleftheriou BE. 1975. Emotionality, exploratory behavior, and locomotion in aging inbred strains of mice. Gerontologia 21:4655.
Erway LC, Willott JF, Archer JR, Harrison DE. 1993. Genetics of agerelated hearing loss in mice: I. Inbred and Fl hybrid strains. Hear Res
65:125-132.
Fanselow MS. 1990. Factors governing one-trial contextual conditioning.
Anim Learn Behav 18:264-270.
Fanselow MS. 1994. Neural organization of the defensive behavior system
responsible for fear. Psych Bull Rev 1:429-438.
Feron C, Baudoin C. 1998. Social isolation induces preference for odours of
oestrous females in sexually naive male staggerer mutant mice. Chem
Senses 23:119-121.
Ferrari PF, Palanza P, Parmigiani S, Rodgers RJ. 1998. Interindividual variability in Swiss male mice: Relationship between social factors, aggression, and anxiety. Physiol Behav 63:821-827.
File SE, Wardill AG. 1975. The reliability of the hole-board apparatus.
Psychopharmacologia 44:47-51.
Rood JF, Morley JE. 1998. Learning and memory in the SAMP8 mouse.
Neurosci Biobehav Rev 22:1-20.
Fox WM. 1965. Reflex-ontogeny and behavioural development of the
mouse. Anim Behav 13:234-241.
Fox RR, Witham BA, editors. 1997. Handbook on Genetically Standardized
JAX Mice. 5th ed. Bar Harbour ME: The Jackson Laboratory.
Frankel WN. 1998. Mouse strain backgrounds: more than black and white.
(Letter). Neuron 20:183.
Galani R, Jarrard LE, Will BE, Kelche C. 1997. Effects of postoperative
housing conditions on functional recovery in rats with lesions of the
hippocampus, subiculum, or entorhinal cortex. Neurobiol Learn Mem
67:43-56.
Gassmann M, Hennett T. 1998. From genetically altered mice to integrative
physiology. News Physiol Sci 13:53-57.
Giuliani A, Ghirardi O, Caprioli A, di Serio S, Ramacci MT, Angelucci L.
1994. Multivariate analysis of behavioral aging highlights some unexpected features of complex systems organization. Behav Neural Biol
61:110-122.
Golub MS, Han B, Keen CL, Gershwin ME. 1994. Auditory startle in Swiss
Webster mice fed excess aluminum in diet. Neurotoxicol Teratol 16:423425.
Gottlieb G. 1998. Naturally occurring environmental and behavioral influences on gene activity: From central dogma to probabilistic epigenesis.
Psychol Rev 105:792-802.
Granger DA, Hood KE, Dceda SC, Reed CL, Block ML. 1996. Neonatal
endotoxin exposure alters the development of social behavior and the
hypothalamic-pituitary-adrenal axis in selectively bred mice. Brain
Behav Immun 10:249-259.
Grant EC, Mackintosh JH. 1963. A comparison of the social postures of
some common laboratory rodents. Behavior 21: 246-259.
Greenough WT. 1986. What's special about development? Thoughts on the
base of experience-sensitive synaptic plasticity. In: Greenough WT,
Juraska JM, editors. Developmental Neuropsychobiology. Toronto
Ontario Canada: Academic Press, p 387-407.
Greenough WT, Volkmar FRR. 1973. Pattern of dendritic branching in
occipital cortex of rats reared in complex environments. Exp Neurol
40:491-504.
Griebel G, Blanchard DC, Blanchard RJ. 1996. Evidence that the behaviors
in the mouse defense test battery relate to different emotional states: A
factor analytic study. Physiol Behav 60:1255-1260.
Grimm MS, Emerman JT, Weinberg J. 1996. Effects of social housing
conditions and behavior on growth of the Shionogi mouse mammary
carcinoma. Physiol Behav 59:633-642.
Hahn ME, Hewitt JK, Adams M, Tully T. 1987. Genetic influences on
ultrasonic vocalizations in young mice. Behav Genet 17:155-166.
Hahn ME, Schanz N. 1996. Issues in the genetics of social behavior: Revisited. Behav Genet 26:463-470.
Volume 4 1 , Number 3
2000
Hall FS. 1998. Social deprivation of neonatal, adolescent and adult rats has
distinct neurochemical and behavioral consequences. Crit Rev Neurobiol
12:129-162.
Harshfield GA, Grim CE. 1997. Stress hypertension: the "wrong" genes in
the "wrong" environment. Acta Physiol Scand Suppl 640:129-132.
Hascoet M, Colombel MC, Bourin M. 1999. Influence of age on behavioural
response in the light/dark paradigm. Physiol Behav 66:567-570.
Hengemihle JM, Long JM, Betkey J, Jucker M, Ingram DK. 1999. Agerelated psychomotor and spatial learning deficits in 129/SvJ mice.
Neurobiol Aging 20:9-18.
Heuze P, Feron C, Baudoin C. 1997. Early behavioral development of mice
is affected by staggerer mutation as soon as postnatal day three. Brain
Res Devel Brain Res 101:81-84.
Hilakivi-Clarke LA, Lister RG. 1992. Are there preexisting behavioral characteristics that predict the dominant status of male NIH Swiss mice (Mus
musculus)! J Comp Psych 106:184-189.
Hoffmann HJ, Schenider R, Crusio WE. 1993. Genetic analysis of isolationinduced aggression. II. Postnatal environment influences in AB mice.
Behav Genet 23:391-394.
Hotchin J, Seegal R. 1977. Virus-induced behavioral alteration of mice.
Science 196:671-674.
Ito M. 1984. The modifiable neuronal network of the cerebellum. Jpn J
Physiol 34:781-792.
Jucker M, Ingram DK. 1997. Murine models of brain aging and age-related
neurodegenerative diseases. Behav Brain Res 85:1-26.
Keppel G. 1991. Design and Analysis. A Researcher's Handbook. 3rd ed.
Englewood Cliffs NJ: Prentice Hall.
Koch M, Schnitzler HU. 1997. The acoustic startle response in rats—Circuits mediating evocation, inhibition and potentiation. Behav Brain Res
89:35-49.
Kodama N. 1993. Behavioral development and strain differences in perinatal mice (Mus muscuclus). J Comp Psychol 107:91-98.
Koolhaas JM, De Boer SF, De Rutter AJ, Meerlo P, Sgoifo A. 1997. Social
stress in rats and mice. Acta Physiol Scand Suppl 640:69-72.
Kriegsfeld LJ, Eliasson MJL, Demas GE, Blackshaw S, Dawson TM, Nelson
RJ, Snyder SH. 1999. Nocturnal motor coordination deficits in neuronal
nitric oxide synthase knock-out mice. Neuroscience 89:311-315.
Lalonde R, Filali M, Bensoula AN, Lestienne F. 1996. Sensorimotor learning in three cerebellar mutant mice. Neurobiol Learn Mem 65:113-120.
Lalonde R, Joyal CC, Botez MI. 1994. Exploration and motor coordination
in dystonia musculorum mutant mice. Physiol Behav 56:277-280.
Lathe R. 1996. Mice, gene targeting and behaviour: More than just genetic
background. Trends Neurosci 19:183-186; Discussion 188-189.
Laviola G, Terranova ML. 1998. The developmental psychobiology of
behavioural plasticity in mice: The role of social experiences in the
family unit. Neurosci Biobehav Rev 23:197-213.
Lavond DG, Kim JJ, Thompson RF. 1993. Mammalian brain substrates of
aversive classical conditioning. Annu Rev Psychol 44:317-342.
Le Marec N, Lalonde R. 1997. Sensorimotor learning and retention during
equilibrium tests in Purkinje cell degeneration mutant mice. Brain Res
768:310-316.
Lister RG. 1987. The use of a plus-maze to measure anxiety in the mouse.
Psychopharmacology (Berl) 92:180-185.
Liu D, Diorio J, Tannenbaum B, Caldji C, Francis D, Freedman A, Sharma
S, Pearson D, Plotsky PM, Meaney MJ. 1997. Maternal care, hippocampal glucocorticoid receptors, and hypothalamic-pituitary-adrenal
responses to stress. Science 277:1659-1662.
Lyons JP, Wahlsten D. 1988. Postnatal development of brain and behavior
of Shaker short-tail mice. Behav Genet 18:35-53.
Markowska AL, Spangler EL, Ingram DK. 1998. Behavioral assessment of
the senescence-accelerated mouse (SAM P8 and Rl). Physiol Behav
64:15-26.
McDonald RJ, White NM. 1993. A triple dissociation of memory systems:
Hippocampus, amygdala, and dorsal striatum. Behav Neurosci 107:322.
McGill TE, Tucker GR. 1964. Genotype and sex drive in intact and in
castrated male mice. Science 145:514-515.
McNaughton N. 1991. The role of the hippocampal RSA: Spatial memory
171
versus anxiety. In: Abraham WC, Corballis M, editors. Memory Mechanisms: A Tribute to G. V. Goddard. Hillsdale NJ: Lawrence Erlbaum
Associates, Inc. p 373.
Messier C. 1997. Object recognition in mice: Improvement of memory by
glucose. Neurobiol Learning Mem 67:172-175.
Meunier M, Saint-Marc M, Destrade C. 1986. The Hebb-Williams test to
assess recovery of learning after limbic lesions in mice. Physiol Behav
37:909-913.
Miczek KA, Thompson ML, Shuster L. 1982. Opioid-like analgesia in
defeated mice. Science 215:1520-1522.
Milligan SR, Sales GD, Khirnykh K. 1993. Sound levels in rooms housing
laboratory animals: An uncontrolled daily variable. Physiol Behav
53:1067-1076.
Mishima N, Higashitani F, Teraoka K, Yoshioka R. 1986. Sex differences in
appetitive learning of mice. Physiol Behav 37:263-268.
Miyamoto M. 1997. Characteristics of age-related behavioral changes in
senescence-accelerated mouse SAMP8 and SAMP 10. Exp Gerontol
32:139-148.
Miyamoto M, Kiyota Y, Yamazaki N, Nagaoka A, Matsuo T, Nagawa Y,
Takeda T. 1986. Age-related changes in learning and memory in the
senescence-accelerated mouse (SAM). Physiol Behav 38:399-406.
Moechars D, Lorent K, Dewachter I, Baekelandt V, De Strooper B, Van
Leuven F. 1998. Transgenic mice expressing an alpha-secretion mutant
of the amyloid precursor protein in the brain develop a progressive CNS
disorder. Behav Brain Res 95:55-64.
Mogil JS, Belknap JK. 1997. Sex and genotype determine the selective
activation of neurochemically-distinct mechanisms of swim stressinduced analgesia. Pharmacol Biochem Behav 56:61-66.
Mohn AR, Gainetdinov RR, Caron MG, Koller BH. 1999. Mice with
reduced NMDA receptor expression display behaviors related to schizophrenia. Cell 98:427-436.
Mondadori C, Weiskrantz L. 1993. NMDA receptor blockers facilitate and
impair learning via different mechanisms. Behav Neural Biol 60:205-210.
Morse AC, Erwin VG, Jones BC. 1993. Strain and housing affect cocaine
self-selection and open-field locomotor activity in mice. Pharmacol
Biochem Behav 45:905-912.
Moser VC. 1989. Screening approaches to neurotoxicity: A functional
observational battery. J Am Coll Toxicol 8:85-93.
Moser VC. 1990. Approaches for assessing the validity of a functional
observational battery. Neurotoxicol Teratol 12:483-488.
Mumby DG, Pinel JP. 1994. Rhinal cortex lesions and object recognition in
rats. Behav Neurosci 108:11-18.
Mumby DG, Wood ER, Pinel JP. 1992. Object-recognition memory is only
mildly impaired in rats with lesions of the hippocampus and amygdala.
Psychobiol 20:18-27.
Nalbantoglu J, Tirado-Santiago G, Lahsaini A, Poirier J, Goncalves O, Verge
G, Momoli F, Welner SA, Massicotte G, Julien JP, and others. 1997.
Impaired learning and LTP in mice expressing the carboxy terminus of
the Alzheimer amyloid precursor protein. Nature 387:500-505.
Nelson RJ. 1997. The use of genetic "knockout" mice in behavioral endocrinology research. Horm Behav 31:188-196.
Nelson RJ, Young KA. 1998. Behavior in mice with targeted disruption of
single genes. Neurosci Biobehav Rev 22:453-462.
Nishii K, Matsushita N, Sawada H, Sano H, Noda Y, Mamiya T, Nabeshima
T, Nagatsu I, Hata T, Kiuchi K, Yoshuzato H, Nakashima K, Nagatusu
T, Kobayashi K. 1998. Motor and learning dysfunction during postnatal
development in mice defective in dopamine neuronal transmission. J
Neurosci Res 54:450-464.
Noel M. 1989. Early development in mice: V. Sensorimotor development of
four coisogenic mutant strains. Physiol Behav 45:21-26.
NRC [National Research Council]. 1996. Guide for the Care and Use of
Laboratory Animals. Washington DC: National Academy Press.
Olivier B, van Dalen D. 1982. Social behaviour in rats and mice: An
ethologically based model for differentiating psychoactive drugs. Aggres
Behav 8:163-168.
Onaivi ES, Chakrabarti A, Gwebu ET, Chaudhuri G. 1995. Neurobehavioral
effects of delta 9-THC and cannabinoid (CB1) receptor gene expression
in mice. Behav Brain Res 72:115-125.
172
Packard MG, Hirsh R, White NM. 1989. Differential effects of fornix and
caudate nucleus lesions on two radial maze tasks: evidence for multiple
memory systems. J Neurosci 9:1465-1472.
Parks CL, Robinson PS, Sibille E, Shenk T, Toth M. 1998. Increased anxiety of mice lacking the serotoninl A receptor. Proc Natl Acad Sci U S A
95:10734-10739.
Paulus MP, Dulawa SC, Ralph RJ, Geyer MA. 1999. Behavioral organization is independent of locomotor activity in 129 and C57 mouse strains.
Brain Res 835:27-36.
Platel A, Porsolt RD. 1982. Habituation of exploratory activity in mice: A
screening test for memory enhancing drugs. Psychopharmacology
78:346-352.
Poon AM, Wu BM, Poon PW, Cheung EP, Chan FH, Lam FK. 1997. Effect
of cage size on ultradian locomotor rhythms of laboratory mice. Physiol
Behav 62:1253-1258.
Provencio I, Wong S, Lederman AB, Argamaso SM, Foster RG. 1994.
Visual and circadian responses to light in aged retinally degenerate mice.
Vision Res 34:1799-1806.
Pryor GT, Uyeno ET, Tilson HA, Mitchell CL. 1983. Assessment of chemicals using a battery of neurobehavioural tests. Neurobehav Toxicol
Teratol 5:91-117.
Ramboz S, Oosting R, Amara DA, Kung HF, Blier P, Mendelson M, Mann
JJ, Brunner D, Hen R. 1998. Serotonin receptor 1A knockout: An animal
model of anxiety-related disorder. Proc Natl Acad Sci U S A 95:1447614481.
Riekkinen Jr. P, Riekkinen M, Sirvio J. 1993. Cholinergic drugs regulate
passive avoidance performance via the amygdala. J Pharmacol Exp Ther
267:1484-1492.
Rogers DC, Fisher EM, Brown SD, Peters J, Hunet AJ, Martin JE. 1997.
Behavioral and functional analysis of mouse phenotype: SHIRPA, a
proposed protocol for comprehensive phenotype assessment. Mamm
Genome 8:711-713.
Rogers DC, Jones DNC, Nelson PR, Jones CM, Quilter CA, Robinson TL,
Hagan JJ. 1999. Use of SHIRPA and discriminant analysis to
characterise marked differences in the behavioural phenotype of six
inbred mouse strains. Behav Brain Res 105:207-217.
Rosenwasser AM, Raibert M, Terman JS, Terman M. 1979. Circadian
rhythm of luminance detectability in the rat. Physiol Behav 23:17-21.
Roubertoux PL, Nosten-Bertrand M, Cohnen-Salmon C, l'Hotellier L. 1992.
Behavioral development: a tool for genetic analysis in mice. In:
Goldowitz D, Walhsten D, Wimer RE, editors. Techniques for the
Genetic Analysis of Brain and Behavior. Amsterdam: Elsevier Science
Publishers, p 423-441.
Roubertoux P, Semal C, Ragueneau S. 1985. Early development in mice: II.
Sensory motor behavior and genetic analysis. Physiol Behav 35:659-666.
Roullet P, Lassalle JM, Jegat R. 1993. A study of behavioral and sensorial
bases of radial maze learning in mice. Behav Neural Biol 59:173-179.
Rubinstein M, Mogil JS, Jap6n M, Chan EC, Allen RG, Low MJ. 1996.
Absence of opioid stress-induced analgesia in mice lacking betaendorphin by site-directed mutagenesis. Proc Natl Acad Sci U S A
93:3995-4000.
Saibaba P, Sales GD, Stodulski G, Hau J. 1996. Behaviour of rats in thenhome cages: Daytime variations and effects of routine husbandry procedures analysed by time sampling techniques. Lab Anim 30:13-21.
Sales GD, Wilson KJ, Spencer KE, Milligan SR. 1988. Environmental ultrasound in laboratories and animal houses: A possible cause for concern in
the welfare and use of laboratory animals. Lab Anim 22:369-375.
Sandman CA, Kastin AJ, Schally AV. 1971. Behavioral inhibition as modified by melanocyte-stimulating hormone (MSH) and light-dark conditions. Physiol Behav 6:45-48.
Schellinck HM, Anand KJS. 1999. Consequences of early experience: Lessons for rodent models of newborn pain. In: McGrath P, Finley AG,
editors. Chronic and Recurrent Pain in Children and Adolescents.
Progress in Pain Research and Management, Vol 13. Seattle WA: ISAP
Press, p 39-55.
Schneider R, Hoffmann HJ, Schicknick H, Moutier R. 1992. Genetic analysis of isolation-induced aggression. I. Comparison between closely related inbred mouse strains. Behav Neural Biol 57:198-204.
WAR Journal
Schuchmann S, Muller W, Heinemann U. 1998. Altered Ca2+ signaling and
mitochondrial deficiencies in hippocampal neurons of trisomy 16 mice:
A model of Down's syndrome. J Neurosci 18:7216-7231.
Schulteis G, Koob GF. 1993. Active avoidance conditioning paradigms for
rodents. In: Sahgal A, editor. Behavioural Neuroscience: A Practical
Approach. New York: IRS Press.
Seo ML, Yamatodani A, Mizutani A, Kiyono S. 1993. Short-term,
preweanling treatment with tetracycline affects physical development
and behavior in rats. Neurotoxicology 14:65-75.
Shastry BS. 1995. Genetic knockouts in mice: An update. Experientia
51:1028-1039.
Shepherd GM. 1994. Neurobiology. Oxford: Oxford University Press.
Siegfried B, Frischknecht HR. 1989. Place avoidance learning and stressinduced analgesia in the attacked mouse: Role of endogenous opioids.
Behav Neural Biol 52:95-107.
Silva AJ, Paylor R, Wehner JM, Tonegawa S. 1992. Impaired spatial learning in alpha-calcium-calmodulin kinase II mutant mice. Science
257:206-211.
Simpson EM, Linder CC, Sargent EE, Davisson MT, Mobraaten LE, Sharp
JJ. 1997. Genetic variation among 129 substrains and its importance for
targeted mutagenesis in mice. Nature Genet 16: 19-29.
Smith GW, Aubry JM, Dellu F, Contarino A, Bilezikjian LM, Gold LH,
Chen R, Marchuk Y, Hauser C, Bentley CA, Sawchenko PE, Koob GF,
Vale W, Lee KF. 1998. Corticotropin releasing factor receptor 1deficient mice display decreased anxiety, impaired stress response, and
aberrant neuroendocrine development. Neuron 20:1093-1102.
Soffie M, Hahn K, Terao E, Eclancher F. 1999. Behavioural and glial
changes in old rats following environmental enrichment. Behav Brain
Res 101:37-49.
Sprengel R, Single FN. 1999. Mice with genetically modified NMDA and
AMPA receptors. Ann NY Acad Sci 868:494-501.
Sprott RL. 1997. Mouse and rat genotype choices. Exp Gerontol 32:79-86.
Sprott RL, Staats J. 1975. Behavioral studies using genetically defined mice.
A bibliography. Behav Genet 5:27-82 .
Sprott RL, Staats J. 1981. Behavioral studies using genetically defined
mice—A bibliography (August 1979-July 1980). Behav Genet 11:7384.
Squire LR. 1987. Memory and Brain. New York: Oxford University Press.
Stanford L, Brown RE, Taylor SW, Neumann PE. 1999. Using "mouse IQ"
testing to assess learning and anxiety behaviours in barrelless mutant
mice. (Abstract 256.8). Society for Neuroscience Abstracts 25:641.
Stanford L, Ward NL, Brown RE, Hagg T. 1998. P75NGFR-deficient mice
show faster learning but decreased working memory in the Hebb-Williams maze. (Abstract 74.14). Society for Neuroscience Abstracts 24:179.
Steel KP. 1995. Inherited hearing defects in mice. Annu Rev Genet 29:675701.
Steiner H, Fuchs S, Accili D. 1997. D3 dopamine receptor-deficient mouse:
Evidence for reduced anxiety. Physiol Behav 63:137-141.
Stephens G, McGaugh JL, Alpern HP. 1967. Periodicity and memory in
mice. Psychonomic Sci 8:201-202.
Strazielle C, Lalonde R. 1998. Grooming in Lurcher mutant mice. Physiol
Behav 64:57-61.
Strohle A, Poettig M, Barden N, Holsboer F, Montkowski A. 1998. Ageand stimulus-dependent changes in anxiety-related behaviour of
transgenic mice with GR dysfunction. Neuroreport 9:2099-2102.
Sundberg H, Doving K, Novikov S, Ursin H. 1982. A method for studying
responses and habituation to odors in rats. Behav Neural Biol 34:113119.
Swanson LW, Petrovich GD. 1998. What is the amygdala? Trends Neurosci
21:323-331.
Tanaka T. 1998. Effects of litter size on behavioral development in mice.
Reprod Toxicol 12:613-617.
Tees R. 1974. Effect of visual deprivation on development of depth perception in the rat. J Comp Physiol Psych 86:300-308.
Thompson RF. 1986. The neurobiology of learning and memory. Science
233:941-947.
Tilson HA 1987. Behavioral indices of neurotoxicity: What can be measured? Neurotoxicol Teratol 9:427-443.
Volume 4 1 , Number 3
2000
Tilson HA, Moser VC. 1992. Comparison of screening approaches.
Neurotoxicology 13:1-14.
Tinbergen N. 1951. The Study of Instinct. New York: Oxford University
Press.
Tremml P, Lipp HP, Mueller U, Ricceri L, Wolfer DP. 1998.
Neurobehavioral development, adult openfield exploration and swimming navigation learning in mice with a modified beta-amyloid precursor protein gene. Behav Brain Res 95:65-76.
Trullas R, Skolnick P. 1993. Differences in fear motivated behaviors among
inbred mouse strains. Psychopharmacology 111:323-331.
Umezawa M, Kogishi K, Tojo H, Yoshimura S, Seriu N, Ohta A, Takeda T,
Hosokawa M. 1999. High-linoleate and high-alpha-linoleate diets affect
learning ability and natural behavior in SAMR1 mice. J Nutr 129:431437.
Upchurch M, Wehner JM. 1988. Differences between inbred strains of mice
in Morris water maze performance. Behav Genet 18:55-68.
Van Abeeleen JHF. 1980. Direct genetic and maternal influences on behavior and growth in two inbred mouse strains. Behav Genet 10:545-551.
Van Abeeleen JHF, Kalkhoven JTR. 1970. Behavioural ontogeny of the
Nijmegen waltzer, a neurological mutant in the mouse. Anim Behav
18:711-718.
Vandenbergh JG, Huggett CL. 1995. The anogenital distance index, a predictor of the intrauterine position effects on reproduction in female house
mice. Lab Anim Sci 45:567-573.
Van de Weerd HA, Baumans V, Koolhaas JM, van Zutphen LF. 1994.
Strain specific behavioural response to environmental enrichment in the
mouse. J Exp Anim Sci 36: 117-127.
Van de Weerd HA, Van Loo PL, Van Zutphen LF, Koolhaas JM, Baumans
V. 1997. Nesting material as environmental enrichment has no adverse
effects on behavior and physiology of laboratory mice. Physiol Behav
62:1019-1028.
Vogel JR, Beer B, Clody DE. 1971. A simple and reliable conflict procedure
for testing anti-anxiety agents. Psychopharmacologia 21:1-7.
Vom Saal FS. 1981. Variation in phenotype due to random intrauterine
positioning of male and female fetuses in rodents. J Reprod Fertil 62:633650.
Vorhees CV. 1983. Influence of early testing on postweaning performance
in untreated F344 rats, with comparisons to Sprague-Dawley rats, using
a standardized battery of tests for behavioral teratogenesis. Neurobehav
Toxicol Teratol 5:587-591.
Vorhees CV. 1985. Comparison of the collaborative behavioral teratology
study and Cincinnati behavioral teratology test batteries. Neurobehav
Toxicol Teratol 7:625-633.
Wahlsten D. 1999. Planning genetic experiments: Power and sample size.
In: Jones B, Mormede P. editors. Neurobehavioral Genetics Methods
and Applications. New York: CRC Press, p 31-43.
Ward R. 1980. Some effects of strain differences in the maternal behavior of
inbred mice. Dev Psychobiol 13:181-190.
Wehner JM, Bowers BJ, Paylor R. 1996. The use of null mutant mice to
study complex learning and memory processes. Behav Genet 26:301312.
White NM. 1997. Mnemonic functions of the basal ganglia. Curr Opin
Neurobiol 7:164-169.
Wilkinson GA, Farinas I, Backus C, Yoshida CK, Reichardt LF. 1996.
Neurotrophin-3 is a survival factor in vivo for early mouse trigeminal
neurons. J Neurosci 16:7661-7669.
Williams E, Scott JP. 1954. The development of social behavior patterns in
the mouse, in relation to natural periods. Behavior 6:35-65.
Winters BD, Matheson W, McGregor IS, Brown RE. 2000. An automated
two choice test of olfactory working memory: The effect of scopolomine.
Psychobiology 28:21-31.
Wolfensohn L, Lloyd M. 1998. Handbook of Laboratory Animal Management and Welfare. Oxford: Blackwell Science.
Wuerbel H, Stauffacher M. 1997. Age and weight at weaning affect corticosterone level and development of stereotypies in ICR-mice. Anim
Behav 53:891-900.
Yagi H, Akiguchi I, Ohta A, Yagi N, Hosokawa M, Takeda T. 1998. Spontaneous and artificial lesions of magnocellular reticular formation of
173
brainstem deteriorate avoidance learning in senescence-accelerated
mouse SAM. Brain Res 791:90-98.
Yayou K, Takeda M, Tsubone H, Sugano S, Doi K. 1993. The disturbance
of water-maze task performance in mice with EMC-D virus infection. J
Vet Med Sci 55:341-342.
Young LJ, Winslow JT, Wang Z, Gingrich B, Guo Q, Matzuk MM, Insel
TR. 1997. Gene targeting approaches to neuroendocrinology: Oxytocin,
maternal behavior, and affiliation. Horm Behav 31:221-231.
174
Zhong DZ, Pei C, Xiu Qin L. 1996. Neurobehavioral study of prenatal
exposure to hyperthermia combined with irradiation in mice.
Neurotoxicol Teratol 18:703-709.
Zhu XO, Brown MW, Aggleton JP. 1995. Neuronal signaling of information important to visual recognition memory in rat rhinal and
neighbouring cortices. Eur J Neurosci 7:753-765.
Zorrilla EP. 1997. Multiparous species present problems (and possibilities)
to developmentalists. Dev Psychobiol 30:141-150.
WAR Journal