1

Applied Animal Behaviour Science (2008) Volume 112: 1-32

2

doi: 10.1016/j.applanim.2008.02.007

3

Final Revision – NOT EDITED by the journal

4

What is it like to be a rat? Rat sensory perception and its

5

implications for experimental design and rat welfare

6

Charlotte C. Burn

7

Department of Clinical Veterinary Science, University of Bristol, Bristol BS40 5DU,

8

UK

9

Running title: Rat sensory perception and its implications

10 11 12 13 14 15 16 17

Correspondence address: Department of Clinical Veterinary Science, University of

18

Bristol, Bristol BS40 5DU, UK

1

19

Abstract

20

This review of rat sensory perception spans eight decades of work conducted across

21

diverse research fields. It covers rat vision, audition, olfaction, gustation, and

22

somatosensation, and describes how rat perception differs from and coincides with

23

ours. As Nagel’s seminal work (1974) implies, we cannot truly know what it is like to

24

be a rat, but we can identify and acknowledge their perceptual biases. These primarily

25

nocturnal rodents are extremely sensitive to light, with artificial lighting frequently

26

causing retinal degeneration, and their vision extends into the ultraviolet. Their

27

olfactory sensitivity and ultrasonic hearing means they are influenced by

28

environmental factors and conspecific signals that we cannot perceive. Rat and human

29

gustation are similar, being opportunistic omnivores, yet this sense becomes largely

30

redundant in the laboratory, where rodents typically consume a single homogenous

31

diet. Rat somatosensation differs from ours in their thigmotactic tendencies and highly

32

sensitive, specialised vibrissae. Knowledge of species-specific perceptual abilities can

33

enhance experimental designs, target resources, and improve animal welfare.

34

Furthermore, the sensory environment has influences from neurone to behaviour, so it

35

can not only affect the senses directly, but also behaviour, health, physiology, and

36

neurophysiology. Research shows that environmental enrichment is necessary for

37

normal visual, auditory, and somatosensory development. Laboratory rats are not

38

quite the simple, convenient models they are sometimes taken for; although very

39

adaptable, they are complex mammals existing in an environment they are not

40

evolutionarily adapted for. Here, many important implications of rat perception are

41

highlighted, and suggestions are made for refining experiments and housing.

42 43

Keywords: Animal Welfare; Communication; Olfaction; Perception; Rats;

44

Refinement; Vision

45 2

46 47 48

1

49

The stimuli that an animal can perceive depend on the available sensory apparatus,

50

while the way stimuli are evaluated in terms of their biological relevance depends on

51

the animal’s innate biases, cognitive abilities and experiences. Perception is therefore

52

a subjective distortion of reality, differing between species and even between

53

individuals within a species. Since rats and mice, which have similar perceptual

54

abilities to each other, constitute over 80% of all research animals in the European

55

Union (Commission of the European Communities, 2003), and they have been bred

56

for research since the late 1800s (Krinke, 2000; Whishaw & Kolb, 2005), much is

57

known about their perceptual biases. However, the information is scattered through

58

time and across different research fields, so it is not easily available to researchers, rat

59

caretakers, and other rat specialists. The resulting lack of awareness can have

60

serious implications, sometimes leading to poorly designed experiments and harming

61

rat welfare. This review brings current information together, to help inform and refine

62

rodent experiments and housing.

63

Introduction

The review concentrates on the laboratory rat, Rattus norvegicus, since

64

summaries of mouse sensory perception are included within several other review

65

papers (Sherwin, 2002; Olsson et al., 2003; Latham & Mason, 2004). Much of the

66

information will also be true for mice and other rodents, but care should still be taken

67

if extrapolating between species. The species’ natural ecology – such as whether they

68

are diurnal or nocturnal, social or solitary, arboreal, burrowing or terrestrial – will

69

profoundly affect their sensory perception. These ethological considerations are

70

highly relevant in laboratory rats despite their domestication; adult laboratory rats

71

retain so many of their wild instincts that, when released into a naturalistic habitat,

3

72

their resulting community and behaviour rapidly resembles that of their wild relatives

73

(Berdoy, 2002). This review is organised around the classic ‘five senses’: vision, audition,

74 75

olfaction, gustation and somatosensation. It should be remembered that these are

76

actually not the only senses; indeed rats may even possess a magnetic compass, like

77

mice (Muheim et al., 2006) and hamsters (Deutschlander et al., 2003), but most

78

published information currently covers the aforementioned five senses. For each

79

sense, the rat’s sensory biases relative to humans are first described, then some

80

practical implications of its perception with respect to welfare and experimental

81

design are discussed. This is an applied review, focussing on the known or suspected

82

implications of each sense, and aiming to provide enough information to allow readers

83

to extrapolate to their own situations. The review cannot be completely

84

comprehensive, and it will become clear that in many cases, rat sensory perception is

85

still poorly understood.

86

2

Vision

87

An obvious difference between human and rat vision is that rats’ eyes are

88

located on the sides of their heads, rather than the front. They therefore have a wider

89

field of view, but less binocular overlap than us: wild rats have a binocular overlap of

90

35o, domestic rats 76o, and humans 105o (Heffner & Heffner, 1992a).

91

Wild rats usually inhabit burrows or other enclosed environments, and tend to

92

be nocturnal or crepuscular, so most of their activities occur under low light

93

conditions (e.g. Calhoun, 1963). Consequently, rats rely relatively little on vision, but

94

they are dramatically more sensitive to dim light than we are, able to discriminate tiny

95

increments in intensity, indiscernible to us, including discriminating ‘total

96

darkness’ from 0.107 lux (Campbell & Messing, 1969).

4

97

Rats, especially albinos, have much poorer visual acuity (Lashley, 1938; Creel

98

et al., 1970; Prusky et al., 2002) and narrower depth perception than humans

99

(O'Sullivan & Spear, 1964; Routtenberg & Glickman, 1964). For example, human

100

acuity can be around 30 c/d (‘cycles per degree’ – a measure of spatial resolution

101

accounting for stimulus size and distance), while pigmented rats’ acuities are only 1–

102

1.5 c/d and albino strains have even lower acuities of 0.5 c/d (Prusky et al., 2002).

103

This presumably gives an extremely blurred image by human standards (Figure 1,

104

reprinted from Prusky and colleagues, 2002). Poor acuity in rats is probably partly

105

due to their eyes’ relatively small size, and partly because their eyes appear to

106

have very limited abilities to focus light from different distances or angles

107

compared with human eyes (Artal et al., 1998). Rats often bob their heads which

108

may help them gain motion cues about the distance of objects (Legg & Lambert,

109

1990).

110

Experiments in the 1930s suggested that, contrary to popular belief, rats possess

111

colour vision (e.g. Munn & Collins, 1936; Walton & Bornemeier, 1938), which has

112

recently been confirmed through electroretinograms and quantitative behavioural tests

113

(Jacobs et al., 2001). Rod cells comprise 99% of rat photoreceptors, but rats also have

114

two cone cell types (Szel & Rohlich, 1992). Around 93% of the cones respond

115

maximally to blue–green light (around 510 nm), while the remaining 7% respond to

116

ultraviolet (UV) (around 360 nm) (Jacobs et al., 2001; Akula et al., 2003). Cone

117

responses are normally distributed, so rats actually perceive hues ranging from

118

ultraviolet (400 nm) to orange-red (around 635 nm) (Jacobs et al., 2001), but they are

119

most responsive to colours near their peak sensitivities (Jacobs et al., 2001; Akula et

120

al., 2003).

121

Flicker fusion thresholds (when emitted light flickers rapidly enough to appear

122

constant) for rats are not yet known, but are relevant for their perception of video

123

images and artificial lighting (D'Eath, 1998). Flicker fusion thresholds decrease with

124

high light intensity, and increase with fatigue. Animals with high proportions of rod 5

125

cells, like rats, generally have high flicker fusion thresholds, so rats might perceive

126

videos, computer monitors, and some fluorescent lighting as flickering (Jarvis et al.,

127

2003).

128

Discussion of the implications of rat vision is separated according to sensitivity

129

to light generally, colour vision, periodicity, and acuity.

130

2.1

131

The sensitivity of rats to light (Campbell & Messing, 1969) means that light levels

132

comfortable for humans can rapidly cause retinal atrophy (reviewed in Schlingmann

133

et al., 1993a; Schlingmann et al., 1993b) and cataract formation in rats (Rao, 1991).

134

Albinos are particularly susceptible because they lack protective melanin in the iris

135

and retinal epithelium, and the entire eyeball is slightly transparent (Schlingmann et

136

al., 1993b). Consequently, even when the iris contracts in bright light, most of the

137

light still enters the eye (Williams et al., 1985). In fact, albino rats may be the most

138

susceptible of all laboratory animals to light-induced retinal degeneration (Bellhorn,

139

1980).

140

Sensitivity to light

To illustrate the relevant range of light intensities, the UK code of practice for

141

the care and use of laboratory animals suggests that “350–400 lux at bench level is

142

adequate for routine experimental and laboratory activities” (Home Office, 1989).

143

Light intensities within cages are commonly between about 150 and 550 lux

144

(Schlingmann et al., 1993c), but are higher in laboratory rooms, with upper limits

145

approaching 10,000 lux due to current technological limitations (e.g. Light Therapy

146

ProductsTM, 2006; Outside In Ltd., 2006). Humans can tolerate still higher intensities

147

– outdoors on sunny days light often exceeds 50,000 lux, and only at this order of

148

magnitude are discomfort and potential retinal damage likely in humans.

149

Light intensities of only 65 lux can cause retinal degeneration in albino rats,

150

even on a 12 h light-dark cycle (Semple-Rowland & Dawson, 1987). Half the

151

photoreceptors were permanently damaged after just 3 days at 133 lux in albinos, but 6

152

pigmented rats were less susceptible, with equivalent damage occurring at 950 lux

153

(Williams et al., 1985). Rod cells are particularly vulnerable to light destruction, but

154

cones often survive even after all rods have been destroyed (Cicerone, 1976; La Vail,

155

1976). Long-term cyclical light intensities of about 500 lux within an animal room

156

can also cause cataracts in albino rats (Rao, 1991). These problems are worst in rats

157

housed closest to the light source, usually those highest in the rack (Rao, 1991; Perez

158

& Perentes, 1994).

159

Surprisingly, some vision can remain after constant long-term light exposure,

160

even when no intact photoreceptor cells can be observed (e.g. Lemmon & Anderson,

161

1979). This might be conferred by a few remaining cones that may be so sparse that

162

they were undetectable by the quantitative techniques used (Cicerone, 1976; La Vail,

163

1976). Even so, under ‘ordinary’ laboratory conditions, visual impairments can

164

confound some tests. For example, in the Morris water maze – a test of cognitive

165

function – rats with incidental light-induced retinal damage perform as poorly as rats

166

with cognitive deficits, both groups displaying difficulties locating the platform

167

(Osteen et al., 1995; Lindner et al., 1997). Also, in commonly used ‘anxiety’ tests,

168

such as open field tests and light-dark boxes, visually impaired individuals might

169

venture into the exposed/light areas more than fully sighted ones, through their lesser

170

ability to discriminate light from dark, but this requires experimental confirmation.

171

Therefore, light-induced retinopathy should be controlled for in such tests, or non-

172

visual tests used alongside the established visual ones.

173

Welfare problems might arise at even lower light levels than those causing

174

retinal damage, because of motivation to hide, as well as to avoid ocular discomfort

175

(Schlingmann et al., 1993c). Rats, especially albinos, reliably choose the lowest light

176

intensities available, even when all the choices are very dim, appearing

177

indistinguishable to humans (Campbell & Messing, 1969; Woodhouse & Greenfeld,

178

1985; Blom et al., 1995). Rats’ aversion to light was clearly demonstrated in a study

179

showing that sleeping pigmented and albino rats awoke and moved to areas of lower 7

180

illumination at thresholds of only 60 and 25 lux, respectively (Schlingmann et al.,

181

1993c). Consistent with such behaviour, chromodacryorrhoea, an aversion-related

182

secretion from the Harderian gland (e.g. Mason et al., 2004), increases with brighter

183

light (Hugo et al., 1987).

184

There is clearly a conflict between human workers needing adequate light to

185

inspect rats, for example for signs of illness, and rats needing to avoid damaging or

186

aversive light levels. Schlingmann (1993a) therefore stresses the importance of

187

providing shelters within cages, allowing rats some control over their light exposure.

188

As described below, coloured shelters exist that allow humans to see rodents, while it

189

supposedly appears dark to the rodents inside the shelter, although their efficacy

190

requires confirmation.

191

Light levels affect commonly used psychological tests, such as elevated plus-

192

mazes in which exploration of the exposed arms is taken to indicate reduced anxiety;

193

rats explore the exposed arms more in dim than bright light (Cardenas et al., 2001;

194

Garcia et al., 2005). Moreover, some effects are only found under certain light

195

conditions. For example, the anxiolytic effects of gentling only emerge in brightly lit

196

open fields (Hirsjarvi & Valiaho, 1995), and some drug effects are influenced by plus

197

maze illumination (Clenet et al., 2006). Therefore, some control and careful

198

description of lighting conditions during these tests is necessary to account for its

199

influence on psychological measures.

200

Surgery presents a difficult situation because good lighting is essential for

201

delicate operations, but the anaesthetised, unblinking rat is unable to protect its eyes

202

from that light. Care should therefore be taken, not only to keep the eyes hydrated, but

203

also to protect them from prolonged bright light. Interestingly, the anaesthetic agent,

204

halothane, prevents retinal degeneration (Keller et al., 2001); other anaesthetics have

205

not yet been investigated. This protection is afforded under white, but not blue, light.

8

206

Despite the above evidence that bright light is harmful to rats, this aspect of

207

their biology is not always considered in some fields of research. An example is the

208

use of rats as models for seasonal affective disorder in humans, exploring whether

209

bright light therapy (up to 11,500 lux for 2 weeks) can cure depression in rats (e.g.

210

Dilsaver & Majchrzak, 1988; Giroux et al., 1991; Humpel et al., 1992; Overstreet et

211

al., 1995). Unsurprisingly, the depression was not cured, and the one study that

212

considered the effects of light on rat vision discovered massive destruction of the

213

albinos’ photoreceptors (Humpel et al., 1992). These examples illustrate how crucial

214

knowledge of species-specific perception is for generating reasonable hypotheses and

215

preventing animal suffering.

216

2.2

217

Rats are not colour-blind (Muenzinger & Reynolds, 1936; Munn & Collins, 1936;

218

Walton & Bornemeier, 1938; Lemmon & Anderson, 1979; Jacobs et al., 2001).

219

However, relative to humans, they perform poorly when discriminating between

220

colours of similar wavelengths (Walton, 1933), and they take longer to learn colour

221

discriminations than light intensity ones (Jacobs et al., 2001).

Colour vision

222

To discuss the implications of rats’ colour sensitivity, the implications for

223

emitted light and that reflected by objects in the environment will be dealt with

224

separately, as their effects are quite distinct.

225

2.2.1

226

Standard artificial lighting rarely emits UV wavelengths (e.g. Bellhorn, 1980; Latham

227

& Mason, 2004), since human cones are insensitive to it. To date, no studies have

228

apparently investigated the effects of UV-deficient light on rats. In some birds, UV

229

light is important for their welfare (Moinard & Sherwin, 1999; Maddocks et al., 2001)

230

and normal behaviour (Bennett & Cuthill, 1994), but laboratory mice appear to have,

231

if anything, a slight aversion to it (C. M. Sherwin, personal communication). Also,

232

high levels of UV can cause cataracts in mice (in Bellhorn, 1980), and can affect

Emitted light

9

233

reproductive and circadian rhythms in rats (reviewed in Brainard et al., 1994). In fact,

234

the colour composition of artificial light can have large effects. In rats, blue light

235

(around 490 nm) causes most retinal degeneration (reviewed in Schlingmann et al.,

236

1993b), and also more disruption to fertility (Tong & Goh, 2000) than any other

237

wavelengths tested; UV light was not included in these studies, but is of a shorter

238

wavelength than blue light so may be more harmful.

239

At the opposite end of the spectrum, dim red light is sometimes used to observe

240

nocturnal behaviour in rats, because it is on the upper edge of the wavelengths visible

241

as colour to them (Jacobs et al., 2001). However, rats’ rod cells are stimulated by

242

similar wavelengths to human rod cells, including red light (Akula et al., 2003). This

243

means that, provided some rod cells remain intact, rats can see red light, even if only

244

as light and dark contrast. This may not be a problem in experiments if rats are

245

habituated to it, since moonlight would provide illumination in the wild. As an

246

alternative to red light, sodium lamps, which emit very narrow peaks of yellow–

247

orange (589 and 589.6 nm) light, can be used (McLennan & Taylor-Jeffs, 2004). Not

248

only is it more visible to humans than red light, but there were no long-term

249

differences between the activity levels of mice when illuminated by this lamp or in

250

darkness. However, in studies unequivocally requiring rats to behave as if in pitch

251

darkness, infra-red light and the necessary viewing equipment should be used.

252

It is also worth noting that most video equipment and computer monitors, which

253

create images using emitted light, include no UV emissions and the colour balance is

254

optimised for human vision (D'Eath, 1998). Even in black-and-white images and light

255

from white artificial light bulbs, ‘white’ is composed of red, green and blue light

256

adjusted for humans, and so would not appear as white to rats. Therefore, any such

257

images presented to species with different colour sensitivities, particularly UV-

258

sensitive animals, could lack important information.

10

259

2.2.2

Colour in the environment

260

Caution is required when presenting images to rats in discrimination tests, even if the

261

cues reflect rather than emit light. Different inks have different spectral properties that

262

may be invisible to the human eye, and some might even reflect UV. Moreover,

263

different pigments might differ in their olfactory qualities, which could be more

264

salient to rats than their visual qualities. Even if this does not harm the experimental

265

purpose, it can make standardisation between experiments difficult.

266

Outside experimental situations, there are also some relevant implications of

267

rodent colour vision within the homecage. In recent years, manufacturers of rodent

268

environmental enrichments have produced transparent shelters in various colours (e.g.

269

Robbins, 2004; Datesand Ltd, 2005). The idea behind them is that, while rodents -

270

supposedly blind to the shelter’s colour - perceive themselves as being sheltered in a

271

dark environment, human carers can inspect them without disturbing them. However,

272

these shelters seem not to have been independently evaluated for their efficacy. Red

273

transparent material might make a suitable shelter, being the least visible colour to

274

rats (Jacobs et al., 2001), but as explained earlier, it would still stimulate rod cells and

275

possibly some cones. The colour of the homecage itself might also affect rats. Sherwin and Glen

276 277

(2003) housed mice in different coloured cages and found that they had significantly

278

different preferences for cage-colours. Moreover, the colour affected their food-to-

279

body mass conversion rates and their elevated plus-maze anxiety. Assuming these

280

effects were due to the colours directly (rather than the scents, tastes, or textures of the

281

dyes used), this study shows that environmental colour can have surprisingly strong

282

effects on mouse behaviour and physiology, and so possibly that of rats too.

283

2.3

284

Rats tend to be most active at dusk and dawn, although their circadian rhythms are

285

relatively flexible (e.g. Calhoun, 1963). Because we are diurnal, many rodent

Periodicity

11

286

experiments are carried out in the light, so much of our knowledge of this species

287

comes from individuals awakened during their resting period, and tested under much

288

brighter conditions than they would voluntarily experience. The implications of this

289

can be profound, but time-shifted experiments are still rare in some fields. The brain

290

state changes radically between sleep and activity, with whole populations of neurons

291

shifting between activity and inactivity (Hobson, 2005; Saper et al., 2005). The time

292

of testing can strongly influence the variables of interest in experiments. For example,

293

during the light phase, rats’ cardiovascular responses to various stressors are more

294

pronounced (Schnecko et al., 1998), and they show less exploratory behaviour in an

295

elevated plus-maze than in the dark phase (Andrade et al., 2003).

296

For most experiments, rats will be in a wakeful state provided they have

297

sufficient time to awaken, but little published information is available on how long

298

rodents require to fully awaken (i.e. be in the same state as during the active phase).

299

Any conclusions drawn from light phase studies of rats as human models could suffer

300

from interpretive problems, because it is unclear whether the observed state would

301

reflect a similar state in our light (active) phase or our (dark) resting phase.

302

Time shifted experiments and husbandry can be made possible by using red or

303

sodium illumination as described above, and also by feeding rats only during the

304

phase when we wish them to be active (cited in Saper et al., 2005); a situation that

305

sometimes occurs in the wild (Calhoun, 1963).

306

2.4

307

As described above, rats have very poor acuity (Figure 1). Their image resolution is at

308

least 20 times poorer than ours (Artal et al., 1998). Note though that the studies

309

investigating rat visual acuity (Lashley, 1938; Creel et al., 1970; Artal et al., 1998;

310

Robinson et al., 2001; Prusky et al., 2002) have used laboratory rats, whose acuity

311

might have been further reduced by their artificially lit environments.

Acuity

12

312

Apart from the damaging effects of light itself, several other factors can affect

313

rat vision, including the early environment. Complete lack of light impairs rats’ visual

314

development (Fagiolini et al., 1994), but providing environmental enrichment to these

315

dark-reared animals can eliminate this effect (Bartoletti et al., 2004). In mice,

316

enriched environments during rearing accelerate visual development and improve

317

adult acuity (Prusky et al., 2000; Cancedda et al., 2004).

318

Also, diet has a large influence on vision (Berson, 2000). For example, caloric

319

restriction can prevent cataracts (e.g. Wolf et al., 2000), and antioxidant intake and

320

consumption of certain vitamins can prevent retinal damage (Li et al., 1985; Berson,

321

2000). Dietary composition is discussed in more detail in the Gustation section of this

322

paper.

323

The research implications of rats’ poor visual acuity depends on the experiment

324

in question, but if visual cues are used they should be relatively large and high

325

contrast, but not too bright as to be aversive. Also, visual cues may not be as salient to

326

rats as cues in other modalities. Few experiments have tested this directly, but rats do

327

remember auditory associations for longer than equivalent visual ones (Wallace et al.,

328

1980), and can more rapidly learn discriminations using multimodal stimuli (floor

329

surfaces differing in appearance, smell, and texture Dymond, 1995; Dymond et al.,

330

1996) or olfactory or tactile cues (Birrell & Brown, 2000). However, vision is often

331

the most appropriate sense for guiding rats in water mazes (Prusky & Douglas, 2005),

332

for comparison with past studies, and for certain models of human activities.

333

3

334

Sound can be described in terms including its frequency, intensity, timbre (frequency

335

spectrum) and envelope (shape of sound pressure through time). While young

336

humans hear frequencies from about 0.02 kHz to 20 kHz (Moore, 2003), hearing in

337

rats is shifted upwards to include the ultrasonic range (Kelly & Masterton, 1977). The

Audition

13

338

lowest frequency rats have been reported to hear is 0.25 kHz and the highest is 80 kHz

339

(Kelly & Masterton, 1977; Heffner & Heffner, 1992b; Heffner et al., 1994). They can

340

also detect lower sound frequencies (Petounis et al., 1977), probably through contact

341

with vibrating surfaces, and can even perceive low frequency sounds using their

342

vibrissae (Neimark et al., 2003) (see the section on Somatosensation). Auditory sensitivity decreases near the extremes of the detectable frequencies,

343 344

so sounds at the lower and higher extremes must be louder before rats can detect

345

them. The rat’s peak sensitivity is estimated to lie between about 8 and 50 kHz (Kelly

346

& Masterton, 1977; Heffner & Heffner, 1992b), although estimates vary, probably

347

due to factors including strain, age, and background noise. Even whether the

348

homecages of rats are barren or environmentally enriched can greatly affect hearing

349

sensitivity; auditory neurone performance is vastly improved by environmental

350

enrichment (Engineer et al., 2004).

351

The implications of rat auditory perception include what sound characteristics

352

are harmful, vocal communication between rats, perception of the human voice, and

353

experimental use of sound cues. There has also been debate about whether rats can

354

echolocate.

355

3.1

356

Interactions between sound intensity and frequency (Fleshler, 1965; Voipio et al.,

357

1998; Björk et al., 2000) make it difficult to determine detection- and safety-

358

thresholds for sound intensities. The decibel (dB) scale is logarithmic, so even small

359

numerical increases represent large increases in the actual intensity. European

360

Union legislation (2003) states that advice and hearing-protection must be provided

361

for human workers frequently exposed to sounds of 80 dB or more. Above about 150

362

dB, auditory damage is inevitable with most perceivable sounds (Gamble, 1982).

363

Equivalent thresholds are unknown for rats, but young rats are more sensitive to

Audiogenic damage in the laboratory

14

364

sounds than older ones, and permanent audiogenic damage is most likely in pups

365

between about 12 and 22 days of age (Voipio, 1997).

366

In the laboratory, audible sounds as loud as 80–90 dB have been recorded; and

367

50–75 dB for ultrasound (Milligan et al., 1993), so conceivably, audiogenic damage

368

could occur in both humans and rats. Husbandry procedures cause the loudest sounds,

369

especially if metallic equipment is involved (Gamble & Clough, 1976; Milligan et al.,

370

1993; Sales et al., 1999). Filling metal food hoppers made 80 dB of (mostly

371

ultrasonic) sound, which would occur about once a week for the rats’ lifetimes (Sales

372

et al., 1999). This was measured from a distance of 50 cm, approximately the furthest

373

that a caged rat could get from the sound.

374

Many apparently silent activities or devices actually produce high levels of

375

ultrasound (Sales et al., 1988; Sales et al., 1999). Examples include computer

376

monitors, making 68–84 dB of broadband ultrasound (Sales et al., 1988), and some

377

fluorescent lighting (G. J. Mason personal communication, and personal observation).

378

Cage washers, hoses, running taps, squeaky chairs, and rotating glass stoppers (Sales

379

et al., 1988) produce both ultrasound and audible sound, as do some air-flow hoods

380

worn to prevent allergy in human workers (Picciotto et al., 1999). Similarly, standard

381

fire alarms produce loud high and low frequency sounds, which laboratory animals

382

cannot escape, so laboratories can be fitted with fire alarms that only emit sound

383

audible to humans but not rodents (Home Office, 1989); although note that even

384

frequencies below rats’ audible range can affect them (Petounis et al., 1977).

385

Whether common laboratory sounds affect rodent welfare has not been

386

investigated directly, but loud noises generally can trigger seizures, reduce fertility,

387

and cause diverse metabolic changes (Sales et al., 1988; Milligan et al., 1993).

388

Repeated short bursts of 2 kHz sound at 120 dB caused ‘behavioural despair’ in rats

389

(Bulduk & Canbeyli, 2004). Longer-lasting sounds can also affect animals, although

390

that has apparently not been tested in rats. In pigs, 90 dB prolonged or intermittent

15

391

broadband noise increased cortisol, ACTH, noradrenaline:adrenaline ratios and time

392

lying down, and decreased growth and social interactions (Otten et al., 2004).

393

Conceivably then, a fluorescent light emitting loud ultrasound could cause significant

394

stress in rats housed near it. The envelopes and timbres of sounds also determine how aversive or damaging

395 396

they are. Noise-type sounds, e.g. white noise or the sound of tearing paper, cause

397

stronger fear reactions in rats than equivalent harmonic or pure tones, or audible rat

398

vocalisations (Voipio, 1997). Sudden sounds are probably also more startling than

399

those with gradual onsets. It should be noted that avoidance of sound occurs at still

400

lower thresholds than those causing startle reactions (in Fleshler, 1965), or physical

401

damage.

402

Ultrasound detectors (e.g. bat detectors), which represent ultrasounds in a form

403

that humans can hear or visualise, would be useful as standard pieces of laboratory

404

equipment to regularly check whether ultrasound of certain frequencies is being

405

emitted in the animal rooms and to test experimental set-ups. Few experimenters

406

would choose to carry out experiments during loud building work, for example,

407

because of potential effects on the animals’ performances, and the same

408

meticulousness should apply to ultrasound. Indeed, background noise levels during

409

behavioural experiments do affect the apparent learning abilities of rats, with louder

410

white noise leading to faster completion of a maze task (Prior, 2006). Moreover, even

411

loud infrasound affects rat behaviour, reducing their activity and triggering sleep

412

(Petounis et al., 1977).

413

3.2

414

As well as audible ‘squeaks’, rats produce at least three types of ultrasonic

415

vocalisations. Firstly, juvenile rats produce a 40–50 kHz vocalisation (Noirot, 1968),

416

which together with olfactory cues, causes pup-retrieval by the mother (e.g. Allin &

417

Banks, 1972; Farrell & Alberts, 2002).

Vocalisations and communication

16

418

The second ultrasonic vocalisation is the ‘22 kHz long-call’, which occurs

419

mainly in aversive situations and might therefore indicate negative affect (Knutson et

420

al., 2002). Examples of such situations include social defeat (Van der Poel & Miczek,

421

1991), exposure to cat odour (Blanchard et al., 1991), administration of naloxone or

422

lithium chloride (Burgdorf et al., 2001), arthritic pain without analgesia (Calvino et

423

al., 1996), acute pain (Jourdan et al., 1995), acoustic startle (Kaltwasser, 1990) and

424

electric shocks (Kaltwasser, 1991). However, male rats make a similar vocalisation

425

after ejaculation (Van der Poel & Miczek, 1991), so this call might occur in two

426

subtly different forms, or might not reliably indicate negative affect.

427

The third ultrasonic vocalisation is the ‘50 kHz chirp’, which is apparently

428

associated with positive events (Knutson et al., 2002), and has even been suggested as

429

a form of laughter (Panksepp & Burgdorf, 2000). It occurs in anticipation of positive

430

social contact (Knutson et al., 1998; Brudzynski & Pniak, 2002), rewarding ‘tickling’

431

by humans (Panksepp & Burgdorf, 2000; Burgdorf & Panksepp, 2001; Panksepp,

432

2006), amphetamine or morphine administration (Knutson et al., 1999), and feeding

433

or rewarding electrical stimulation of the brain (Burgdorf et al., 2000), and also during

434

play (Knutson et al., 1998; Brudzynski & Pniak, 2002; Burn, 2006). However, again,

435

this vocalisation does not reliably indicate positive affect because it occurs in some

436

aversive situations, e.g. during morphine withdrawal (Vivian & Miczek, 1991),

437

aggression (Sales, 1972), and in certain painful situations (Hawkins et al., 2005).

438

Surprisingly little work has investigated the audible squeak. There may in fact

439

be several different types of squeak, because subjectively there is variation in the

440

quality of sounds produced (O. H. P. Burman, personal communication; personal

441

observation). Pups and their mothers make audible squeaks in the nest (e.g. Voipio,

442

1997), but this may be different from squeaking in other contexts. Squeaks occur

443

during nociception as they persist even when central nervous system analgesics are

444

given, which might suggest that they are detached from the emotional experience of

445

pain (Jourdan et al., 1995). They also occur during playing and fighting (Voipio, 17

446

1997; Burn et al., 2006a), and sometimes during handling, especially alongside

447

struggling behaviour (van Driel et al., 2004; Burn, 2006). They generally seem to

448

indicate negative affect, but do not necessarily occur alongside the 22 kHz long-call,

449

so there must be some qualitative or quantitative difference between the motivations

450

behind the two call types.

451

All of these vocalisations could have practical implications. Procedures or

452

environments that cause rats to vocalise could affect the behaviour and physiology of

453

all neighbouring rats within audible range. For example, playbacks of 22 kHz long-

454

calls caused freezing and decreased activity (Sales, 1991; Brudzynski & Chiu, 1995)

455

and increased latencies to emerge into an arena (Burman et al., 2007). Playbacks of

456

audible squeaks also caused conspecifics to orientate towards the speaker and

457

occasionally to squeak themselves (Voipio, 1997).

458

3.3

459

An awareness that rats can hear our voices is important, because of affects on

460

experimental results and rat welfare. Rats can hear and discriminate many elements of

461

the human voice (e.g. Pons, 2006), and pet rats can learn to respond to verbal

462

commands (e.g. Fox, 1997). In fact, rats can distinguish between some languages

463

(Toro et al., 2003), so the pitches, rhythms and accents of different human workers

464

could be at least partly responsible for rats being able to distinguish between

465

individual humans (McCall et al., 1969; Morlock et al., 1971; Davis et al., 1997; van

466

Driel & Talling, 2005). Shouting causes stress responses in farm animals

467

(Hemsworth, 2003), so this may also be true for laboratory rats, especially because

468

when humans speak with more emotional content, the higher-pitched and ultrasonic

469

content of our speech increases (Mason, 1969).

470

3.4

471

By default, most standard recording devices and speakers include no ultrasound, so

472

specialised equipment is necessary, such as ‘tweeter’ speakers and ultrasonic

Perception of the human voice

Sound recordings and playbacks

18

473

microphones (Björk et al., 2000). White noise, although aversive to rats (Voipio,

474

1997), is commonly used to standardise background noise in experiments, but

475

different speakers differ in their ultrasonic output, so comparisons across studies

476

might sometimes be invalid. Even a study that specifically investigated how

477

background noise affected rat behaviour in a maze, neither mentioned their ultrasonic

478

hearing abilities, nor used specialist equipment to produce the experimental white

479

noise (Prior, 2006), indicating that awareness of these auditory issues may be lacking

480

in some fields.

481

3.5

Echolocation There has been some debate about whether rats can echolocate (e.g. Rosenzweig

482 483

et al., 1955; Riley & Rosenzweig, 1957; Kaltwasser & Schnitzler, 1981; Forsman &

484

Malmquist, 1988). Blind rats can use self-generated sounds, reflected off solid

485

objects, to guide them in mazes (Rosenzweig et al., 1955; Riley & Rosenzweig,

486

1957). Also, sighted rats in darkness can discriminate between shelves close enough

487

to jump to and those too far away, but not if they are deafened (Chase, 1980). Some

488

studies described quiet ultrasonic ‘clicks’ (Chase, 1980; Graver et al., 2004), which

489

were produced more in darkness than in light, more before rats jumped to the platform

490

than after, and the decision to jump was faster in rats that clicked more (Graver et al.,

491

2004). However, rats seem not to have anything like the specialised echolocation

492

abilities of mammals such as bats or cetaceans. Indeed, some blind and blindfolded

493

humans can ‘echolocate’ using reflected sound, similar to rats (in Riley &

494

Rosenzweig, 1957), but there is no evidence that either species can use sound to build

495

up a detailed picture of their environment, as bats or cetaceans can.

496

4

497

Rats rely heavily on olfaction (e.g. Doty, 1986). They can quickly associate olfactory

498

cues with food rewards (Le Magnen, 1999a; Birrell & Brown, 2000), with this ability

499

even making them a suitable alternative to ‘sniffer’ dogs for locating contraband

Olfaction

19

500

substances (Otto et al., 2002). Rats can locate the direction of odorants, without

501

moving their heads, three orders of magnitude more quickly than we can (Rajan et al.,

502

2006). It is sometimes stated that albinism dampens olfaction, because albinos show

503

weaker avoidance of garlic than pigmented rats do (Keeler, 1942), but of course they

504

might simply be less averse to the scent. Humans are unusual mammals because a much smaller proportion of our

505 506

genome is devoted to olfaction, than other species (Gilad et al., 2003; Emes et al.,

507

2004; Rat Genome Sequencing Project Consortium, 2004; Quignon et al., 2005), and

508

our vomeronasal organ is vestigial or non-existent (e.g. Brennan & Keverne, 2004). In

509

contrast, rats not only possess main olfactory epithelia, but also well-developed

510

vomeronasal organs. Although the two systems overlap (reviewed in Shepherd, 2006),

511

the vomeronasal organ seems specialised for instinctive recognition of pheromones

512

and evolutionarily relevant compounds (Dulac, 1997; Holy et al., 2000; Brennan &

513

Keverne, 2004), while the olfactory epithelium is specialised for learned associations

514

between volatile scents and their implications (Dulac, 1997). The vomeronasal system

515

detects relatively non-volatile compounds, requiring the rat to lick or imbibe some

516

compounds before it can detect them (Brennan & Keverne, 2004). Here ‘olfaction’

517

includes both systems, because in most cases the specific odorant or detection

518

mechanism is currently unknown. The focus is on olfactory communication, but some

519

significant scents within laboratory environments are also discussed.

520

4.1

521

Rat olfactory communication is well-developed, yet remains little understood by

522

humans. Much communication is mediated through urine, but rats have many scent

523

glands, including the sebaceous, preputial, clitoral, perineal, salivary, anal, plantar,

524

and Harderian glands. Through scent, rats can gain information about each others’

525

gender (Alberts & Galef, 1973; Moore, 1985; Brown, 1992; Garcia-Brull et al., 1993),

526

reproductive state (Gawienowski et al., 1975; Manzo et al., 2002; Zala et al., 2004),

Overview of rat olfactory communication

20

527

genetic relatedness (Wills, 1983; Hurst et al., 2005), dominance (Krames et al., 1969),

528

health status (Zala et al., 2004), and individual identity (Hopp et al., 1985; Gheusi et

529

al., 1997). Rats also recognise familiar conspecifics using olfaction (Burman &

530

Mendl, 2003), not through a shared ‘colony scent’, but through remembering

531

individual odours (Alberts & Galef, 1973; Carr et al., 1976). These odours can be

532

determined genetically or be acquired from the environment (Schellinck et al., 1991;

533

Schellinck & Brown, 2000; Hurst et al., 2005). Laboratory rats may not be completely isolated from conspecifics even when

534 535

individually housed, because scents from neighbouring cages, or experimental

536

apparatus and instruments can influence them (unless they are in individually

537

ventilated cages). These scents can profoundly affect rats, as described below,

538

although it should be mentioned that isolation itself also affects these social animals

539

(e.g. Day et al., 1982; Hurst et al., 1997; Sharp et al., 2002; Westenbroek et al., 2005).

540

4.2

541

Much sexual behaviour in rodents is olfactorily mediated. The ‘Bruce effect’,

542

whereby female mice abort their offspring upon encountering the volatile scent of

543

unfamiliar males (Bruce & Parrott, 1960), seems not to occur in rats. However, the

544

‘Whitten effect’, in which volatile male scents trigger oestrus in females (Whitten,

545

1959), and the ‘Lee–Boot effect’, when females housed without males show

546

suppressed, irregular oestrus cycles (Van Der Lee & Boot, 1956) do occur relatively

547

weakly in rats. In rats and mice, male odour accelerates the onset of puberty in

548

females, in a phenomenon labelled the ‘Vandenbergh effect’ (Vandenbergh, 1969,

549

1976).

550

Scent and reproduction

The scent of female rats, especially those in oestrus, stimulates male sexual

551

behaviour, but also urinary-marking (Manzo et al., 2002) and competitive aggression

552

(Alberts & Galef, 1973). It is possible therefore, that housing males where they can

553

smell females could affect their physiology and behaviour, affecting research, and 21

554

might affect their welfare either way. The vomeronasal system, probably responsible

555

for detecting these scents, habituates to stimuli less easily than most sensory systems

556

(Holy et al., 2000), so the effects might be persistent. However, since the vomeronasal

557

organ requires direct physical contact to detect some pheromones (Brennan &

558

Keverne, 2004), the problem might only exist if the scent is volatile. Other important scents here include those mediating the mother–pup

559 560

relationship. For example, diodecyl proprionate, a pup preputial gland pheromone,

561

induces maternal licking (Brouettelahlou et al., 1991). Mother rats produce various

562

odours aiding pup survival, including those guiding pups to the nipples, and those

563

deposited in the bedding that reduce pup activity, keeping them in the nest (Porter &

564

Winberg, 1999). Also, pregnant females release a non-volatile pheromone that

565

prevents infanticide by cohabiting males (Mennella & Moltz, 1988). Perhaps it is the

566

removal of these scents that increases the likelihood of pups being cannibalised when

567

rats’ cages are cleaned within the first few days of birth (Burn & Mason, in press).

568

4.3

569

Aggression in male rodents can be triggered by novel (usually male) scents, so rats

570

rendered anosmic show little aggression in resident–intruder tests (Alberts & Galef,

571

1973). Habituation to familiar or self-scents plays a large role in reducing aggression

572

between familiar or related individuals. For example, aggression is reduced between

573

more familiar individuals (Alberts & Galef, 1973; Garcia-Brull et al., 1993) and

574

between more closely related individuals (Nevison et al., 2003). Some inbred mouse

575

strains cannot discriminate between familiar and unfamiliar conspecific odours,

576

resulting in reduced aggression (Nevison et al., 2003). This could also be true for rats.

577

In fact, unfamiliar male scents not only stimulate aggression, but also defensive

578

behaviour in subordinate males encountering dominant male odours. Rats defeated by

579

an alpha-male, subsequently show avoidance and fear behaviour upon encountering

580

the scent of other alpha-males (Williams & Groux, 1993; Williams, 1999).

Olfactory modulation of aggression

22

581

This said, while cage-cleaning – which removes scent marks – provokes

582

aggression in male mice (Gray & Hurst, 1995; Van Loo et al., 2000), in familiar rats it

583

merely provokes non-aggressive skirmishing (Burn et al., 2006a; Burn et al., 2006b);

584

perhaps for this reason cage-cleaning frequency seemingly has no long-term effects

585

on male rat welfare.

586

When unfamiliar rats are to be housed together, exposing them to each other’s

587

scents for a few days before allowing physical contact may prevent aggression (e.g.

588

Bulla, 1999). Alternatively, aggression can sometimes be prevented by masking

589

unfamiliar conspecifics using another unfamiliar, neutral scent. In rats evidence is

590

anecdotal, but in a controlled study of mice, chocolate or sheep’s wool odours reduced

591

resident–intruder aggression (Kemble et al., 1995).

592

Finally, it is worth mentioning that odour-mediated aggression does not only

593

occur between males. For example, mother rats able to smell their own pups show

594

aggression towards intruders – neither visual, tactile, nor auditory cues from the pups

595

elicit this aggression (Ferreira & Hansen, 1986).

596

4.4

597

Rats are generally attracted to areas smelling of conspecifics (e.g. Galef & Heiber,

598

1976; Mackay-Sim & Laing, 1980), but scents released during negative or positive

599

experiences, can make those areas aversive or more attractive, respectively.

600

Communication about experiences

Rats produce ‘alarm’ odour when they experience electric shocks (Mackay-Sim

601

& Laing, 1980; Abel & Bilitzke, 1990; Williams & Groux, 1993; Kiyokawa et al.,

602

2004), transport between rooms (Beynen, 1992), and the events and disturbances

603

accompanying carbon dioxide euthanasia (Ware & Mason, 2003). They probably also

604

produce it in forced-swim tests (Abel & Bilitzke, 1990), but no unstressed controls

605

were used so rats may simply have been responding to odours left by an unfamiliar

606

male. Alarm odour is more powerful with more severe stressors (Mackay-Sim &

607

Laing, 1980). The molecule(s) involved have not yet been identified, but a candidate 23

608

is 2-heptanone; more of this is present in urine from stressed rats, but diazepam during

609

the stressor does not reduce the amount produced (Gutiérrez-García et al., 2006).

610

In recipients, alarm odour increases freezing behaviour (Williams, 1999;

611

Kikusui et al., 2001), activity (Mackay-Sim & Laing, 1980; Abel & Bilitzke, 1990;

612

Kikusui et al., 2001; Ware & Mason, 2003), body temperature (Kikusui et al., 2001),

613

hypothalamic–pituitary–adrenal activity (Takahashi et al., 1990; but see Mackay-Sim

614

& Laing, 1980), urination (Stevens & Koster, 1972), and latency to approach rewards

615

(Mackay-Sim & Laing, 1981; Ware & Mason, 2003). It also causes avoidance

616

compared with the scent of unstressed conspecifics (Mackay-Sim & Laing, 1980).

617

Experience can affect responses to alarm odour, with rats avoiding the odour of

618

shocked rats more if they have experienced shock themselves, but not necessarily if

619

they have experienced defeat by an alpha-male (Williams & Groux, 1993).

620

A somewhat separate body of literature describes ‘frustration’ or ‘non-reward’

621

odour, produced when anticipated rewards are withheld (Collerain & Ludvigson,

622

1972; Ludvigson et al., 1985; Taylor & Ludvigson, 1987). Again this odour causes

623

avoidance, but unlike alarm odour, no fear responses to it have been reported. It seems

624

not to exist in urine (Collerain & Ludvigson, 1972), unlike alarm odour (Mackay-Sim

625

& Laing, 1981), but both are also produced from other bodily sources yet to be

626

identified (Mackay-Sim & Laing, 1981; Weaver et al., 1982).

627

Rats probably also produce a ‘reward’ odour, although this has mainly been

628

tested against non-reward situations (i.e. frustration odour), with no neutral rat odour

629

control. Nevertheless, a rat’s trail is more attractive if laid down after the rat receives

630

a reward than before (Galef & Buckley, 1996), and when it perceives a signal that

631

reliably predicts reward (Ludvigson et al., 1985). However, the attraction of rats to

632

reward odour is much weaker than the avoidance of frustration odour, when compared

633

against the same ‘no odour’ control (Taylor & Ludvigson, 1980).

24

634

The release of alarm odour means that rat welfare and experimental aims might

635

be compromised if neighbouring conspecifics are distressed by illness, injury, or

636

experimental procedures (Beynen, 1992). Any of these odours can bias rats’ decisions

637

in choice tests (Collerain & Ludvigson, 1972; Aoyama & Okaichi, 1994; Mitchell et

638

al., 1999), increase ‘baseline’ stress in subsequently tested rats or supposed control

639

ones (Beynen, 1992; Kikusui et al., 2001), and alter behaviour in tests such as swim

640

tests (Abel & Bilitzke, 1990), and open field or novelty tests (Mackay-Sim & Laing,

641

1981; Takahashi et al., 1990; Ware & Mason, 2003). There has apparently been no evaluation of effective ways to clean experimental

642 643

apparatus; various cleaning agents are used, which probably vary in efficacy and may

644

have intrinsic odours that affect rats. Alcohol is commonly used, but in pigs, its

645

volatile components can reduce cortisol levels in open field tests (Thodberg et al.,

646

2006).

647

4.5

648

Rats can learn about specific foods from conspecific odours. Carbon disulphide,

649

present in rats’ breath (Galef et al., 1988), causes rats to strongly prefer novel foods

650

eaten by their cagemates versus other novel foods (e.g. Strupp & Levitsky, 1984). The

651

preferences can persist for at least 30 days, even without opportunity to sample the

652

foods during that time (Galef & Whiskin, 2003b).

653

Communication about food

Aversion to novel foods can be caused by the ‘poisoned partner effect’ (Lavin et

654

al., 1980). Here if a novel food is eaten by a rat, which then encounters the odour of a

655

poisoned conspecific, the healthy rat will subsequently avoid the novel food, even if

656

the poisoned rat did not eat it (Stierhoff & Lavin, 1982). Strangely, the healthy rat

657

only avoids food that it itself has eaten, rather than that eaten by the poisoned rat, and

658

therefore not necessarily the poisonous food (Galef et al., 1990). In fact, exactly as

659

described above, the healthy rat actually prefers novel foods after smelling them on

660

the poisoned rat’s breath (Galef et al., 1990). 25

661

Lactating rats also avoid novel foods ingested just before their pups become ill,

662

because of an odour released by pups with gastrointestinal illness (Gemberling, 1984).

663

The odour causes no aversion in males or nulliparous females, and is not released by

664

pups stressed in other ways, so it seems more specific than the poisoned partner

665

effect.

666

4.6

667

Most of the scents relevant to laboratory rats are those within the cage itself. Apart

668

from those produced by conspecifics or food, others could include detergent residues,

669

bedding materials, and microbial products from the breakdown of food or excreta.

670

Cage-cleaning abruptly changes the olfactory environment, which might contribute to

671

post-cleaning changes in rat behaviour and physiology (Burn et al., 2006a). Also, like

672

gerbils, rats might more accurately discriminate scents in a test arena on days when

673

their cages are clean rather than soiled (Dagg et al., 1971).

Scents in the laboratory

674

Another salient source of smell for laboratory rodents might be their human

675

handlers. Rats respond differently towards different humans (McCall et al., 1969;

676

Morlock et al., 1971; Davis et al., 1997; van Driel & Talling, 2005), mostly because

677

of differences in odour (McCall et al., 1969). People smell different due to genetic

678

factors and environmental ones, such as diet, smoking, perfume, soap, and deodorant.

679

Regular rodent handlers may also be ‘marked’ with odours from previously handled

680

rodents, sometimes including reward or alarm odours.

681

Additionally, rats might fear humans carrying scents from their pets, especially

682

if the pet is a predatory species. Rats innately fear predator odours, including cats and

683

mustelids (reviewed in Blanchard et al., 2003), but apparently not dogs. Rats cannot

684

easily habituate to predator odours (Blanchard et al., 1998), showing increased

685

corticosterone, freezing and vigilance, elevated plus-maze anxiety and endogenous

686

opioid analgesia, and suppressed electric-prod burying, and impaired working

687

memory (Williams, 1999; Blanchard et al., 2003). Predator odours also elicit fear26

688

related fast-waves and reduce cell-proliferation in the dentate gyrus (Heale et al.,

689

1994; Tanapat et al., 2001).

690

It is even possible that rats would instinctively fear human odour – wild rats

691

usually avoid close human contact, and any such fear of humans might have escaped

692

our notice, of course, it would require a controlled experiment not involving human

693

presence.

694

Many odours from synthetic products used in laboratories could affect rodents.

695

While several reviews compare the efficacy of detergents for cleaning animal cages

696

(e.g. Heuschele, 1995), none discuss their potential olfactory impacts on the animals.

697

Yet, some organic solvents (e.g. xylene, toluene, diethyl ether, and methyl

698

methacrylate) cause avoidance and fast-waves in the dentate gyrus, just as predator

699

odours do (Heale et al., 1994). These solvents constitute many everyday substances,

700

including some inks, glues, and paints; indeed, identification-marking rodents with

701

inks or dyes can affect their anxiety profiles (Burn et al., in press) and cause them to

702

become submissive to unmarked cagemates (Lacey et al., 2007).

703

Many odorants that smell subjectively pleasant to humans, often therefore being

704

present in perfumed products or human diets, can also influence hypothalamo-

705

pituitary-adrenal activity and immune responsiveness, positively or negatively

706

(Komori et al., 2003). Rose oil (de Almeida et al., 2004) and ‘green odour’, trans-2-

707

hexenal (Nakashima et al., 2004), are anxiolytic to rats. Citrus oils are analgesic

708

(Aloisi et al., 2002), but can have complex effects on rodent anxiety (Komori et al.,

709

2003; Ceccarelli et al., 2004). In rat pups, peppermint increases mortality and

710

decreases activity (Pappas et al., 1982), and rats avoid the scent of garlic (Keeler,

711

1942) and rosemary (R. M. J. Deacon, personal communication). Many of these

712

effects could inadvertently introduce variation between experiments, but some could

713

be used as non-nutritive environmental enrichments or rewards. Also, anxiolytic

27

714

scents could be easily administered to rats in mildly stressful situations (de Almeida et

715

al., 2004; Nakashima et al., 2004).

716

5

717

Like us, rats are opportunistic omnivores; their ecological niche is characterised by

718

sampling diverse food substances and remembering their nutritional consequences

719

(e.g. Capaldi, 1996). They rapidly learn aversions to harmful novel foods, which can

720

be a problem in pest control situations when they ingest sub-lethal quantities of bait.

721

Rats, particularly wild strains, are neophobic, being reluctant to consume novel food

722

(Galef & Whiskin, 2003a). They initially sample only small amounts of novel food (if

723

any at all), but if it proves safe, they later readily consume it, often in preference to

724

more familiar foods (Calhoun, 1963). Under natural conditions, this cautious but

725

explorative behaviour might help them obtain a full nutritional complement, reducing

726

reliance on any one food type, while avoiding poisoning.

727

Gustation

Rats detect similar taste dimensions to humans, i.e. sweetness (carbohydrates

728

and artificial sweeteners), saltiness (sodium salts), sourness (hydrogen ions),

729

bitterness (quinine, caffeine, most natural toxins, and some others) (Grill & Norgren,

730

1978), and umami (amino acids, such as glutamate) (e.g. Smith & Margolskee, 2001).

731

As with humans, sweetness and umami are rewarding, bitterness is usually aversive,

732

and saltiness and sourness are only pleasant at low concentrations (Grill & Norgren,

733

1978; Berridge, 2000). They also initially strongly avoid capsaicin, the ‘hot’ taste of

734

chilli, but often consume it readily once it becomes familiar (Jensen et al., 2003).

735

However, rats do not perceive certain artificial sweeteners as being ‘sweet’ (Sclafani

736

& Abrams, 1986; Dess, 1993; Sclafani & Clare, 2004), and they may have separate

737

receptors for sugars and starch (Sclafani, 1987). Their bitterness thresholds for some

738

compounds differ from ours (Glendinning, 1994; Mueller et al., 2005), allowing

739

denatonium benzoate – which tastes less bitter to rats than to humans and some other

740

animals – to be added to baits to prevent its consumption by non-target species 28

741

(Hansen et al., 1993). There are also some strain and sex differences in rat gustation

742

(Boakes et al., 2000; Clarke et al., 2001).

743

In fact, ‘flavour’ involves not only gustation, but also olfaction and tactile

744

sensations (Smith & Margolskee, 2001). For completeness, these senses are not

745

separated here when discussing the practical implications of rat gustatory biases.

746

5.1

747

Laboratory rodents usually have no opportunity to sample different foods, typically

748

being fed a palatable, dry, nutritionally complete diet, in powder form or as pellets.

749

These diets are easily stored, inexpensive, and require little preparation (Lane-Petter,

750

1975), and they aid standardisation between experiments. Laboratory rats will also

751

taste their mothers’ milk, bodily secretions from themselves or conspecifics (if

752

socially housed), their cage surfaces, and perhaps human hands or gloves, and

753

bedding material (if provided). Hence, scope for learning taste–nutrient associations is

754

very limited, rendering the gustatory sense largely redundant in laboratories.

755

Taste in the laboratory

For other sensory modalities, sensory deprivation reduces the volume and

756

functioning of the associated brain regions. For example, the visual cortices of rats

757

reared in darkness are permanently underdeveloped (Fagiolini et al., 1994), while

758

sensory deprivation only temporarily limits olfactory bulb (Cummings et al., 1997)

759

and barrel cortex development (Polley et al., 2004; but see Rema et al., 2003).

760

However, despite rats frequently being used as models in taste research, precisely

761

because their gustatory perception is supposedly similar to ours, the effects of

762

gustatory deprivation on the brain and behaviour are apparently unknown. The effects

763

may be minimal if taste is tightly genetically controlled, but alternatively, lack of

764

gustatory experience could, for example, exaggerate rats’ neophobia or diminish their

765

gustatory learning abilities.

29

766

5.2

767

It is unclear whether rats can appropriately self-regulate their nutritional intake, given

768

the opportunity. Most discrepancies between findings are probably due to differences

769

between the diets offered to rats (Naim et al., 1985; Sclafani, 1987; Prats et al., 1989),

770

and circadian variations in intake patterns (Larue-Achagiotis et al., 1992). Rats

771

generally do select foods appropriate for their changing nutritional needs, but like

772

humans, they are biased towards sugary or fatty foods. They are consequently also

773

prone to obesity if offered palatable, calorific diets (Naim et al., 1985; Sclafani, 1987;

774

Prats et al., 1989).

775

Nutritional regulation

Because laboratory rodent diets are homogenous, they allow no qualitative

776

nutritional regulation. Generally, this is unproblematic because the diets have

777

sustained rodent populations for many decades, without apparent negative effects on

778

breeding, health, or longevity. However, although special formulations are available,

779

many widely used diets cover all age and sex categories: oestrus females, weanling

780

pups, and elderly males alike. Moreover, they are often common to rats and mice.

781

Thus, within this diversity, individuals might sometimes have different nutritional

782

requirements from that provided. In standardising diets to this extent, we might

783

inadvertently increase, rather than decrease, variation in rodents’ internal nutritional

784

states because they have no opportunity to regulate them.

785

Some dietary supplements can enhance laboratory rat health, calling into

786

question the completeness of homogenous diets. For example, blueberries, high in

787

antioxidants, prevent cognitive deficits in aging rats (Casadesus et al., 2004), and as

788

mentioned previously, other dietary supplements prevent retinal damage (Li et al.,

789

1985). Also, in hamsters, supplementation with seeds and rabbit chow increased pup

790

growth, and reduced cannibalism by the mothers (Day et al., 2002).

30

791

5.3

Refinement within the homecage

792

Palatable diets may provide rats with ‘enjoyment’ (Lane-Petter, 1975) or hedonic

793

experiences, with palatable and unpalatable foods eliciting distinctive behavioural

794

expressions that are homologous to human gustatory expressions (Berridge, 2000).

795

Most welfare efforts concentrate on reducing negative welfare, but facilitating

796

positive welfare, such as pleasure from food or foraging, should not be neglected (e.g.

797

Balcombe, 2005). Food-related environmental enrichments might be particularly

798

relevant for generalists, like rats, because their natural ecology incorporates diverse

799

food types, varying through time and space. However, the idea of food-related

800

enrichment has been little explored for laboratory rats, and yet it could improve their

801

welfare (Johnson & Patterson Kane, 2003), provided obesity is avoided (e.g. Mattson,

802

2005). There are three main aspects of food that could be varied for enrichment

803

purposes: nutritional content, flavour, and physical presentation.

804

5.3.1

805

Providing rodents with very nutritionally diverse diets may be undesirable for

806

practical reasons (Lane-Petter, 1975; Key, 2004), and because they encourage obesity

807

(Mattson, 2005), and may increase variation. Nevertheless, offering some opportunity

808

to nutritionally self-regulate could be beneficial, as suggested above. In some animal

809

facilities, seeds and nuts are scattered onto rats’ bedding; rats become very active

810

upon hearing them being scattered in neighbouring cages, and continue foraging for

811

many hours (Key, 2004). Since the seeds would constitute only a very small

812

proportion of the diet, they are unlikely to impact heavily on nutritional regulation,

813

but could allow some relevant gustatory stimulation and regular hedonic experiences.

814

Proper evaluation of the effects is necessary however; the most relevant study so far

815

seems to be one, mentioned earlier, when seed supplements enhanced hamster pup

816

growth and decreased cannibalism (Day et al., 2002).

Nutritional content

31

817

5.3.2

818

Even without nutritional value, gustatory enrichment could be achieved; providing

819

daily non-nutritional pina-colada flavour treats to breeding mice increased the number

820

of pups weaned (Inglis et al., 2004), suggesting that the hedonistic aspects alone of

821

scatter-feeding are beneficial.

822

Flavour

Domesticated rats value variety, and will substitute a preferred food that has

823

been their sole diet for several days for a less preferred, newly available food (Galef

824

& Whiskin, 2003a, 2005). They also consume more food if provided as a succession

825

of varied ‘meals’ rather than homogenous meals (Treit et al., 1983; Clifton et al.,

826

1987; Le Magnen, 1999b). These preferences exist even when foods differ primarily

827

in flavour not nutritional value, such as when cinnamon, cocoa, ‘all spice’, or

828

marjoram are added to normal chow (as in the above five studies). These additives

829

presumably have negligible bioactivity, being common non-nutritive components of

830

human diets, but confirmation in rats is required. The above studies suggest that

831

obesity might be a risk because of the increased food consumption, but they were all

832

relatively short-term, so rats might down-regulate their intake of variable food over

833

time. Le Magnen (1999b) found that if ‘variable days’ were alternated with

834

‘homogenous days’, rats ate less food than normal on homogenous days, perhaps

835

compensating for over-eating on variable days.

836

5.3.3

837

Finally, enrichment might be achieved through varying dietary presentation. Soft ‘wet

838

mash’ (chow soaked in water) is often used to help sick or weak rats gain weight, and

839

usually any healthy cage-mates also prefer the mash to freely available pellets.

840

However, it is an impractical enrichment for healthy rats, being messy and

841

encouraging microbial growth (Lane-Petter, 1975). Occasionally scattering chow

842

pellets within the cage allows rats to eat in their natural posture, holding the pellet in

Physical presentation

32

843

their forepaws (Bruce, 1965), and they more readily consume these pellets than those

844

in the hopper (personal observation). Captive rats also ‘contra-freeload’, choosing food that requires handling and

845 846

preparation, even when prepared food is available (Carder & Berkowitz, 1970). This

847

may be because most of a wild rat’s time and effort would be devoted to foraging

848

(Johnson & Patterson Kane, 2003). Scattering small food items, such as the

849

aforementioned seed mixes or chow pellets, in bedding allows rats to forage, which

850

may be rewarding in itself. Scatter-feeding rarely triggers competitive aggression

851

because the food is spatially distributed. Commercially available rodent puzzle-

852

feeders are also available, although they are uncommon in laboratories and are not

853

always easily sourced.

854

5.4

855

The generalist feeding habits of rats can be exploited in research, improving

856

experiments ethically, enhancing rats’ cooperation, and reducing interference from

857

stress. Drugs and inoculants are often delivered by gavage, a tube inserted via the

858

mouth into the stomach, which can be technically difficult, and causes stress,

859

respiratory distress, and occasionally even death (Balcombe et al., 2004). However,

860

substances can be successfully delivered within palatable vehicles that rats will

861

voluntarily consume, provided there is no interference with the active ingredient.

862

Fruit- or beef-flavoured gelatine is commonly used but some rats only reluctantly

863

consume it, so it can be worth trying several alternatives (Hawkins et al., 2004).

864

Another example is to use small amounts of chocolate (Huang-Brown & Guhad,

865

2002). Taste aversion can develop if the vehicle becomes associated with illness, but

866

giving rats prior experience with the unadulterated food can prevent this. Some

867

substances can also be microencapsulated and added to chow for long-term studies

868

(Melnick et al., 1987; Dieter et al., 1993; Yuan et al., 1993).

Refinement of experiments

33

869

Preferred rewards can often be used to motivate rats to perform tasks in

870

experiments, rather than using punishments or prior deprivation. Deprivation is a

871

powerful motivator, but can undesirably affect behaviour, physiology,

872

neurochemistry, and drug efficacy (Slawecki & Roth, 2005). Moreover, it is

873

sometimes unnecessary, because undeprived rats will often work – albeit to a limited

874

extent – for preferred rewards, including commercially available reward pellets,

875

sucrose solution (Slawecki & Roth, 2005), or breakfast cereals (e.g. Ellis, 1984). Prats

876

and colleagues (1989) found that rats did not readily consume cheese, chocolate or

877

fruit-candy, and instead preferred other foods offered, including banana, cookies,

878

standard chow pellets, and liver pâté. Large quantities of dairy products (DiBattista,

879

1990) and chocolate (Huang-Brown & Guhad, 2002) should be avoided as they harm

880

rodent health. Undeprived rats are particularly motivated to earn rewards if

881

experiments coincide with their active period (Hyman & Rawson, 2001), with a

882

shifted light cycle enabling practical working hours (see the section on Vision).

883

Neophobia can be eliminated by providing the palatable incentive in the homecages of

884

rats several days before experiments.

885

Finally, food must often be withheld overnight before surgery or intraperitoneal

886

injections. This deprivation causes weight loss, and reduced hepatic weight and blood

887

glucose, and potentially, emotional distress from hunger. However, providing sugar

888

cubes to the rats can prevent these problems, while gastrointestinal volume is still

889

reduced, as required (Levine & Saltzman, 1998).

890

6

891

Rat somatosensation could be considered from many different angles. Here, the focus

892

is on that relating to the ability of rats to explore and interact with their environments.

893

In the rat somatosensory cortex, the vibrissae (sensory whiskers), nose and mouth,

894

forepaws, and sinus hairs on its wrists, are particularly well-represented. In fact, the

895

forepaws are represented twice each, and the whiskers and sinus hairs have

Somatosensation

34

896

specialised granular aggregates devoted to them (Hermer-Vazquez et al., 2005). In

897

general, rat and human somatosensation seem similar, but there are two main

898

differences that noticeably affect rat behaviour. Firstly, rats’ vibrissae are extremely

899

sensitive (Arabzadeh et al., 2005), being comparable to primate fingertips (Carvell &

900

Simons, 1990). Rats can whisk them independently of each other across surfaces to

901

make fine tactile discriminations (Guic-Robles et al., 1989; Carvell & Simons, 1990).

902

In a study investigating rats’ numerical competencies, subjects could not discriminate

903

between two, three or four tactile stimuli delivered to the body, but they succeeded

904

when the stimuli were delivered to a single vibrissal hair (Davis et al., 1989). The

905

vibrissae also detect differences in mechanical resonant frequencies, with the shorter

906

anterior vibrissae detecting higher frequencies than the longer posterior ones

907

(Neimark et al., 2003).

908

The second obvious difference from humans relates to thigmotaxis; the bias of

909

rats towards maintaining physical contact with vertical surfaces. In fact, thigmotaxis

910

underlies many tests of ‘anxiety’ (Treit & Fundytus, 1988), because when rats

911

perceive environments as threatening, they stay closer to vertical surfaces, such as the

912

boundaries of open field arenas, or the closed arms of elevated plus-mazes. The

913

thigmotactic bias may not be strictly somatosensory, perhaps also incorporating visual

914

preferences for avoiding light exposure. Rats that lack vibrissae on one side prefer to

915

maintain wall-contact on their intact side, suggesting the vibrissae play a role (Meyer

916

& Meyer, 1992). The implications of rat somatosensation include the impact of environmental

917 918

enrichment on rat somatosensory development generally, and implications of the

919

vibrissal sense for experiments and housing.

920

6.1

921

Environmental enrichment profoundly affects the somatosensory and barrel cortices.

922

In rats kept in enriched rather than barren environments, the primary somatosensory

Environmental enrichment and somatosensation

35

923

cortex representing the forepaws becomes 1.5 times larger (Xerri et al., 1996; Coq &

924

Xerri, 1998, 2001). The barren cages in these studies contained bedding, exerting their

925

effect despite rats being able to dig with their forepaws, so the difference might be

926

even more pronounced in rats housed on wire floors. Environmental enrichment

927

seemingly does not enhance textural discrimination abilities, but it does increase the

928

rate of learning such discriminations (Bourgeon et al., 2004). Enrichment can also

929

counteract age-related declines in hind-paw representation in the somatosensory

930

cortex, which is otherwise associated with impaired walking in aged rats (Godde et

931

al., 2002). Finally, in naturalistic environments, the representation of each whisker in

932

the barrel cortex becomes dramatically more well-defined compared with standard

933

cages (Polley et al., 2004).

934

The above studies combined several enrichment types, including social

935

contact, foraging opportunities, structural features and novelty, so it is unclear what

936

relative contributions were made by each enrichment type. It is lack of tactile contact

937

with conspecifics that apparently leads to the self-biting and tail manipulation seen in

938

isolated rats (Day et al., 1982; Hurst et al., 1997).

939

6.2

940

The sensitivity of the vibrissal sense (Davis et al., 1989; Guic-Robles et al., 1989;

941

Carvell & Simons, 1990; Arabzadeh et al., 2005) is probably under exploited in

942

learning tasks, where less salient visual cues are currently more widely used

943

(Dymond, 1995; Dymond et al., 1996; Birrell & Brown, 2000). However, laboratory

944

rats can sometimes lack vibrissae for various reasons, including ‘barbering’, when

945

hairs and often whiskers are removed by conspecifics (Garner et al., 2004). This

946

occurs in rats, albeit to a much lesser extent than in mice (Bresnahan et al., 1983;

947

Wilson et al., 1995). Other rats may lack whiskers due to their strain; some nude

948

rodent strains have no whiskers at all (e.g. Sundberg et al., 2000), but most have short,

949

kinked whiskers, giving a limited sensory range (e.g. Festing et al., 1978; Moemeka et

Vibrissae and the laboratory environment

36

950

al., 1998). Nude strains also lack the sensitive guard hairs otherwise dispersed through

951

the coat, and which would convey proprioceptive information. Both vibrissal absence and barrel cortex impairment through lack of

952 953

environmental enrichment (as described above), could have practical consequences.

954

Rats lacking vibrissae show impaired orientation towards tactile stimuli, and –

955

provided they have environmental enrichment – compensate by orienting towards

956

visual stimuli more than controls (Symons & Tees, 1990). Whiskers also aid

957

swimming, enabling animals to keep their heads above water (Ahl, 1986; Meyer &

958

Meyer, 1992), and consequently, rats lacking vibrissal sensation can drown in water

959

mazes and swim tests (Hughes et al., 1978).

960

Finally, vibrissae are important in social interactions, with whiskerless rats

961

being unable to avoid bites to their faces during fighting (Blanchard et al., 1977a;

962

Blanchard et al., 1977b). Because aggression between familiar rats is uncommon

963

(Burn et al., 2006b), whiskerless rats need not be socially isolated, except in cases

964

where aggression is observed. However, whiskerless rats may be injured if introduced

965

to unfamiliar conspecifics, when fighting is more likely.

966

7

967

It is impossible for us to know what it is like to be a `rat` (Nagel, 1974), but

968

knowledge of their sensory biases allows us to imagine what it might be like, as a

969

human, to have those biases within a laboratory rat’s environment. This insight, while

970

imperfect, could help predict how rats might be affected by different situations,

971

improving our experimental design and their welfare. In summing up then, an overall

972

theoretical picture of a rat’s perception of the laboratory could be as follows.

973

Summary

The rat’s sensitive eyes, shunning the intense artificial light, provide it with a

974

hazy view in predominantly grey, ultra-violet and green hues. From within its cage, it

975

hears the chirps, squeaks and whines of its neighbours, gaining information that we 37

976

cannot hear unaided and are yet to understand. Background noise consists of the low

977

babbles and hisses of distinctively scented humans, and the unregulated drones and

978

blasts of ultrasonic sounds. Scents provide visceral warnings and enticements, induce

979

new motivations, and inform the rat about social possibilities outside the cage. The

980

environment wafts a succession of scents, from pleasant, calming fragrances to the

981

innately alarming odours of intangible predators. The rat tastes little apart from its

982

dry, satiating homogenous diet. Its vibrissae provide a protective, finely tuned force-

983

field to feel the details of the cage surfaces; with the rat perceiving security from close

984

contact with the solid walls.

985

8

986

Conclusion Knowledge of the sensory gulfs and similarities between ourselves and this

987

commonly used research animal can improve science and enhance rat welfare. More

988

work is still necessary to understand rat perception, and even more so for less well-

989

researched species. The aim of this review is to make current knowledge accessible to

990

researchers, rat caretakers and rodent specialists, in the hope that it will enable

991

tangible improvements in experimental design and rat welfare.

992

Acknowledgements

993

Many thanks to Georgia Mason for her detailed comments and encouragement,

994

and also to Robert Deacon, Mark Ungless, Jennifer Bizley, and Alex Weir for their

995

comments.

996

References

997 998 999 1000 1001 1002 1003

Abel, E. L., Bilitzke, P. J., 1990. A possible alarm substance in the forced swimming test. Physiol. Behav. 48, 233-239. Ahl, A. S., 1986. The role of vibrissae in behavior: a status review. Vet. Res. Commun. 10, 245-268. Akula, J. D., Lyubarsky, A. L., Naarendorp, F., 2003. The sensitivity and spectral identity of the cones driving the b-wave of the rat electroretinogram. Vis. Neurosci. 20, 109-117. 38

1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053

Alberts, J. R., Galef, B. G., 1973. Olfactory cues and movement: stimuli mediating intraspecific aggression in the wild Norway rat. J. Comp. Physiol. Psychol. 85, 233-242. Allin, J. T., Banks, E. M., 1972. Functional aspects of ultrasound production by infant albino rats (Rattus norvegicus). Anim. Behav. 20, 175-185. Aloisi, A. M., Ceccarelli, I., Masi, F., Scaramuzzino, A., 2002. Effects of the essential oil from citrus lemon in male and female rats exposed to a persistent painful stimulation. Behav. Brain Res. 136, 127-135. Andrade, M. M., Tome, M. F., Santiago, E. S., Lucia-Santos, A., de Andrade, T. G., 2003. Longitudinal study of daily variation of rats' behavior in the elevated plus-maze. Physiol. Behav. 78, 125-133. Aoyama, K., Okaichi, H., 1994. The influence of conspecific distress responses on the lever choice behavior in the rat. Jpn. J. Psychol. 65, 286-294. Arabzadeh, E., Zorzin, E., Diamond, M. E., 2005. Neuronal encoding of texture in the whisker sensory pathway. PLoS Biology 3, e17. Artal, P., Herreros De Tejada, P., Muñoz Tedó, C., Green, D. G., 1998. Retinal image quality in the rodent eye. Vis. Neurosci. 15, 597-605. Balcombe, J., 2005. Pleasure: the neglected experience (Poster presentation). Paper presented at From Darwin to Dawkins: the science and implications of animal sentience, London. Balcombe, J. P., Barnard, N. D., Sandusky, C., 2004. Laboratory routines cause animal stress. Contemp. Top. Lab. Anim. Sci. 43, 42-51. Bartoletti, A., Medini, P., Berardi, N., Maffei, L., 2004. Environmental enrichment prevents effects of dark-rearing in the rat visual cortex. Nat. Neurosci. 7, 215216. Bellhorn, R. W., 1980. Lighting in the animal environment. Lab. Anim. Sci. 30, 440450. Bennett, A. T., Cuthill, I. C., 1994. Ultraviolet vision in birds: what is its function? Vision Res. 34, 1471-1478. Berdoy, M. (2002). The laboratory rat: a natural history. Retrieved 24th Feb 2006, from www.ratlife.org Berridge, K. C., 2000. Measuring hedonic impact in animals and infants: microstructure of affective taste reactivity patterns. Neurosci. Biobehav. Rev. 24, 173-198. Berson, E. L., 2000. Nutrition and retinal degenerations. Int. Ophthalmol. Clin. 40, 93-111. Beynen, A. C., 1992. Communication between rats of experiment-induced stress and its impact on experimental results. Anim. Welf. 1, 153-160. Birrell, J. M., Brown, V. J., 2000. Medial frontal cortex mediates perceptual attentional set shifting in the rat. J. Neurosci. 20, 4320-4324. Björk, E., Nevalainen, T., Hakumäki, M., Voipio, H.-M., 2000. R-weighting provides better estimation for rat hearing sensitivity. Lab. Anim. 34, 136-144. Blanchard, D. C., Griebel, G., Blanchard, R. J., 2003. Conditioning and residual emotionality effects of predator stimuli: Some reflections on stress and emotion. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 27, 1177-1185. Blanchard, R. J., Blanchard, D. C., Takahashi, T., Kelley, M. J., 1977a. Attack and defensive behaviour in the albino rat. Anim. Behav. 25, 622-634. Blanchard, R. J., Takahashi, L. K., Fukunaga, K. K., Blanchard, D. C., 1977b. Functions of the vibrissae in the defensive and aggressive behavior of the rat. Aggress. Behav. 3, 231-240. 39

1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103

Blanchard, R. J., Blanchard, D. C., Agullana, R., Weiss, S. M., 1991. Twenty-two kHz alarm cries to presentation of a predator, by laboratory rats living in visible burrow systems. Physiol. Behav. 50, 967-972. Blanchard, R. J., Nikulina, J. N., Sakai, R. R., McKittrick, C., McEwen, B., Blanchard, D. C., 1998. Behavioral and endocrine change following chronic predatory stress. Physiol. Behav. 63, 561-569. Blom, H. J. M., Van Tintelen, G., Baumans, V., Van Den Broek, J., Beynen, A. C., 1995. Development and application of a preference test system to evaluate housing conditions for laboratory rats. Appl. Anim. Behav. Sci. 43, 279-290. Boakes, R. A., Boot, B., Clarke, J. V., Carver, A., 2000. Comparing albino and hooded Wistar rats of both sexes on a range of behavioral and learning tasks. Psychobiology 28, 339-359. Bourgeon, S., Xerri, C., Coq, J.-O., 2004. Abilities in tactile discrimination of textures in adult rats exposed to enriched or impoverished environments. Behav. Brain Res. 153, 217-231. Brainard, G. C., Barker, F. M., Hoffman, R. J., Stetson, M. H., Hanifin, J. P., Podolin, P. L., Rollag, M. D., 1994. Ultraviolet regulation of neuroendocrine and circadian physiology in rodents. Vision Res. 34, 1521-1533. Brennan, P. A., Keverne, E. B., 2004. Something in the air? New insights into mammalian pheromones. Curr. Biol. 14, R81-R89. Bresnahan, J. F., Kitchell, B. B., Wildman, M. F., 1983. Facial hair barbering in rats. Lab. Anim. Sci. 33, 290-291. Brouettelahlou, I., Amouroux, R., Chastrette, F., Cosnier, J., Stoffelsma, J., Vernetmaury, E., 1991. Dodecyl propionate, attractant from rat pup preputial gland - characterization and identification. J. Chem. Ecol. 17, 1343-1354. Brown, R. E., 1992. Responses of dominant and subordinate male-rats to the odors of male and female conspecifics. Aggress. Behav. 18, 129-138. Bruce, H. M., Parrott, D. M., 1960. Role of olfactory sense in pregnancy block by strange males. Science 131, 1526. Bruce, H. M., 1965. Comments on the design of food hoppers in current use for laboratory animals. J. Anim. Technicians Assoc. 16, 32-33. Brudzynski, S. M., Chiu, E. M., 1995. Behavioural responses of laboratory rats to playback of 22 kHz ultrasonic calls. Physiol. Behav. 57, 1039-1044. Brudzynski, S. M., Pniak, A., 2002. Social contacts and production of 50-kHz short ultrasonic calls in adult rats. J. Comp. Psychol. 116, 73-82. Bulduk, S., Canbeyli, R., 2004. Effect of inescapable tones on behavioral despair in Wistar rats. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 28, 471-475. Bulla, G., 1999. Fancy rats: a complete pet owner's manual (E. A. Bye, Trans.). New York: Barrons Educational Series Inc. Burgdorf, J., Knutson, B., Panksepp, J., 2000. Anticipation of rewarding electrical brain stimulation evokes ultrasonic vocalization in rats. Behav. Neurosci. 114, 320-327. Burgdorf, J., Knutson, B., Panksepp, J., Shippenberg, T. S., 2001. Evaluation of rat ultrasonic vocalizations as predictors of the conditioned aversive effects of drugs. Psychopharmacology 155, 35-42. Burgdorf, J., Panksepp, J., 2001. Tickling induces reward in adolescent rats. Physiol. Behav. 72, 167-173. Burman, O. H. P., Mendl, M., 2003. The influence of preexperimental experience on social discrimination in rats (Rattus norvegicus). J. Comp. Psychol. 117, 344349. 40

1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151

Burman, O. H. P., Ilyat, A., Jones, G., Mendl, M., 2007. Ultrasonic vocalizations as indicators of welfare for laboratory rats (Rattus norvegicus). Appl. Anim. Behav. Sci. 104, 116-129. Burn, C. C., 2006. Effects of husbandry manipulations and the laboratory environment on rat health and welfare. Unpublished D.Phil. thesis, University of Oxford, Oxford, UK. Burn, C. C., Peters, A., Mason, G. J., 2006a. Acute effects of cage cleaning at different frequencies on laboratory rat behaviour and welfare. Anim. Welf. 15, 161-172. Burn, C. C., Day, M. J., Peters, A., Mason, G. J., 2006b. Long-term effects of cagecleaning frequency and bedding type on laboratory rat health, welfare, and handleability: a cross-laboratory study. Lab. Anim. 40, 353-370. Burn, C. C., Deacon, R., Mason, G. J., in press. Marked for life? Effects of early cage cleaning frequency, delivery batch and identification tail-marking on adult rat anxiety profiles. Dev. Psychobiol. Burn, C. C., Mason, G. J., in press. Effects of cage-cleaning frequencies on rat reproductive performance, infanticide, and welfare. Appl. Anim. Behav. Sci. Calhoun, J. B., 1963. The ecology and sociology of the Norway rat (Vol. 1008). Bethesda, Md., USA: U.S. Department of Health Education and Welfare Public Health Service. Calvino, B., Besson, J. M., Boehrer, A., Depaulis, A., 1996. Ultrasonic vocalization (22-28 kHz) in a model of chronic pain, the arthritic rat: Effects of analgesic drugs. Neuroreport 7, 581-584. Campbell, B. A., Messing, R. B., 1969. Aversion thresholds and aversion difference limens for white light in albino and hooded rats. J. Exp. Psychol. 82, 353-359. Cancedda, L., Putignano, E., Sale, A., Viegi, A., Berardi, N., Maffei, L., 2004. Acceleration of visual system development by environmental enrichment. J. Neurosci. 24, 4840-4848. Capaldi, E. D., 1996. Conditioned food preferences. In E. D. Capaldi (Ed.), Why we eat what we eat (pp. 53-82). Washington: American Psychological Association. Cardenas, F., Lamprea, M. R., Morato, S., 2001. Vibrissal sense is not the main sensory modality in rat exploratory behavior in the elevated plus-maze. Behav. Brain Res. 122, 169-174. Carder, B., Berkowitz, K., 1970. Rats' preference for earned in comparison with free food. Science 167, 1273-1274. Carr, W. J., Yee, L., Gable, D., Marasco, E., 1976. Olfactory recognition of conspecifics by domestic norway rats. J. Comp. Physiol. Psychol. 90, 821-828. Carvell, G. E., Simons, D. J., 1990. Biometric analyses of vibrissal tactile discrimination in the rat. J. Neurosci. 10, 2638-2648. Casadesus, G., Shukitt-Hale, B., Stellwagen, H. M., Zhu, X. W., Lee, H. G., Smith, M. A., Joseph, J. A., 2004. Modulation of hippocampal plasticity and cognitive behavior by short-term blueberry supplementation in aged rats. Nutr. Neurosci. 7, 309-316. Ceccarelli, I., Lariviere, W. R., Fiorenzani, P., Sacerdote, P., Aloisi, A. M., 2004. Effects of long-term exposure of lemon essential oil odor on behavioral, hormonal and neuronal parameters in male and female rats. Brain Res. 1001, 78-86.

41

1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201

Chase, J., 1980. Rat echolocation: correlations between object detection and click production. In R. G. Busnel, J. F. Fish (Eds.), Animal Sonar Systems (pp. 875877). New York: Plenum Press. Cicerone, C. M., 1976. Cones survive rods in the light-damaged eye of the albino rat. Science 194, 1183-1185. Clarke, S. N. D. A., Koh, M. T., Bernstein, I. L., 2001. NaCl detection thresholds: comparison of Fischer 344 and Wistar rats. Chem. Senses 26, 253-257. Clenet, F., Bouyon, E., Hascoet, M., Bourin, M., 2006. Light/dark cycle manipulation influences mice behaviour in the elevated plus maze. Behav. Brain Res. 166, 140-149. Clifton, P. G., Burton, M. J., Sharp, C., 1987. Rapid loss of stimulus-specific satiety after consumption of a second food. Appetite 9, 149-156. Collerain, I., Ludvigson, H. W., 1972. Aversion of conspecific odor of frustrative nonreward in rats. Psychon. Sci. 27, 54-56. Commission of the European Communities. 2003. Third report from the Commission to the Council and the European Parliament on the statistics on the number of animals used for experimental and other scientific purposes in the member states of the European Union (No. COM/2003/0019). Brussels: Commission of the European Communities. Coq, J. O., Xerri, C., 1998. Environmental enrichment alters organizational features of the forepaw representation in the primary somatosensory cortex of adult rats. Exp. Brain Res. 121, 191-204. Coq, J. O., Xerri, C., 2001. Sensorimotor experience modulates age-dependent alterations of the forepaw representation in the rat primary somatosensory cortex. Neuroscience 104, 705-715. Creel, D. J., Dustman, R. E., Beck, E. C., 1970. Differences in visually evoked responses in albino versus hooded rats. Exp. Neurol. 29, 246-260. Cummings, D. M., Henning, H. E., Brunjes, P. C., 1997. Olfactory bulb recovery after early sensory deprivation. J. Neurosci. 17, 7433-7440. D'Eath, R. B., 1998. Can video images imitate real stimuli in animal behaviour experiments? Biol. Revs 73, 267-292. Dagg, A. I., Bell, W. L., Windsor, D. E., 1971. Urine marking of cages and visual isolation as possible sources of error in behavioural studies of small mammals. Lab. Anim. 5, 163-167. Datesand Ltd. 2005. Environmental enrichment. In Product Catalogue (pp. 23-26). Manchester: Datesand Ltd. Davis, H., MacKenzie, K. A., Morrison, S., 1989. Numerical discrimination by rats (Rattus norvegicus) using body and vibrissal touch. J. Comp. Psychol. 103, 45-53. Davis, H., Taylor, A. A., Norris, C., 1997. Preference for familiar humans by rats. Psychon. Bull. Rev. 4, 118-120. Day, D. E., Mintz, E. M., Bartness, T. J., 2002. Diet choice exaggerates food hoarding, intake and pup survival across reproduction. Physiol. Behav. 75, 143-157. Day, H. D., Seay, B. M., Hale, P., Hendricks, D., 1982. Early social deprivation and the ontogeny of unrestricted social behavior in the laboratory rat. Dev. Psychobiol. 15, 47-59. de Almeida, R. N., Motta, S. C., de Brito Faturi, C., Catallani, B., Leite, J. R., 2004. Anxiolytic-like effects of rose oil inhalation on the elevated plus-maze test in rats. Pharmacol. Biochem. Behav. 77, 361-364. 42

1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241 1242 1243 1244 1245 1246 1247 1248 1249

Dess, N. K., 1993. Saccharin's aversive taste in rats - evidence and implications. Neurosci. Biobehav. Rev. 17, 359-372. Deutschlander, M. E., Freake, M. J., Borland, S. C., Phillips, J. B., Madden, R. C., Anderson, L. E., Wilson, B. W., 2003. Learned magnetic compass orientation by the Siberian hamster, Phodopus sungorus. Anim. Behav. 65, 779-786. DiBattista, D., 1990. Conditioned taste avoidance induced by lactose ingestion in adult rats. Physiol. Behav. 47, 253-257. Dieter, M. P., Goehl, T. J., Jameson, C. W., Elwell, M. R., Hildebrandt, P. K., Yuan, J. H., 1993. Comparison of the toxicity of citral in F344 rats and B6C3F1 mice when administrated by microencapsulation in feed or by corn-oil gavage. Food Chem. Toxicol. 31, 463-474. Dilsaver, S. C., Majchrzak, M. J., 1988. Bright artificial light produces subsensitivity to nicotine. Life Sci. 42, 225-230. Doty, R. L., 1986. Odor-guided behavior in mammals. Experientia 42, 257-271. Dulac, C., 1997. Molecular biology of pheromone perception in mammals. Semin. Cell Dev. Biol. 8, 197-205. Dymond, S., 1995. Conditional discrimination responding in non-humans. Ir. J. Psych. 16, 334-345. Dymond, S., Gomez-Martin, S., Barnes, D., 1996. Multi-modal conditional discrimination in rats: Some preliminary findings. Ir. J. Psych. 17, 269-281. Ellis, M. E., 1984. Exhaustive memory scanning in Rattus norvegicus: recognition for food items. J. Comp. Psychol. 98, 194-200. Emes, R. D., Beatson, S. A., Ponting, C. P., Goodstadt, L., 2004. Evolution and comparative genomics of odorant- and pheromone-associated genes in rodents. Genome Res. 14, 591-602. Engineer, N. D., Percaccio, C. R., Pandya, P. K., Moucha, R., Rathbun, D. L., Kilgard, M. P., 2004. Environmental enrichment improves response strength, threshold, selectivity, and latency of auditory cortex neurons. J. Neurophysiol. 92, 73-82. European Union, 2003. Directive 2003/10/EC on the minimum health and safety requirements regarding the exposure of workers to the risks arising from physical agents (noise), The European Parliament and the Council of the European Union Fagiolini, M., Pizzorusso, T., Berardi, N., Domenici, L., Maffei, L., 1994. Functional postnatal development of the rat primary visual cortex and the role of visual experience: dark rearing and monocular deprivation. Vision Res. 34, 709-720. Farrell, W. J., Alberts, J. R., 2002. Stimulus control of maternal responsiveness to Norway rat (Rattus norvegicus) pup ultrasonic vocalizations. J. Comp. Psychol. 116, 297-307. Ferreira, A., Hansen, S., 1986. Sensory control of maternal aggression in Rattus norvegicus. J. Comp. Psychol. 100, 173-177. Festing, M. F., May, D., Connors, T. A., Lovell, D., Sparrow, S., 1978. An athymic nude mutation in the rat. Nature 274, 365-366. Fleshler, M., 1965. Adequate acoustic stimulus for startle reaction in the rat. J. Comp. Physiol. Psychol. 60, 200-207. Forsman, K. A., Malmquist, M. G., 1988. Evidence for echolocation in the common shrew. J. Zool. 216, 655-662. Fox, S., 1997. The guide to owning a rat. Neptune City: TFH Publications Inc.

43

1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298

Galef, B. G., Jr., Heiber, L., 1976. Role of residual olfactory cues in the determination of feeding site selection and exploration patterns of domestic rats. J. Comp. Physiol. Psychol. 90, 727-739. Galef, B. G., Jr., Mason, J. R., Preti, G., Bean, N. J., 1988. Carbon disulfide: A semiochemical mediating socially-induced diet choice in rats. Physiol. Behav. 42, 119-124. Galef, B. G., Jr., McQuoid, L. M., Whiskin, E. E., 1990. Further evidence that Norway rats do not socially transmit learned aversions to toxic baits. Anim. Learn. Behav. 18, 199-205. Galef, B. G., Jr., Buckley, L. L., 1996. Use of foraging trails by Norway rats. Anim. Behav. 51, 765-771. Galef, B. G., Jr., Whiskin, E. E., 2003a. Preference for novel flavors in adult Norway rats (Rattus norvegicus). J. Comp. Psychol. 117, 96-100. Galef, B. G., Jr., Whiskin, E. E., 2003b. Socially transmitted food preferences can be used to study long-term memory in rats. Learn. Behav. 31, 160-164. Galef, B. G., Jr., Whiskin, E. E., 2005. Differences between golden hamsters (Mesocricetus auratus) and Norway rats (Rattus norvegicus) in preference for the sole diet that they are eating. J. Comp. Psychol. 119, 8-13. Gamble, M. R., Clough, G., 1976. Ammonia build-up in animal boxes and its effect on rat tracheal epithelium. Lab. Anim. 10, 93-104. Gamble, M. R., 1982. Noise and laboratory animals. J. Inst. Anim. Technicians 33, 515. Garcia-Brull, P. D., Nunez, J., Nunez, A., 1993. The effect of scents on the territorial and aggressive-behavior of laboratory rats. Behav. Process. 29, 25-36. Garcia, A. M. B., Cardenas, F. P., Morato, S., 2005. Effect of different illumination levels on rat behavior in the elevated plus-maze. Physiol. Behav. 85, 265-270. Garner, J. P., Dufour, B., Gregg, L. E., Weisker, S. M., Mench, J. A., 2004. Social and husbandry factors affecting the prevalence and severity of barbering ('whisker trimming') by laboratory mice. Appl. Anim. Behav. Sci. 89, 263-282. Gawienowski, A. M., Orsulak, P. J., Stacewicz-Sapuntzakis, M., Joseph, B. M., 1975. Presence of sex pheromone in preputial glands of male rats. J. Endocrinol. 67, 283-288. Gemberling, G. A., 1984. Ingestion of a novel flavor before exposure to pups injected with lithium chloride produces a taste aversion in mother rats (Rattus norvegicus). J. Comp. Psychol. 98, 285-301. Gheusi, G., Goodall, G., Dantzer, R., 1997. Individually distinctive odours represent individual conspecifics in rats. Anim. Behav. 53, 935-944. Gilad, Y., Man, O., Paabo, S., Lancet, D., 2003. Human specific loss of olfactory receptor genes. Proc. Natl. Acad. Sci. USA 100, 3324-3327. Giroux, M. L., Malatynska, E., Dilsaver, S. C., 1991. Bright light does not alter muscarinic receptor binding parameters. Pharmacol. Biochem. Behav. 38, 695697. Glendinning, J. I., 1994. Is the bitter rejection response always adaptive? Physiol. Behav. 56, 1217-1227. Godde, B., Berkefeld, T., David-Jurgens, M., Dinse, H. R., 2002. Age-related changes in primary somatosensory cortex of rats: evidence for parallel degenerative and plastic-adaptive processes. Neurosci. Biobehav. Rev. 26, 743-752. Graver, L., Mason, G. J., Burman, O. H. P., 2004. Do rats use ultrasound? Unpublished B.A. Honours dissertation, University of Oxford, Oxford.

44

1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332 1333 1334 1335 1336 1337 1338 1339 1340 1341 1342 1343 1344 1345 1346 1347

Gray, S., Hurst, J. L., 1995. The effect of cage cleaning on aggression within groups of male laboratory mice. Anim. Behav. 49, 821-826. Grill, H. J., Norgren, R., 1978. The taste reactivity test: I. Mimetic responses to gustatory stimuli in neurologically normal rats. Brain Res. 143, 263-279. Guic-Robles, E., Valdivieso, C., Guajardo, G., 1989. Rats can learn a roughness discrimination using only their vibrissal system. Behav. Brain. Res. 31, 285289. Gutiérrez-García, A. G., Contreras, C. M., Mendoza-López, M. R., Cruz-Sánchez, S., García-Barradas, O., Rodríguez-Landa, J. F., Bernal-Morales, B., 2006. A single session of emotional stress produces anxiety in Wistar rats. Behav. Brain. Res. 167, 30-35. Hansen, S. R., Janssen, C., Beasley, V. R., 1993. Denatonium benzoate as a deterrent to ingestion of toxic substances: toxicity and efficacy. Vet. Hum. Toxicol. 35, 234-236. Hawkins, P., Grant, G., Raymond, R., Hughes, G., Morton, D., Mason, G. J., Playle, L., Hubrecht, R., Jennings, M., 2004. Reducing suffering through refinement of procedures: Report of the 2003 RSPCA/UFAW rodent welfare group meeting. Anim. Technol. Welf. 3, 79-85. Hawkins, P., Nicholson, J., Burn, C. C., Mean, J., Leach, M., Strudley, I., Van Loo, P., Bolam, S., Anderson, D., Hubrecht, R., Jennings, M., 2005. Report of the 2004 RSPCA/UFAW Rodent Welfare Group meeting. Anim. Technol. 4, 7989. Heale, V. R., Vanderwolf, C. H., Kavaliers, M., 1994. Components of weasel and fox odors elicit fast wave bursts in the dentate gyrus of rats. Behav. Brain Res. 63, 159-165. Heffner, R. S., Heffner, H. E., 1992a. Visual factors in sound localization in mammals. J. Comp. Neurol. 317, 219-232. Heffner, H. E., Heffner, R. S., 1992b. Auditory Perception. In P. Clive, D. Piggins (Eds.), Farm animals and the environment (pp. 159-184). Wallingford, UK: CAB International. Heffner, H. E., Heffner, R. S., Contos, C., Ott, T., 1994. Audiogram of the hooded Norway rat. Hear. Res. 73, 244-247. Hemsworth, P. H., 2003. Human-animal interactions in livestock production. Appl. Anim. Behav. Sci. 81, 185-198. Hermer-Vazquez, L., Hermer-Vazquez, R., Chapin, J. K., 2005. Somatosensation. In I. Q. Whishaw, B. Kolb (Eds.), The behavior of the laboratory rat: a handbook with tests (pp. 60-68). Oxford: Oxford University Press. Heuschele, W. P., 1995. Use of disinfectants in zoos and game parks. Revue Scientifique et Technique 14, 447-454. Hirsjarvi, P., Valiaho, T., 1995. Effects of gentling on open-field behaviour of Wistar rats in fear-evoking test situation. Lab. Anim. 29, 380-384. Hobson, J. A., 2005. Sleep is of the brain, by the brain and for the brain. Nature 437, 1254-1256. Holy, T. E., Dulac, C., Meister, M., 2000. Responses of vomeronasal neurons to natural stimuli. Science 289, 1569-1572. Home Office, 1989. Code of practice for the housing and care of animals used in scientific procedures, Home Office, UK Hopp, S. L., Owren, M. J., Marion, J. R., 1985. Olfactory discrimination of individual littermates in rats (Rattus norvegicus). J. Comp. Psychol. 99, 248-251.

45

1348 1349 1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364 1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 1379 1380 1381 1382 1383 1384 1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397

Huang-Brown, K. M., Guhad, F. A., 2002. Chocolate, an effective means of oral drug delivery in rats. Lab Animal 31, 34-36. Hughes, C. W., Stein, E. A., Lynch, J. J., 1978. Hopelessness-induced sudden death in rats: Anthropomorphism for experimentally induced drownings? J. Nerv. Ment. Dis. 166, 387-401. Hugo, J., Krijt, J., Vokurka, M., Janousek, V., 1987. Secretory response to light in rat Harderian gland: possible photoprotective role of Harderian porphyrin. Gen. Physiol. Biophys. 6, 401-404. Humpel, C., Neudorfer, C., Philipp, W., Steiner, H. J., Haring, C., Schmid, K. W., Schwitzer, J., Saria, A., 1992. Effects of bright artificial-light on monoamines and neuropeptides in 8 different brain-regions compared in a pigmented and nonpigmented rat strain. J. Neurosci. Res. 32, 605-612. Hurst, J. L., Barnard, C. J., Nevison, C. M., West, C. D., 1997. Housing and welfare in laboratory rats: Welfare implications of isolation and social contact among caged males. Anim. Welf. 6, 329-347. Hurst, J. L., Thom, M. D., Nevison, C. M., Humphries, R. E., Beynon, R. J., 2005. MHC odours are not required or sufficient for recognition of individual scent owners. Proc. R. Soc. Lond. B 272, 715-724. Hyman, S., Rawson, N. E., 2001. Preliminary results of olfactory testing in rats without deprivation. Lab Animal 30, 38-39. Inglis, C. A., Campbell, E. R., Auciello, S. L., Sarawar, S. R., 2004. Effects of enrichment devices on stress-related problems in mouse breeding: Final report for The Centre for Alternatives to Animal Testing, John Hopkins University. Jacobs, G. H., Fenwick, J. A., Williams, G. A., 2001. Cone-based vision of rats for ultraviolet and visible lights. J. Exp. Biol. 204, 2439-2446. Jarvis, J. R., Prescott, N. B., Wathes, C. M., 2003. A mechanistic inter-species comparison of flicker sensitivity. Vision Res. 43, 1723-1734. Jensen, P. G., Curtis, P. D., Dunn, J. A., Austic, R. E., Richmond, M. E., 2003. Field evaluation of capsaicin as a rodent aversion agent for poultry feed. Pest Manag. Sci. 59, 1007-1015. Johnson, S. R., Patterson Kane, E. G., 2003. Foraging as environmental enrichment for laboratory rats: a theoretical review. Anim. Technol. Welf. 2, 13-22. Jourdan, D., Ardid, D., Chapuy, E., Eschalier, A., Lebars, D., 1995. Audible and ultrasonic vocalization elicited by single electrical nociceptive stimuli to the tail in the rat. Pain 63, 237-249. Kaltwasser, M. T., Schnitzler, H. U., 1981. Echolocation signals confirmed in rats. Zeit. Saugetierkunde 46, 394-395. Kaltwasser, M. T., 1990. Startle-inducing acoustic stimuli evoke ultrasonic vocalization in the rat. Physiol. Behav. 48, 13-17. Kaltwasser, M. T., 1991. Acoustic startle induced ultrasonic vocalization in the rat - a novel animal-model of anxiety. Behav. Brain Res. 43, 133-137. Keeler, C. E., 1942. The association of the black (non-agouti) gene with behavior in the Norway rat. J. Hered. 33, 371-384. Keller, C., Grimm, C., Wenzel, A., Hafezi, F., Reme, C. E., 2001. Protective effect of halothane anesthesia on retinal light damage: Inhibition of metabolic rhodopsin regeneration. Invest. Ophth. Vis. Sci. 42, 476-480. Kelly, J. B., Masterton, B., 1977. Auditory sensitivity of the albino rat. J. Comp. Physiol. Psychol. 91, 930-936. Kemble, E. D., Garbe, C. M., Gordon, C., 1995. Effects of novel odors on intermale attack behavior in mice. Aggress. Behav. 21, 293-299. 46

1398 1399 1400 1401 1402 1403 1404 1405 1406 1407 1408 1409 1410 1411 1412 1413 1414 1415 1416 1417 1418 1419 1420 1421 1422 1423 1424 1425 1426 1427 1428 1429 1430 1431 1432 1433 1434 1435 1436 1437 1438 1439 1440 1441 1442 1443 1444 1445 1446

Key, D., 2004. Environmental enrichment options for laboratory rats and mice. Lab Anim. Eur. 4, 30-38. Kikusui, T., Takigami, S., Takeuchi, Y., Mori, Y., 2001. Alarm pheromone enhances stress-induced hypothermia in rats. Physiol. Behav. 72, 45-50. Kiyokawa, Y., Kikusui, T., Takeuchi, Y., Mori, Y., 2004. Modulatory role of testosterone in alarm pheromone release by male rats. Horm. Behav. 45, 122127. Knutson, B., Burgdorf, J., Panksepp, J., 1998. Anticipation of play elicits highfrequency ultrasonic vocalizations in young rats. J. Comp. Psychol. 112, 6573. Knutson, B., Burgdorf, J., Panksepp, J., 1999. High-frequency ultrasonic vocalizations index conditioned pharmacological reward in rats. Physiol. Behav. 66, 639643. Knutson, B., Burgdorf, J., Panksepp, J., 2002. Ultrasonic vocalizations as indices of affective states in rats. Psychol. Bull. 128, 961-977. Komori, T., Miyahara, S., Yamamoto, M., Matsumoto, T., Zhang, K., Nakagawa, M., Nomura, S., Motomura, E., Shiroyama, T., Okazaki, Y., 2003. Effects of odorants on the hypothalamic-pituitary-adrenal axis and interleukin-6 (IL-6) and IL-6 receptor mRNA expression in rat hypothalamus after restraint stress. Chem. Senses 28, 767-771. Krames, L., Carr, W. J., Bergman, B., 1969. A pheromone associated with social dominance among male rats. Psychon. Sci. 16, 11-12. Krinke, G., 2000. The laboratory rat. London: Academic Press. La Vail, M. M., 1976. Survival of some photoreceptor cells in albino rats following long-term exposure to continuous light. Inv. Opthal. 15, 64-70. Lacey, J. C., Beynon, R. J., Hurst, J. L., 2007. The importance of exposure to other male scents in determining competitive behaviour among inbred male mice. Appl. Anim. Behav. Sci. 104, 130-142. Lane-Petter, W., 1975. Presentation of compound diets. Anim. Technol. 26, 83-85. Larue-Achagiotis, C., Martin, C., Verger, P., Louis-Sylvestre, J., 1992. Dietary selfselection vs. complete diet: body weight gain and meal pattern in rats. Physiol. Behav. 51, 995-999. Lashley, K. S., 1938. The mechanisms of vision III: The comparative visual acuity of pigmented and albino rats. J. Genet. Psych. 37, 481-484. Latham, N., Mason, G., 2004. From house mouse to mouse house: the behavioural biology of free-living Mus musculus and its implications in the laboratory. Appl. Anim. Behav. Sci. 86, 261-289. Lavin, M. J., Freise, B., Coombes, S., 1980. Transferred flavor aversions in adult rats. Behav. Neur. Biol. 28, 15-33. Le Magnen, J., 1999a. Efficacy of olfactory, tactile, and other food stimuli in the acquisition and manifestation of appetite in rats. Appetite 33, 43-51. Le Magnen, J., 1999b. Increased food intake induced in rats by changes in the satiating sensory input from food. Appetite 33, 33-35. Legg, C. R., Lambert, S., 1990. Distance estimation in the hooded rat: Experimental evidence for the role of motion cues. Behav. Brain Res. 41, 11-20. Lemmon, V., Anderson, K. V., 1979. Behavioral correlates of constant light-induced retinal degeneration. Exp. Neurol. 63, 35-49. Levine, S., Saltzman, A., 1998. An alternative to overnight withholding of food from rats. Contemp. Top. Lab. Anim. Sci. 37.

47

1447 1448 1449 1450 1451 1452 1453 1454 1455 1456 1457 1458 1459 1460 1461 1462 1463 1464 1465 1466 1467 1468 1469 1470 1471 1472 1473 1474 1475 1476 1477 1478 1479 1480 1481 1482 1483 1484 1485 1486 1487 1488 1489 1490 1491 1492 1493 1494

Li, Z. Y., Tso, M. O., Wang, H. M., Organisciak, D. T., 1985. Amelioration of photic injury in rat retina by ascorbic acid: a histopathologic study. Invest. Ophth. Vis. Sci. 26, 1589-1598. Light Therapy ProductsTM. (2006). MI, USA. Retrieved 5th February 2006, from http://www.lighttherapyproducts.com/ Lindner, M. D., Plone, M. A., Schaller, T., Emerich, D. F., 1997. Blind rats are not profoundly impaired in the reference memory Morris water maze and cannot be clearly discriminated from rats with cognitive deficits in the cued platform task. Cogn. Brain Res. 5, 329-333. Ludvigson, H. W., Mathis, D. A., Choquette, K. A., 1985. Different odors in rats from large and small rewards. Anim. Learn. Behav. 13, 315-320. Mackay-Sim, A., Laing, D. G., 1980. Discrimination of odors from stressed rats by non-stressed rats. Physiol. Behav. 24, 699-704. Mackay-Sim, A., Laing, D. G., 1981. The sources of odors from stressed rats. Physiol. Behav. 27, 511-513. Maddocks, S. A., Cuthill, I. C., Goldsmith, A. R., Sherwin, C. M., 2001. Behavioural and physiological effects of absence of ultraviolet wavelengths for domestic chicks. Anim. Behav. 62, 1013-1019. Manzo, J., Garcia, L. I., Hernandez, M. E., Carrillo, P., Pacheco, P., 2002. Neuroendocrine control of urine-marking behavior in male rats. Physiol. Behav. 75, 25-32. Mason, G., Wilson, D., Hampton, C., Wurbel, H., 2004. Non-invasively assessing disturbance and stress in laboratory rats by scoring chromodacryorrhoea. Alt. Lab. Anim. 32, 153-159. Mason, R. K., 1969. The influence of noise on emotional states. J. Psychosom. Res. 13, 275-282. Mattson, M. P., 2005. Energy intake, meal frequency, and health: a neurobiological perspective. Annu. Rev. Nutr. 25, 237-260. McCall, R. B., Lester, M. L., Corter, C. M., 1969. Caretaker effect in rats. Dev. Psychol. 1, 771. McLennan, I. S., Taylor-Jeffs, J., 2004. The use of sodium lamps to brightly illuminate mouse houses during their dark phases. Lab. Anim. 38, 384-392. Melnick, R. L., Jameson, C. W., Goehl, T. J., Maronpot, R. R., Collins, B. J., Greenwell, A., Harrington, F. W., Wilson, R. E., Tomaszewski, K. E., Agarwal, D. K., 1987. Application of microencapsulation for toxicology studies. II. Toxicity of microencapsulated trichloroethylene in Fischer 344 rats. Fundam. Appl. Toxicol. 8, 432-442. Mennella, J. A., Moltz, H., 1988. Infanticide in rats: male strategy and female counter-strategy. Physiol. Behav. 42, 19-28. Meyer, M. E., Meyer, M. E., 1992. The effects of bilateral and unilateral vibrissotomy on behavior within aquatic and terrestrial environments. Physiol. Behav. 51, 877-880. Milligan, S. R., Sales, G. D., Khirnykh, K., 1993. Sound levels in rooms housing laboratory-animals - an uncontrolled daily variable. Physiol. Behav. 53, 10671076. Mitchell, C. J., Heyes, C. M., Gardner, M. R., Dawson, G. R., 1999. Limitations of a bidirectional control procedure for the investigation of imitation in rats: Odour cues on the manipulandum. Q. J. Exp. Psychol. B 52, 193-202.

48

1495 1496 1497 1498 1499 1500 1501 1502 1503 1504 1505 1506 1507 1508 1509 1510 1511 1512 1513 1514 1515 1516 1517 1518 1519 1520 1521 1522 1523 1524 1525 1526 1527 1528 1529 1530 1531 1532 1533 1534 1535 1536 1537 1538 1539 1540 1541 1542 1543 1544

Moemeka, A. N., Hildebrandt, A. L., Radaskiewicz, P., King, T. R., 1998. Shorn (shn): a new mutation causing hypotrichosis in the Norway rat. J. Hered. 89, 257-260. Moinard, C., Sherwin, C. M., 1999. Turkeys prefer fluorescent light with supplementary ultraviolet radiation. Appl. Anim. Behav. Sci. 64, 261-267. Moore, B. C. J., 2003. An introduction to the psychology of hearing (5th ed.). London: Academic Press. Moore, C. L., 1985. Sex differences in urinary odors produced by young laboratory rats (Rattus norvegicus). J. Comp. Psychol. 99, 336-341. Morlock, G. W., McCormick, C. E., Meyer, M. E., 1971. The effect of a stranger's presence on the exploratory behavior of rats. Psychon. Sci. 22, 3-4. Mueller, K. L., Hoon, M. A., Erlenbach, I., Chandrashekar, J., Zuker, C. S., Ryba, N. J. P., 2005. The receptors and coding logic for bitter taste. Nature 434, 225229. Muenzinger, K. F., Reynolds, H. E., 1936. Color vision in white rats: I. Sensitivity to red. J. Genet. Psych. 48, 58-71. Muheim, R., Edgar, N. M., Sloan, K. A., Phillips, J. B., 2006. Magnetic compass orientation in C57BL/6J mice. Learn. Behav. 34, 366-373. Munn, N. L., Collins, M., 1936. Discrimination of red by white rats. J. Genet. Psych. 48, 72-87. Nagel, T., 1974. What is it like to be a bat? Philos. Rev. 83, 435-450. Naim, M., Brand, J. G., Kare, M. R., Carpenter, R. G., 1985. Energy intake, weight gain, and fat deposition in rats fed flavored, nutritionally controlled diets in a multichoice ("cafeteria") design. J. Nutr. 115, 1447-1458. Nakashima, T., Akamatsu, M., Hatanaka, A., Kiyohara, T., 2004. Attenuation of stress-induced elevations in plasma ACTH level and body temperature in rats by green odor. Physiol. Behav. 80, 481-488. Neimark, M. A., Andermann, M. L., Hopfield, J. J., Moore, C. I., 2003. Vibrissa resonance as a transduction mechanism for tactile encoding. J. Neurosci. 23, 6499-6509. Nevison, C. M., Barnard, C. J., Hurst, J. L., 2003. The consequence of inbreeding for modulating social relationships between competitors. Appl. Anim. Behav. Sci. 81, 387-398. Noirot, E., 1968. Ultrasounds in young rodents. II. Changes with age in albino rats. Anim. Behav. 16, 129-134. O'Sullivan, D. J., Spear, N. E., 1964. Comparison of hooded and albino rats on the visual cliff. Psychon. Sci. 1, 87-88. Olsson, I. A. S., Nevison, C. M., Patterson-Kane, E. G., Sherwin, C. M., Van de Weerd, H. A., Wurbel, H., 2003. Understanding behaviour: the relevance of ethological approaches in laboratory animal science. Appl. Anim. Behav. Sci. 81, 245-264. Osteen, W. K., Spencer, R. L., Bare, D. J., Mcewen, B. S., 1995. Analysis of severe photoreceptor loss and Morris water-maze performance in aged rats. Behav. Brain Res. 68, 151-158. Otten, W., Kanitz, A. E., Puppe, B., Tuchscherer, M., Brüssow, K. P., Nürnberg, G., Stabenow, B., 2004. Acute and long term effects of chronic intermittent noise stress on hypothalamic-pituitary-adrenocortical and sympathoadrenomedullary axis in pigs. Anim. Sci. 78, 271. Otto, J., Brown, M. F., Long, W., 2002. Training rats to search and alert on contraband odors. Appl. Anim. Behav. Sci. 77, 217-232. 49

1545 1546 1547 1548 1549 1550 1551 1552 1553 1554 1555 1556 1557 1558 1559 1560 1561 1562 1563 1564 1565 1566 1567 1568 1569 1570 1571 1572 1573 1574 1575 1576 1577 1578 1579 1580 1581 1582 1583 1584 1585 1586 1587 1588 1589 1590 1591 1592 1593

Outside In Ltd. (2006). Cambridge, UK. Retrieved 5th February 2006, from http://www.outsidein.co.uk/ Overstreet, D. H., Pucilowski, O., Rezvani, A. H., Janowsky, D. S., 1995. Administration of antidepressants, diazepam and psychomotor stimulants further confirms the utility of Flinders sensitive line rats as an animal-model of depression. Psychopharmacology 121, 27-37. Panksepp, J., Burgdorf, J., 2000. 50-kHz chirping (laughter?) in response to conditioned and unconditioned tickle-induced reward in rats: Effects of social housing and genetic variables. Behav. Brain Res. 115, 25-38. Panksepp, J., 2006. Affective neuroscience and the ancestral sources of socioemotional feelings within animal minds. Paper presented at the 40th International Congress of the ISAE, Bristol. Pappas, B. A., Vickers, G., Buxton, M., Pusztay, W., 1982. Infant rat hyperactivity elicited by home cage bedding is unaffected by neonatal telencephalic dopamine or norepinephrine depletion. Pharmacol. Biochem. Behav. 16, 151154. Perez, J., Perentes, E., 1994. Light-induced retinopathy in the albino rat in long-term studies. An immunohistochemical and quantitative approach. Exp. Toxicol. Pathol. 46, 229-235. Petounis, A., Spyrakis, C., Varonos, D., 1977. Effects of infrasound on activity levels of rats. Physiol. Behav. 18, 153-155. Picciotto, M. R., Self;, D. W., Pohorecky;, L. A., Dawson, G. R., Flint, J., Wilkinson;, L. S., Hen;, R., Tordoff, M. G., Bachmanov, A. A., Friedman, M. I., Beauchamp;, G. K., Wahlsten, D., Crabbe, J., Dudek, B., 1999. Testing the genetics of behavior in mice. Science 285, 2067. Polley, D. B., Kvasnak, E., Frostig, R. D., 2004. Naturalistic experience transforms sensory maps in the adult cortex of caged animals. Nature 429, 67-71. Pons, F., 2006. The effects of distributional learning on rats' sensitivity to phonetic information. J. Exp. Psychol. Anim. Behav. Process. 32, 97-101. Porter, R. H., Winberg, J., 1999. Unique salience of maternal breast odors for newborn infants. Neurosci. Biobehav. Rev. 23, 439-449. Prats, E., Monfar, M., Castella, J., Iglesias, R., Alemany, M., 1989. Energy intake of rats fed a cafeteria diet. Physiol. Behav. 45, 263-272. Prior, H., 2006. Effects of the acoustic environment on learning in rats. Physiol. Behav. 87, 162-165. Prusky, G. T., Reidel, C., Douglas, R. M., 2000. Environmental enrichment from birth enhances visual acuity but not place learning in mice. Behav. Brain Res. 114, 11-15. Prusky, G. T., Harker, K. T., Douglas, R. M., Whishaw, I. Q., 2002. Variation in visual acuity within pigmented, and between pigmented and albino rat strains. Behav. Brain Res. 136, 339-348. Prusky, G. T., Douglas, R. M., 2005. Vision. In I. Q. Whishaw, B. Kolb (Eds.), The behavior of the laboratory rat: a handbook with tests (pp. 49-59). Oxford: Oxford University Press. Quignon, P., Giraud, M., Rimbault, M., Lavigne, P., Tacher, S., Morin, E., Retout, E., Valin, A. S., Lindblad Toh, K., Nicolas, J., Galibert, F., 2005. The dog and rat olfactory receptor repertoires. Genome Biol. 6, R83. Rajan, R., Clement, J. P., Bhalla, U. S., 2006. Rats smell in stereo. Science 311, 666670.

50

1594 1595 1596 1597 1598 1599 1600 1601 1602 1603 1604 1605 1606 1607 1608 1609 1610 1611 1612 1613 1614 1615 1616 1617 1618 1619 1620 1621 1622 1623 1624 1625 1626 1627 1628 1629 1630 1631 1632 1633 1634 1635 1636 1637 1638 1639 1640 1641 1642 1643

Rao, G. N., 1991. Light intensity-associated eye lesions of Fischer-344 rats in longterm studies. Toxicol. Pathol. 19, 148-155. Rat Genome Sequencing Project Consortium. 2004. Genome sequence of the Brown Norway rat yields insights into mammalian evolution. Nature 428, 493-521. Rema, V., Armstrong-James, M., Ebner, F. F., 2003. Experience-dependent plasticity is impaired in adult rat barrel cortex after whiskers are unused in early postnatal life. J. Neurosci. 23, 358-366. Riley, D. A., Rosenzweig, M., 1957. Echolocation in rats. J. Comp. Physiol. Psychol. 50, 323-328. Robbins, K. (2004). Critter critiques - Bio-Serv’s Mouse Igloo and Crawl Ball. AFRMA Rat & Mouse Tales Magazine, 2004, http://www.afrma.org/cc_bioserv.htm. Robinson, L., Bridge, H., Riedel, G., 2001. Visual discrimination learning in the water maze: a novel test for visual acuity. Behav. Brain. Res. 119, 77-84. Rosenzweig, M. R., Riley, D. A., Krech, D., 1955. Evidence for echolocation in the rat. Science 121, 600. Routtenberg, A., Glickman, S. E., 1964. Visual cliff behavior in albino and hooded rats. J. Comp. Physiol. Psychol. 58, 140-142. Sales, G. D., 1972. Ultrasound and aggressive behaviour in rats and other small mammals. Anim. Behav. 20, 88-100. Sales, G. D., Wilson, K. J., Spencer, K. E. V., Milligan, S. R., 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. Sales, G. D., 1991. The effect of 22 kHz calls and artificial 38 kHz signals on activity in rats. Behav. Process. 24, 83-93. Sales, G. D., Milligan, S. R., Khirnykh, K., 1999. Sources of sound in the laboratory animal environment: A survey of the sounds produced by procedures and equipment. Anim. Welf. 8, 97-115. Saper, C. B., Scammell, T. E., Lu, J., 2005. Hypothalamic regulation of sleep and circadian rhythms. Nature 437, 1257-1263. Schellinck, H. M., Brown, R. E., Slotnick, B. M., 1991. Training rats to discriminate between the odors of individual conspecifics. Anim. Learn. Behav. 19, 223233. Schellinck, H. M., Brown, R. E., 2000. Selective depletion of bacteria alters but does not eliminate odors of individuality in Rattus norvegicus. Physiol. Behav. 70, 261-270. Schlingmann, F., De Rijk, S. H. L. M., Pereboom, W. J., Remie, R., 1993a. Light intensity in animal rooms and cages in relation to the care and management of albino rats. Anim. Technol. 44, 97-107. Schlingmann, F., Pereboom, W. J., Remie, R., 1993b. The sensitivity of albino and pigmented rats to light: a mini review. Anim. Technol. 44, 71-85. Schlingmann, F., De Rijk, S. H. L. M., Pereboom, W. J., Remie, R., 1993c. 'Avoidance' as a behavioural parameter in the determination of distress amongst albino and pigmented rats at various light intensities. Anim. Technol. 44, 87-96. Schnecko, A., Witte, K., Lemmer, B., 1998. Effects of routine procedures on cardiovascular parameters of Sprague-Dawley rats in periods of activity and rest. J. Exp. Anim. Sci. 38, 181-190. Sclafani, A., Abrams, M., 1986. Rats show only a weak preference for the artificial sweetener aspartame. Physiol. Behav. 37, 253-256. 51

1644 1645 1646 1647 1648 1649 1650 1651 1652 1653 1654 1655 1656 1657 1658 1659 1660 1661 1662 1663 1664 1665 1666 1667 1668 1669 1670 1671 1672 1673 1674 1675 1676 1677 1678 1679 1680 1681 1682 1683 1684 1685 1686 1687 1688 1689 1690 1691 1692 1693

Sclafani, A., 1987. Carbohydrate taste, appetite, and obesity: An overview. Neurosci. Biobehav. Rev. 11, 131-153. Sclafani, A., Clare, R. A., 2004. Female rats show a bimodal preference response to the artificial sweetener sucralose. Chem. Senses 29, 523-528. Semple-Rowland, S. L., Dawson, W. W., 1987. Retinal cyclic light damage threshold for albino rats. Lab. Anim. Sci. 37, 289-298. Sharp, J. L., Zammit, T. G., Azar, T. A., Lawson, D. M., 2002. Stress-like responses to common procedures in male rats housed alone or with other rats. Contemp. Top. Lab. Anim. Sci. 41, 8-14. Shepherd, G. M., 2006. Behaviour: Smells, brains and hormones. Nature 439, 149151. Sherwin, C. M., 2002. Comfortable quarters for mice in research institutions. In V. Reinhardt, A. Reinhardt (Eds.), Comfortable Quarters for Laboratory Animals (9 ed., pp. 6-17). Washington: Animal Welfare Institute. Sherwin, C. M., Glen, E. F., 2003. Cage colour preferences and effects of home cage colour on anxiety in laboratory mice. Anim. Behav. 66, 1085-1092. Slawecki, C. J., Roth, J., 2005. Assessment of sustained attention in ad libitum fed Wistar rats: Effects of MK-801. Physiol. Behav. 85, 346-353. Smith, D. V., Margolskee, R. F., 2001. Making sense of taste. Sci. Am. 284, 32. Stevens, D. A., Koster, E. P., 1972. Open-field responses of rats to odors from stressed and nonstressed predecessors. Behav. Biol. 7, 519-525. Stierhoff, K. A., Lavin, M. J., 1982. The influence of rendering rats anosmic on the poisoned-partner effect. Behav. Neur. Biol. 34, 180-189. Strupp, B. J., Levitsky, D. A., 1984. Social transmission of food preferences in adult hooded rats (Rattus norvegicus). J. Comp. Psychol. 98, 257-266. Sundberg, J. P., Boggess, D., Bascom, C., Limberg, B. J., Shultz, L. D., Sundberg, B. A., King, L. E., Jr., Montagutelli, X., 2000. Lanceolate hair-J (lahJ): a mouse model for human hair disorders. Exp. Dermatol. 9, 206-218. Symons, L. A., Tees, R. C., 1990. An examination of the intramodal and intermodal behavioral consequences of long-term vibrissae removal in rats. Dev. Psychobiol. 23, 849-867. Szel, A., Rohlich, P., 1992. Two cone types of rat retina detected by anti-visual pigment antibodies. Exp. Eye Res. 55, 47-52. Takahashi, L. K., Kalin, N. H., Baker, E. W., 1990. Corticotropin-releasing factor antagonist attenuates defensive-withdrawal behavior elicited by odors of stressed conspecifics. Behav. Neurosci. 104, 386-389. Tanapat, P., Hastings, N. B., Rydel, T. A., Galea, L. A. M., Gould, E., 2001. Exposure to fox odor inhibits cell proliferation in the hippocampus of adult rats via an adrenal hormone-dependent mechanism. J. Comp. Neurol. 437, 496-504. Taylor, R. D., Ludvigson, H. W., 1980. Selective removal of reward and nonreward odors to assess their control of patterned responding in rats. Bull. Psychon. Soc. 16, 101-104. Taylor, R. D., Ludvigson, H. W., 1987. Airborne differences in odors emitted by Rattus norvegicus in response to reward and nonreward. J. Chem. Ecol. 13, 1147-1161. Thodberg, K., Malmkvist, J., Herskin, M. S., 2006. The acute effect of a synthetic pig appeasing hormone on open field and intruder test behaviour. Paper presented at the 40th International Congress of the ISAE, Bristol. Tong, T. Y. Y., Goh, V. H. H., 2000. Effect of selected wavelengths of light on the fertility status of rats. Anim. Technol. 51, 1-7. 52

1694 1695 1696 1697 1698 1699 1700 1701 1702 1703 1704 1705 1706 1707 1708 1709 1710 1711 1712 1713 1714 1715 1716 1717 1718 1719 1720 1721 1722 1723 1724 1725 1726 1727 1728 1729 1730 1731 1732 1733 1734 1735 1736 1737 1738 1739 1740 1741 1742 1743

Toro, J. M., Trobalon, J. B., Sebastián-Gallés, N., 2003. The use of prosodic cues in language discrimination tasks by rats. Anim. Cogn. 6, 131-136. Treit, D., Spetch, M. L., Deutsch, J. A., 1983. Variety in the flavor of food enhances eating in the rat: A controlled demonstration. Physiol. Behav. 30, 207-211. Treit, D., Fundytus, M., 1988. Thigmotaxis as a test for anxiolytic activity in rats. Pharmacol. Biochem. Behav. 31, 959-962. Van Der Lee, S., Boot, L. M., 1956. Spontaneous pseudopregnancy in mice. II. Acta Physiol. Pharmacol. Neerl. 5, 213-215. Van der Poel, A. M., Miczek, K. A., 1991. Long ultrasonic calls in male-rats following mating, defeat and aversive-stimulation - Frequency-modulation and bout structure. Behaviour 119, 127-142. van Driel, K., Talling, J., Pickersgill, M., Hendrie, C., Lane, J., Owen, D., 2004. Effect of experience of the handler on stress in rats. Paper presented at the 9th FELASA Symposium: Internationalization and harmonization in laboratory animal care and use issues, Nantes, France. van Driel, K. S., Talling, J. C., 2005. Familiarity increases consistency in animal tests. Behav. Brain Res. 159, 243-245. Van Loo, P. L. P., Kruitwagen, C. L. J. J., Van Zutphen, L. F. M., Koolhaas, J. M., Baumans, V., 2000. Modulation of aggression in male mice: Influence of cage cleaning regime and scent marks. Anim. Welf. 9, 281-295. Vandenbergh, J. G., 1969. Male odor accelerates female sexual maturation in mice. Endocrinology 84, 658-660. Vandenbergh, J. G., 1976. Acceleration of sexual maturation in female rats by male stimulation. J. Reprod. Fertil. 46, 451-453. Vivian, J. A., Miczek, K. A., 1991. Ultrasounds during morphine-withdrawal in rats. Psychopharmacology 104, 187-193. Voipio, H.-M., 1997. How do rats react to sounds? Scand. J. Lab. Anim. Sci. 24, 1-80. Voipio, H.-M., Björk, E., Hakumäki, M., Nevalainen, T., 1998. Sound: from noise to communication. Paper presented at the Scand-LAS Symposium, Hafjell. Wallace, J., Steinert, P. A., Scobie, S. R., Spear, N. E., 1980. Stimulus modality and short-term memory in rats. Anim. Learn. Behav. 8, 10-16. Walton, W. E., 1933. Color vision and color preference in the albino rat. II. The experiments and results. J. Comp. Psychol. 15, 373-394. Walton, W. E., Bornemeier, R. W., 1938. Further evidence of color discrimination in rodents. J. Genet. Psych. 52, 165-181. Ware, N., Mason, G. J., 2003. Does euthanasia by CO2 elicit the production of fear odours in laboratory rats? Using conspecific reaction as a tool for data collection. Unpublished B.A. Honours dissertation, University of Oxford, Oxford. Weaver, M. S., Whiteside, D. A., Janzen, W. C., Moore, S. A., Davis, S. F., 1982. A preliminary investigation into the source of odor-cue production. Bull. Psychon. Soc. 19, 284-286. Westenbroek, C., Snijders, T. A. B., den Boer, J. A., Gerrits, M., Fokkema, D. S., Ter Horst, G. J., 2005. Pair-housing of male and female rats during chronic stress exposure results in gender-specific behavioral responses. Horm. Behav. 47, 620-628. Whishaw, I. Q., Kolb, B., 2005. The behavior of the laboratory rat: a handbook with tests. Oxford: Oxford University Press. Whitten, W. K., 1959. Occurrence of anoestrus in mice caged in groups. J. Endocrinol. 18, 102-107. 53

1744 1745 1746 1747 1748 1749 1750 1751 1752 1753 1754 1755 1756 1757 1758 1759 1760 1761 1762 1763 1764 1765 1766 1767 1768 1769 1770 1771 1772 1773 1774

Williams, J. L., Groux, M. L., 1993. Exposure to various stressors alters preferences for natural odors in rats (Rattus Norvegicus). J. Comp. Psychol. 107, 39-47. Williams, J. L., 1999. Effects of conspecific and predator odors on defensive behavior, analgesia, and spatial working memory. Psychol. Rec. 49, 493-536. Williams, R. A., Howard, A. G., Williams, T. P., 1985. Retinal damage in pigmented and albino rats exposed to low levels of cyclic light following a single mydriatic treatment. Curr. Eye Res. 4, 97-102. Wills, G. D., 1983. Discrimination by olfactory cues in albino rats reflecting familiarity and relatedness among conspecifics. Behav. Neur. Biol. 38, 139143. Wilson, A., Hubrecht, R., Bradley, W. A., Buist, D., Smith, D., Francis, R., James, R., 1995. Housing husbandry and welfare provision for animals used in toxicology studies: results of a UK questionnaire on current practice (1994). Report of the Toxicology and Welfare Working Group, Universities Federation for Animal Welfare. Wolf, N. S., Li, Y., Pendergrass, W., Schmeider, C., Turturro, A., 2000. Normal mouse and rat strains as models for age-related cataract and the effect of caloric restriction on its development. Exp. Eye Res. 70, 683-692. Woodhouse, R., Greenfeld, N., 1985. Responses of albino and hooded rats to various illumination choices in a six-chambered alleyway. Percept. Mot. Skills 61, 343-354. Xerri, C., Coq, J. O., Merzenich, M. M., Jenkins, W. M., 1996. Experience-induced plasticity of cutaneous maps in the primary somatosensory cortex of adult monkeys and rats. J. Physiol. 90, 277-287. Yuan, J., Dieter, M. P., Bucher, J. R., Jameson, C. W., 1993. Application of microencapsulation for toxicology studies. III. Bioavailability of microencapsulated cinnamaldehyde. Fundam. Appl. Toxicol. 20, 83-87. Zala, S. M., Potts, W. K., Penn, D. J., 2004. Scent-marking displays provide honest signals of health and infection. Behav. Ecol. 15, 338-344.

54

1775

Figure 1

1776

Visual perception of rat strains in visual-based behavioral tasks, reprinted from

1777

Prusky and colleagues (2002) (with permission from Elsevier and the authors). The

1778

original image (top-left) has been blurred to model the perception of rats with acuities

1779

of 1.5 c/d (top-right; Fisher–Norway), 1.0 c/d (bottom-left; Dark Agouti, Long-Evans,

1780

wild) and 0.5 c/d (bottom-right; Fisher-344, Sprague–Dawley, Wistar) when the

1781

image subtends 10 degrees. This approximates the size of the image if it were used as

1782

a visual cue in a typical visuo-behavioural task (see Prusky et al., 2002 for details).

1783

1784

55

What is it like to be a rat

This review of rat sensory perception spans eight decades of work conducted across diverse research fields. It covers rat vision, audition, olfaction, gustation, and somatosensation, and describes how rat perception differs from and coincides with ours. As Nagel's seminal work (1974) implies, we cannot truly know what it is ...

387KB Sizes 3 Downloads 454 Views

Recommend Documents

What is it like to be a bat
it is doubtful that any meaning can be attached to the supposition that I should possess the internal ... And we know that while it includes an enormous amount of ...

What is it like to be a bat
What is it like to be a bat http://members.aol.com/NeoNoetics/Nagel_Bat.html. 2 of 9. 20/04/2004 16.12 be ascribed to robots or automata that behaved like ...

What is it like to have ADHD?
Page 1 ... worse during the day, which makes learning even more difficult. At night time it is hard to stop ... muddled and their handwriting is hard to read.

What Is Required To Be A Biomedical Engineer
BookMyEssay is ready to provide the online facility for students, you can easily get every kind of Biomedical Engineering assignment directly from website at lowest cost with several additional benefits. visit us: https://www.bookmyessay.com/biomedic

What is it?
Student's answers are recorded using their plicker cards along with the teachers device and displays the results in real time. plickers. “Plickers is a powerfully simple tool that lets teachers collect real-time formative assessment data without th

The Design Process What it Looks Like
The main problem in software design is what to create in the first place. This process is useful for all kinds of things, but we are going to focus on using it for app ...

What is right has to be done
come, recovery, of a professionally possible and economically affordable treatment .... more easily be identified through data on contribution payments trans-.

What is it made of.pdf
Whoops! There was a problem loading more pages. Whoops! There was a problem previewing this document. Retrying... Download. Connect more apps.

In Response to: What Is a Grid?
Correspondence: Ken Hall, BearingPoint, South Terraces Building, ... the dynamic creation of multiple virtual organizations and ... Int J. Supercomp App. 2001 ...

In Response to: What Is a Grid?
uted and cost-effective way to boost computational power to ... Correspondence: Ken Hall, BearingPoint, South Terraces Building, .... Int J. Supercomp App. 2001 ...

What is Computational Intelligence and what could it ...
formatics, connectionism, data mining, graphical methods, intelligent agents and in- .... nity, presented at conferences, and published in CI journals. .... ject recognition, auditory and visual scene analysis, spatial orientation, memory, motor.