Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Jun 4.
Published in final edited form as: Neurosci Biobehav Rev. 2012 Jan 2;36(3):1001–1014. doi: 10.1016/j.neubiorev.2011.12.012

The use of a running wheel to measure activity in rodents: Relationship to energy balance, general activity, and reward

Colleen M Novak a,*,#, Paul R Burghardt b,#, James A Levine c
PMCID: PMC4455940  NIHMSID: NIHMS694965  PMID: 22230703

Abstract

Running wheels are commonly employed to measure rodent physical activity in a variety of contexts, including studies of energy balance and obesity. There is no consensus on the nature of wheel-running activity or its underlying causes, however. Here, we will begin by systematically reviewing how running wheel availability affects physical activity and other aspects of energy balance in laboratory rodents. While wheel running and physical activity in the absence of a wheel commonly correlate in a general sense, in many specific aspects the two do not correspond. In fact, the presence of running wheels alters several aspects of energy balance, including body weight and composition, food intake, and energy expenditure of activity. We contend that wheel-running activity should be considered a behavior in and of itself, reflecting several underlying behavioral processes in addition to a rodent's general, spontaneous activity. These behavioral processes include defensive behavior, predatory aggression, and depression- and anxiety-like behaviors. As it relates to energy balance, wheel running engages several brain systems—including those related to the stress response, mood, and reward, and those responsive to growth factors—that influence energy balance indirectly. We contend that wheel-running behavior represents factors in addition to rodents' tendency to be physically active, engaging additional neural and physiological mechanisms which can then independently alter energy balance and behavior. Given the impact of wheel-running behavior on numerous overlapping systems that influence behavior and physiology, this review outlines the need for careful design and interpretation of studies that utilize running wheels as a means for exercise or as a measurement of general physical activity.

Keywords: Running wheel, Wheel-running activity, Energy balance, Reward, Brain

1. Introduction

Measurement of physical activity in laboratory rodents has enabled the widespread use of the running wheel. Most rodents readily run in wheels, and it has become an uncomplicated, easily quantifiable measure of physical activity (Sherwin, 1998). The running wheel is commonly used in studies of obesity and energy balance as a surrogate for general activity—that is, the tendency of a rodent to be more or less active. The interest in this area has increased along with evidence that non-exercise activity and the associated energy expenditure (non-exercise activity thermogenesis, or NEAT) is associated with resistance to obesity (Church et al., 2007; Hamilton et al., 2007; Levine et al., 1999, 2005; van Baak et al., 2003; Weinsier et al., 1998). Perhaps even more compelling are suggestions that low physical activity (i.e., sitting) is in itself a risk factor for cardiovascular and metabolic disease (Chomistek et al., 2011; Danaei et al., 2009; Hamilton et al., 2007; Sisson et al., 2009; Stamatakis et al., 2011; Stephens et al., 2011). In humans and in animals, the tendency to be more or less active is shaped by both genetic and environmental factors (Bassett et al., 2004; Joosen et al., 2005; Kaprio et al., 1981; Lanningham-Foster et al., 2003; Novak and Levine, 2007). In order to tease apart the neural, endocrine, and physiological mechanisms underlying individual differences in physical activity, experimental models are employed and physical activity is assessed. The running wheel is often used to assess levels of general physical activity, and other times to model the effects of exercise (Haskell-Luevano et al., 2009; Patterson and Levin, 2008). The purpose of this review is to systematically examine how running wheels complicate the investigation of energy balance and behavior in rodents. Previous reviews have described a range of behavioral effects that a running wheel had on animals and pondered what wheel running might represent in a rodent (Sherwin, 1998); the neurobiology underlying physical activity and wheel running (Garland et al., 2011); and translational studies on the use of the running wheel to model the effects of exercise on health, the brain, and behavior (Haskell-Luevano et al., 2009; Patterson and Levin, 2008). Here, we hypothesize that wheel-running behavior is not solely reflective of the tendency to be physically active, but is a complex and dynamic behavior that interacts with genetics and the environment. We will describe how the running wheel alters rodent behavior and several components of energy balance, delineate how and why this may occur, consider the implications to study design and interpretation, and finally give recommendations regarding the use of running wheels to examine rodent behavior.

2. Access to a running wheel amplifies activity and alters elements of energy balance

Exercise has several well-known benefits to health and fitness (Haskell-Luevano et al., 2009; Patterson and Levin, 2008), as well as neural and cognitive effects (Cotman and Engesser-Cesar, 2002), in humans and laboratory animals; these are outside of the scope of this review. In this section, we will focus on how a running wheel can have unintended consequences on rodent energy balance. Here, we define energy balance as the biological homeostatis of energy in an organism, encompassing energy intake, storage, efficiency, and internal and external work produced. Table 1 outlines studies documenting the effects of wheel-running activity on these different aspects of energy balance. Commonly, when a running wheel is introduced, activity levels are intensified. The amount of wheel running shown by rodents exhibits extreme variability both between and within species, however (Friedman et al., 1992). Though it is rarely noted in publications, it is not unusual to have a small but conspicuous fraction of rodents fail to engage in any significant wheel running, particularly when dealing with mice, which show large strain-related differences in wheel running (de Visser et al., 2006; Lerman et al., 2002). Other rodents are more consistent. For example, Syrian hamsters show very large amounts of wheel running, and the resultant increase in energy expenditure may drive the increased food intake in hamsters (Coutinho et al., 2006b). Body fat mass also decreases in hamsters with wheel running, demonstrating that, even with the increased food intake, negative energy balance is achieved. Similarly, when rats are given free access to a running wheel, they will show high levels of wheel-running activity and increase their food intake compared to rats without wheels (Dixon et al., 2003). Lastly, mice show increased food intake when given access to a wheel (Swallow et al., 2001).

Table 1.

Summary of alterations in energy balance caused by wheel running in rodents.

Species Sex Compartment of energy balance altered Effect Reference
Rat Female Amount of activity; food intake Increased amount of activity; on a restricted diet, induced activity-based anorexia Dixon et al. (2003)
OLETF rat Male Activity; food intake Increased amount of activity; decreased food intake Bi et al. (2005)
Mouse Female and male Food intake Increased food intake, specific to mice bred for low body fat Simoncic et al. (2008), Swallow et al. (2001)
Syrian hamster Female and male Amount of activity Increased amount of activity; increased food intake; weight loss; loss of body fat Coutinho et al. (2006a,b)
A. nitloticus Female and male Activity pattern, amount Increased amount of activity; altered diurnal pattern of activity in subset of animals due to masking Blanchong et al. (1999), Katona and Smale (1997), Redlin and Mrosovsky (2004)
Octodon degus Female and male Activity pattern, amount Increased amount of activity; altered diurnal pattern of activity in subset of animals Kas and Edgar (1999)

The alterations in food intake and energy balance that occur after wheel access serve to underscore the fact that access to running wheels amplifies activity in most rodents (see Table 1 for examples), thereby increasing the need for calories. Even though some studies consider them interchangeable, measuring wheel-running activity is not the same as measuring general cage activity. On the contrary, running wheels have been conclusively demonstrated to alter home cage activity in both mice and rats (de Visser et al., 2005; Hoffmann et al., 1987). In mice with access to wheels, less time is spent in their shelter and ambulating about their home cage compared to mice that were not provided a wheel; the temporal pattern of activity was also altered in mice with access to wheels (de Visser et al., 2005). Recently, it has become more common to investigate physical activity using both wheels as well as general cage activity in succession (Pistilli et al., 2011); this allows for direct comparison between the two methods. Several investigators have identified correlations between wheel running and spontaneous or home-cage physical activity, either according to group or, less commonly, within individual animals (Coyle et al., 2008; Malisch et al., 2008; Morgan and Pfaff, 2001, 2002; Pistilli et al., 2011; Richter et al., 2008; Schmidt et al., 2008; Simoncic et al., 2008). While the two types of activity are usually correlated in a general sense, wheel running intensifies activity, increases variability between individuals, and can amplify group differences in activity. Though variation is seen between species and individuals in the magnitude of the effect, running wheels have a general tendency to increase activity and activity energy expenditure in rodents (Table 1). This then has downstream effects on other aspects of energy balance. When investigating physical activity as a component of energy balance, it is prudent to use a method other than, or in addition to, the running wheel.

Running wheels have the potential to influence several components of energy balance apart from the effects of wheel running on energy expenditure. In female rats over their estrous cycle, “scalloping” of wheel running is often observed, with an advance in the onset of running and increases in wheel-running activity occurring on the morning of estrus (Dixon et al., 2003). This also coincides with decreased energy intake. This effect is subject to species differences which may be linked to high circulating concentrations of estradiol (Cushing and Cawthorn, 1996; Cushing et al., 1995). The influence of the estrous cycle on activity is much more salient when observing wheel running compared to home cage activity. When female rats are placed on a restricted feeding schedule, they develop activity-based anorexia: activity will increase sharply but food intake will not increase to compensate, and weight loss results (Dixon et al., 2003). This in turn disrupts estrous cyclicity. Similarly, an altered response to wheel-induced disruption of energy balance is also seen in rats lacking CCK-A receptors (OLETF rats). These rats are normally obese, diabetic, and hyperphagic. The introduction of wheels at 8 weeks of age normalized food intake, body weight, and circulating leptin and glucose (Bi et al., 2005). The negative energy balance seen in OLETF rats with running wheels was much more pronounced than that seen in their control counterparts: the OLETF rats ate less when presented with a running wheel, whereas the control rats ate more. These studies highlight how the effect of a running wheel on specific aspects of energy balance varies depending on the species, sex, genetic background, and environment. As detailed below, wheel running corresponds more with exercise than general non-exercise activity, at least in how it affects metabolism.

In general, when a rodent is presented with a running wheel, activity will increase sharply, then continue to increase before plateauing, with a concomitant increase in food intake that is presumed to meet the increased energy demands (Cabeza de Vaca et al., 2007; Engel et al., 2009; Hickey et al., 2008; Murray et al., 2010; Pham et al., 2005; Richter et al., 2008) (see Table 1 for examples). This highlights the difference between activity in a running wheel compared to general cage activity. Often, a wheel is introduced to assess the effects of exercise rather to assess physical activity levels (Coutinho et al., 2006a,b; Moraska and Fleshner, 2001; Wood, 2002). Indeed, the physiological results of wheel-running activity in rodents closely parallel the results of exercise in humans, including increased muscle insulin sensitivity, increased skeletal muscle citrate synthase, and decreased visceral fat relative to times when the rats cannot run in wheels (Booth et al., 2008; Davidson et al., 2006; Henriksen and Halseth, 1995). Though voluntary wheel running appears to be a fitting model of physiological effects of aerobic exercise (Booth et al., 2008; Cotman and Engesser-Cesar, 2002; Meeusen, 2005), this does not imply that it is an equally fitting model for the motivation to exercise. What about the processes underlying the motivation of the rodent to voluntarily run in the wheel? Is it analogous to the motivation to exercise in humans? Are the mechanistic processes and motivations that predispose a rodent to run in their wheel for extended distances and time periods the same processes and motivations that result in a more active rodent in the absence of a wheel? In general, evidence does not support this assertion; several factors are known to differentially alter physical activity versus wheel running, which will be discussed in the following sections.

3. Running wheels can alter daily activity patterns

Activity wheels are customarily used to measure circadian activity rhythms in all suborders of rodents. The introduction of an activity wheel can also alter the temporal pattern of activity (see Table 1). The murid rodent Arvicanthis niloticus shows diurnal activity in the field (Blanchong et al., 1999); a study using timed trapping devices showed that A. niloticus were found outside of their burrow almost exclusively during the daytime hours (Blanchong et al., 1999), leading to the conclusion that they are strictly diurnal. This parallels their general cage activity in the absence of a wheel (Blanchong et al., 1999). In the laboratory setting, however, access to a running wheel induces a very distinct pattern of activity. A subset of the animals will “switch” to a pattern of activity that appears to be nocturnal; removal of the wheel returns the animal to its previous diurnal pattern of activity (Blanchong et al., 1999; Katona and Smale, 1997). The change in activity pattern is abrupt, indicating that a circadian phase shift is unlikely to underlie this phenomenon. Blanchong et al. (1999) also found that the propensity to switch activity patterns in the presence of a wheel is inherited (Blanchong et al., 1999). The effect of a wheel on activity in A. niloticus is likely to be due to masking, or the direct influence of environmental lighting conditions on activity rather than underlying circadian processes. When lights are turned on intermittently throughout the light phase of the cycle (subjective day) the animals will run on their wheels primarily during the dark periods. Without wheels, however, the animals will be active preferentially when lights are on (Redlin and Mrosovsky, 2004). In this diurnal species, access to a running wheel profoundly alters the daily pattern of activity, and it is likely that brain regions activated by arousal and reward mediate this “switch” (Castillo-Ruiz et al., 2010). A similar wheel-induced alteration in the temporal pattern of activity can be seen in another rodent, Octodon degus. A subset of degus invert their phase preference from diurnal to nocturnal when housed with a running wheel. The activity pattern immediately switches back to diurnal upon removal of the wheel (Kas and Edgar, 1999). The running wheel has the ability to dramatically alter the temporal pattern of activity in these rodent species. In one case, this phenomenon is subject to individual differences which appear to be heritable, potentially through single-gene inheritance of an autosomal dominant gene with incomplete penetrance, sex-linked inheritance, or epigenetic mechanisms (Blanchong et al., 1999). These findings support the assertion that wheel-running activity is distinctly different from general activity. Moreover, it is likely that the neural systems underlying these different types of activity overlap, but not completely. This theme is reinforced when one examines how caloric restriction differentially modulates wheel-running versus general physical activity.

4. Food availability affects wheel running and spontaneous physical activity differently

Alterations in food availability affect physical activity, both with and without the running wheel. In general, acute starvation increases general physical activity over the short term in several species, presumably as a foraging response (Lynn et al., 2003; Novak et al., 2005; Williams et al., 2002). In contrast, long-term starvation decreases levels of physical activity (Novak et al., 2005; Severinsen and Munch, 1999), apparently to conserve energy. The complete absence of food (starvation/food restriction) has different effects than caloric restriction, wherein rodents are fed a reduced number of calories each day, most commonly 60–70% of pre-restriction levels. Caloric restriction suppresses physical activity levels in rats (Severinsen and Munch, 1999). This is generally interpreted as an effort to conserve energy in the face of decreased food availability. In mice, on the other hand, caloric restriction decreases dark-phase physical activity but increases light-phase activity, resulting in no net change in daily physical activity (Overton and Williams, 2004; Williams et al., 2002).

How does complete or partial caloric restriction affect activity in a wheel in rodents? As described above, in female rats, starvation and caloric restriction both increase wheel running to the detriment of the animal's health (Hebebrand et al., 2003; Morse et al., 1995; Pirke et al., 1993; Russell et al., 1987). [Activity-based anorexia or starvation-induced hyperactivity (Hebebrand et al., 2003) is not evident in mice (Rikke et al., 2003).] This paradoxical effect has been attributed to a response to peripheral energy balance signals such as leptin (Exner et al., 2000), an attempt to thermoregulate in response to change in ambient temperature (Gutierrez et al., 2002; Williams et al., 2002), or as a representation of foraging. Indeed, the running wheel has been used as a model of foraging whereby rodents must work (i.e., run) for food (Day and Bartness, 2001). Circadian influences should also be taken into account: caloric restriction usually entails giving the rodents food at a specific time each day. Because of a food-entrainable oscillator, the rodent will anticipate the time of food availability and increase their activity in the time period before the food arrives (Stephan, 2002).

Thus, several factors interact to determine how food availability alters general, spontaneous activity in rodents, including species, strain or genetic background, sex, phase of the estrous cycle, ambient temperature, the type and magnitude of food restriction, time of day, and of course the presence of a running wheel. Though innate general cage activity levels appear to be roughly correlated with wheel-running activity levels (Coyle et al., 2008; Waters et al., 2008), it is clear that the two are not equivalent. It is also apparent that wheel-running activity does not accurately represent general physical activity in rodents during caloric restriction, and that the presence of the wheel has profound effects on energy balance, especially in rats. In addition to the general tendency to be physically active, what other factors account for wheel-running activity in rodents? To answer this question, we will examine behavioral aspects of wheel running which, at first glance, may appear to be unrelated to the regulation of energy balance. Though the capacity for voluntary wheel running to influence stress and/or behavior has been reviewed several times (Dishman, 1997; Greenwood and Fleshner, 2008; Sothmann et al., 1996; Stranahan et al., 2008), here, we will focus on behaviors which have underlying neurochemical systems in common with energy balance regulation.

5. Behavioral aspects of wheel running: wheel running affects motivation and reward systems

Increased attention is being paid to the importance of brain reward system activation in energy balance regulation, especially as it relates to appetite (Fulton, 2010; Zheng et al., 2009). These same brain reward systems play a role in wheel-running behavior; examples are given in Table 2. In fact, the act of wheel running itself can become a self-perpetuating behavior that has the capacity to reach obsessive levels. Wheel-running behavior in rodents follows a fairly consistent pattern, increasing for several weeks until the total distance covered in a given period of time plateaus. Total running duration then decreases over time or with ageing. Beyond this general pattern of wheel-running behavior, it is important to recognize that a fair amount of individual variability occurs. This variance is likely due to the capacity of the animal to engage in wheel-running behavior coupled with the amount of reward the animal experiences from wheel running, as well as other environmental factors (Table 2). Taken together, these factors will influence an animal's propensity and motivation to engage in wheel running.

Table 2.

Summary of reward-related behaviors altered by wheel running in rodents.

Species Sex Duration of access to wheel Effect on behavior Reference
Rat (Wistar) Female and male 22 weeks unrestricted access Food restriction increased wheel running, while increased access to food decreased wheel running Finger (1951)
Rat (Sprague–Dawley) Male 10 min per day Rats lever-pressed for access to running wheels Kagan and Berkun (1954)
Rat (albino from Holtzman Research) Male Up to 14 days, unrestricted access Food was ignored upon presentation of wheel, resulting in self-starvation Routtenberg and Kuznesof (1967)
Rat (Sprague–Dawley) Male Varied Rats lever-pressed for access to running wheels Collier and Hirsch (1971)
Rat (Long Evans) Male 2 h per day, 5 days/week for 8 weeks Increased brain catecholamine levels related to running, weight loss, and hyperphagia de Castro and Duncan (1985)
Rat (Sprague–Dawley) Female and male Varied Food deprivation increased lever pressing for access to a running wheel; acute wheel exposure during food deprivation decreased responding for food reward Pierce et al. (1986)
Rat (Sprague–Dawley) Male Up to 70% of starting body-weight, or 20 days unrestricted access Activity-induced anorexia (and death) was reduced by propylene glycol but not ethanol Spigelman et al. (1991)
Rat (Long–Evans) Female Varied Rats lever-pressed for access to running wheels Iversen (1993)
Rat (Sprague–Dawley) Male Alternating access Wheel access decreased oral amphetamine intake Kanarek et al. (1995)
Rat (Wistar) Male Varied Rats lever-pressed for access to running wheels; running behavior was dependent on duration of opportunity to run Belke (1997)
Rat (Lewis and Fischer) 30 days, unrestricted Access Addiction-prone rats exhibited higher levels of running compared to non-addiction-prone rats Werme et al. (1999)
Rat (Sprague–Dawley) Male 2-h controlled access on alternating days, over ~2 weeks Wheel running induced conditioned place preference, which was reversible by the opioid antagonist naloxone Lett et al. (2001)
Rat (Sprague–Dawley) Male 2-h controlled access on alternating days, over ~2 weeks Wheel running blocked the ability of morphine to produce conditioned place preference Lett et al. (2002)
Rat (Long Evans) Female 8 weeks unrestricted access Wheel running decreased breakpoint on a progressive ratio schedule for cocaine self-administration Cosgrove et al. (2002)
Rat (Sprague–Dawley) Male 24 days, 2-h per day or unrestricted access Escalated running with unrestricted access Lattanzio and Eikelboom (2003)
Rat (Wistar) Male Varied Rats developed conditioned place preference to context associated with wheel Belke and Wagner (2005)
Rat (Sprague–Dawley) Male 25 days Wheel access suppressed sucrose and food intake Satvat and Eikelboom (2006)
Rat (Long Evans) Female 8 weeks unrestricted access Wheel running increased conditioned place preference to cocaine Smith et al. (2008)
Rat (Fischer 344) Male 2 and 6 weeks Increased conditioned place preference after 6 weeks of wheel running Greenwood et al. (2011)
Mice (BALB-c) Male 90 min Wheel running induced c-Fos in nucleus accumbens (core) Vargas-Perez et al. (2003)
Mice (C57BL/6J) Female Varied Wheel running increased in the absence of ethanol Ozburn et al. (2008)
Mice (BALB-c-type and D2L receptor-deficient) Male 30 min/day for 20 days Wheel running engages dopamine and opioid reward systems; motivational state affects ability of wheel to engage these systems Vargas-Perez et al. (2004), Vargas-Perez et al. (2008)
Syrian hamster Male Up to 35 days Amount of wheel running was positively correlated with self-administration of testosterone, which has reinforcing properties Wood (2002)

5.1. Voluntary wheel running is rewarding and interacts with other reinforcers

Table 2 gives examples of evidence supporting the assertion that the running wheel is rewarding to rodents. Rats will lever-press for access to running wheels (Belke, 1997; Collier and Hirsch, 1971; Iversen, 1993; Kagan and Berkun, 1954), indicating that, similar to drug self-administration paradigms, wheel running is rewarding and rats will work for the opportunity to run. The rewarding aspect of voluntary wheel running suggests that neural pathways regulating reward to substances of abuse may also be activated by wheel running. This is supported by the finding that rats who exhibit the highest levels of lever-pressing also exhibit the highest levels of wheel running (Iversen, 1993), suggesting that those individuals that receive the most reward from wheel running are willing to work the hardest for it. In addition to rats' motivation to work for access to running wheels, it is apparent that unrestricted wheel running can reach levels that could be considered compulsive and extreme (Lattanzio and Eikelboom, 2003). Due to the reinforcing effect of wheel running, the running wheel is becoming a more frequently used tool to investigate the rewarding properties of drugs of abuse. For example, rats that are considered “addiction prone”—they are more likely to self-administer addictive drugs than rats from other strains—also exhibit excessive levels of voluntary wheel running (Werme et al., 1999); examples are cited in Table 2. Female rats given access to running wheels in combination with food restriction will further escalate their running behavior and will forego the opportunity to eat, ultimately leading to their death (Routtenberg and Kuznesof, 1967; Spigelman et al., 1991). This behavioral pattern suggests that animals are willing to engage in wheel-running behavior in a way that parallels humans' willingness to engage in drug use despite dangerous or dire consequences (Robinson, 2004). Another parallel is the heightened activity seen with decreased caloric intake in human anorexics, resulting in profound negative energy balance (Hebebrand et al., 2003; Scheurink et al., 2010). Interestingly, anorexic patients also show alterations in the processing of rewarding stimuli (Jappe et al., 2011).

Similar to other reinforcers, the running wheel can be used in hedonic substitution as an alternate reinforcer (see examples in Table 2). For example, wheel running increases when mice are denied access to ethanol (Ozburn et al., 2008). It has been suggested that the running wheel can serve as an alternative source of motivation, replacing drugs of abuse such as cocaine (Smith et al., 2008) because voluntary wheel running can substantially decrease intravenous self-administration of cocaine in rodents (Cosgrove et al., 2002; Smith et al., 2008). Wheel running also decreases oral consumption of amphetamine in rats (Kanarek et al., 1995). More relevant to energy balance, studies demonstrate that wheel running also decreases motivation for natural reinforcers such as food (Iversen, 1993) and a sucrose solution (Satvat and Eikelboom, 2006). Conversely, food deprivation increases lever pressing for both cocaine and access to running wheels (Carroll, 1984; Finger, 1951; Pierce et al., 1986). Therefore, in addition to directly impacting metabolism through exercise and its downstream impact on energy expenditure and appetite, wheel running may also alter energy balance via indirect means: modulating appetite by affecting the rewarding and reinforcing aspects of food. This hypothesized indirect alteration of energy balance by wheel running differs from the effect of general physical activity on energy balance, and is another source of individual variability in wheel running. Perhaps most importantly, these data support the idea that neural mechanisms controlling appetite and reward overlap, and that these brain systems are activated by wheel running.

5.2. Wheel running involves brain reward systems activated by drugs of abuse

The impact of the running wheel on reward and other reinforcing stimuli is reflected in alterations in brain reward mechanisms. Individual differences in wheel-running activity in rats mirrored differences in response to amphetamine, and exhibited a “cross-sensitization” between wheel running and amphetamine (Ferreira et al., 2006). This implies similar underlying mechanisms mediating wheel-running activity and addiction or reward (Ferreira et al., 2006). The mesolimbic and mesocortical dopamine systems have long been known to play a role in substance abuse (Feltenstein and See, 2008; Self, 1998), and the same dopaminergic systems are influenced by voluntary exercise (Meeusen and De Meirleir, 1995). Thirty minutes of voluntary wheel running increases dopamine metabolism in the striatum of mice (Bliss and Ailion, 1971). Eight weeks of wheel running resulted in elevated catecholamine concentrations (dopamine and norepinephrine) in the brain that were correlated with weight loss and hyperphagia induced by wheel running (de Castro and Hill, 1988).

In addition to the dopaminergic system, wheel running is known to influence endogenous opioids. Endogenous opioid systems are integrally related to mesolimbic and mesocortical dopaminergic function and regulate the response of these dopaminergic systems to drugs of abuse. Exposure to running wheels for 2 h a day over eight days promotes tolerance to morphine as measured by conditioned place preference (Lett et al., 2002), and wheel running itself induces conditioned place preference which is reversible by administration of the opioid receptor antagonist naloxone (Lett et al., 2001). Chronic wheel running decreases the analgesic properties of morphine (D'Anci et al., 2000; Kanarek et al., 1998; Mathes and Kanarek, 2001) with pain regulatory regions of the brain implicated in this response (Mathes and Kanarek, 2006). Further, dynorphin mRNA is increased in the medial caudate putamen following wheel running and cocaine administration (Werme et al., 2000). Recently, alterations in both dopamine and opioid systems in the basal fore-brain (Bjornebekk et al., 2008; Greenwood et al., 2011), as well as increased tyrosine hydroxylase mRNA levels in the ventral tegmental area, were reported following 6 weeks of voluntary wheel running (Greenwood et al., 2011). Levels of ΔFosB protein, known to be involved in drug reward (Hope et al., 1994; Nestler et al., 2001) and food reward (Teegarden et al., 2008), as well as c-Fos, were both increased following voluntary wheel running (Greenwood et al., 2011; Vargas-Perez et al., 2003). The authors also reported conditioned place preference in rats following six weeks, but not two weeks, of voluntary wheel running (Greenwood et al., 2011), linking the changes in neurochemistry to a behavioral endpoint. Lastly, reward system activation by wheel running depends on genetic background (Bjornebekk et al., 2006), thus the distinct neurocircuitry activation in different lines and strains of rodents may contribute to differential behavioral and neurochemical responses to voluntary running.

6. Behavioral aspects of wheel running: stress and anxiety-like/depression-like behavior

As summarized in Table 3, a number of studies have reported alterations in behavior following access to running wheels in both rats and mice. Often, these changes are attributed to exercise, and many of these effects depend on the duration of exposure to the running wheel. Eight weeks, but not 4 weeks, of voluntary wheel running increased measures of defensive behavior and risk assessment in rats (Burghardt et al., 2004). Lines of mice selected for high levels of voluntary wheel running also showed heightened levels of predatory aggression (Gammie et al., 2003). Extended access to running wheels appears to reduce learned helplessness: six weeks, but not 3 weeks, of running wheel access resulted in decreased escape latency in the shuttle-box test (Greenwood et al., 2005, 2003). Consistent with these results, changes in activation patterns in the amygdala, a limbic region involved in fear and anxiety, are found between sedentary and wheel-running groups following contextual fear conditioning. In this case, eight weeks of wheel running altered c-Fos levels in the amygdala following contextual fear conditioning (Burghardt et al., 2006b).

Table 3.

Summary of behaviors altered by wheel running in rodents.

Species Sex Duration of access to wheel Effect on behavior Reference
Rat (JCR:LA-cp) Male Intermittent Induced conditioned taste aversion; not secondary to novelty Heth et al. (2001)
Rat (Long–Evans) Male 22 days Increased conditioned freezing Van Hoomissen et al. (2004)
Rat (Sprague–Dawley) Male 8 weeks (effect not seen at 4 weeks) Increased defensive behavior; Increased risk assessment Burghardt et al. (2004)
Rat (Flinders Sensitive/Resistant Lines) Male 30 days Decreased immobility in FSL rats; showed a trend toward increased immobility in FRL rats Bjornebekk et al. (2005)
Rat (Sprague–Dawley (2003) and Fischer F344 (2005) Male 6 weeks (effect not seen at 3 weeks) Reduced learned helplessness Greenwood et al. (2005), Greenwood et al. (2003)
Rat (Sprague–Dawley) Male 8 weeks Increased contextual conditioned freezing Burghardt et al. (2006a,b)
Rat (Sprague–Dawley) Male 5–8 weeks unrestricted access Protective against chronic unpredictable stress: reduced stress effect on spatial learning and consumption of sucrose solution Zheng et al. (2006)
Rat (Flinders Sensitive Line) Female 30 days Decreased depression-like behavior Bjornebekk et al. (2008)
Rat (Sprague–Dawley) Male 4 weeks unrestricted access Increased immobility in the forced swim test Collins et al. (2009)
Mouse (C57BL/6J) Not reported 9 weeks unrestricted access Decreased immobility in forced swim test; delayed decreases in sucrose intake following chronic stress Solberg et al. (1999)
Mouse Female and male N/A (artificially selected for high wheel running) Increased predatory and maternal aggression Gammie et al. (2003)
Mouse (C57BL/6J) Male 4 weeks unrestricted Increased anxiety-like behavior in open field, decreased anxiety-like behavior in elevated plus maze Binder et al. (2004)
Mouse (C57BL/6J) Male 3–4 weeks Decreased depression-like behavior Duman et al. (2008)
Mouse (CF1) Not reported 4 weeks Increased hippocampal-dependent learning Muller et al. (2011)

Access to a running wheel also alters depression-like behavior or coping strategies in rodents (Table 3). These behaviors are commonly assessed using a number of different tests including the forced swim test; in this test, immobility, or “floating,” is considered a measure of depression-like behavior (Cryan et al., 2005). Generally, decreases in immobility in the forced swim test have been reported following wheel running (Bjornebekk et al., 2005; Solberg et al., 1999). Further, wheel running produced a larger reduction in immobility compared to the antidepressant escitalopram in the forced swim test, and the effects of antidepressant compounds on immobility in the forced swim test are augmented by wheel running (Bjornebekk et al., 2008). In contrast, a recent study has reported increased immobility in the forced swim test following voluntary wheel running, which the authors attributed to increases in `passive coping' following extended wheel running (Collins et al., 2009). Wheel running in mice has been shown to decrease measures of anxiety-like behavior (Binder et al., 2004; Duman et al., 2008) and had an anti-depression-like effect (Duman et al., 2008).

The relationship between stress and weight gain has been recognized for some time, and the vilification of hormones in the stress axis (e.g., cortisol) tends to be fodder for a number of late-night infomercial remedies for obesity. The ability of stress-axis hormones (e.g., corticosterone, cortisol) to influence body weight and energy mobilization has been recognized for several decades. Further, there is evidence that wheel running can affect several levels of the HPA-axis in rats and mice (de Rijke et al., 2005; Droste et al., 2007, 2003). There also appears to be a reciprocal relationship between stress and wheel running as social isolation stress prior to wheel access inhibits hippocampal neurogenesis (Stranahan et al., 2006). Wheel running protects against decreases in hippocampal brain-derived neurotrophic factor (BDNF) mRNA expression following a single session of restraint stress (Adlard and Cotman, 2004). Wheel running also protects against reductions in growth factor mRNA levels in the hippocampus, reduced spatial learning, and consumption of sucrose solution following four weeks of chronic unpredictable stress (Zheng et al., 2006). In addition, stress-induced reductions in BDNF levels do not appear to be directly related to circulating corticosterone (Adlard and Cotman, 2004; Zheng et al., 2006). Interestingly, the ability of wheel running to prevent decreases in hippocampal BDNF mRNA following stress may not be functionally relevant for the beneficial effects of wheel running to block learned helplessness behavior (Greenwood et al., 2007). These factors can be affected by low-intensity running as well, perhaps to an even greater degree than with high-intensity exercise, at least in juvenile rats (Lou et al., 2008). This opens up the possibility that non-exercise activity could affect neurogenesis and related factors in the brain, but possibly in a different manner than does high-intensity activity or wheel running.

Thus, the effects of wheel running on rodent behavior do not fit easily into one unified behavioral profile–increased predatory aggression and defensive behaviors, but decreased depression- and anxiety-like behaviors (Table 3). The lack of consensus may be due to several factors. First, there may be a tendency to anthropomorphize rodent behavior (Holmes, 2003), ultimately resulting in the development of misleading hypotheses. Second, the `control' group to which voluntary running is compared may influence the interpretation of results (Dubreucq et al., 2011) as would the time of day when the behavioral testing occurs (Hopkins and Bucci, 2010). Third, food restriction prior to behavioral testing can substantially influence behavior (Heiderstadt et al., 2000; Inoue et al., 2004; Jahng et al., 2007; McDermott and Kelly, 2008), although the magnitude and direction of this effect may very well depend on the metabolic profile of the animal. Fourth, the use of different species, strains of rats or mice, or selective breeding for various traits, can produce markedly different behavioral responses to a given behavioral test like the elevated plus maze, forced swim test, locomotor response to novelty, or wheel-running behavior (Berton et al., 1997; Dichter et al., 1996; Kabbaj et al., 2000; Overstreet and Rezvani, 1996; Tejani-Butt et al., 2003). With these caveats in mind, the interpretation of behavioral and metabolic effects of exercise must be clearly defined within the context of the experimental design as it is likely that these factors interact. When using wheel running to assess rodent energy balance, it is important to recognize that wheel-running behavior may both cause and/or be affected by changes in metabolic function. The use of the running wheel provides an exciting opportunity to address hypotheses relating to the relationship between mood and metabolism, particularly when considering the neurovegetative symptoms associated with several forms of depression, and given the comorbidity of affective/mood disorders with metabolic dysfunction (Lawrence and Kopelman, 2004; Sullivan et al., 1993) and emerging data indicating overlap in the brain systems that shape mood and energy balance (Lutter et al., 2008a,b; Novak et al., 2006; Teske et al., 2006; Zheng et al., 2009).

7. Behavioral aspects of wheel running: wheel running influences energy balance and behavior by changing the physical structure of the brain

Changes in brain structure are associated with changes in brain function and behavior, including behaviors that alter energy balance. Growth factors play a number of essential roles in brain function by influencing neuroplasticity, vascularization, and neurogenesis, ultimately leading to restructuring and rewiring of the brain. One of the most studied regions of the brain exhibiting plasticity is the hippocampus, an area rich in growth factors that plays an integral role in memory and emotion. The hippocampus provides a robust example of the convergence of genetics and environment in altering plasticity of neural circuitry, and is one of the main regions where neurogenesis can be observed in the adult. Although most commonly associated with learning, memory, and emotion, there is evidence that the hippocampus is also involved in the regulation of energy balance (Davidson et al., 2007). Strikingly, the hippocampus contains both insulin and leptin receptors (Lathe, 2001), indicating this structure may be a direct target for peripheral peptides that signal energy status. In line with this, rats and humans with damage to the hippocampal formation show decreased inhibitory control of food intake (Hebben et al., 1985; Rozin and Dow, 1998) and increased appetitive behavior (Clifton et al., 1998; Schmelzeis and Mittleman, 1996), respectively. Therefore changes in hippocampal structure may have implications for a variety of behaviors including those related to energy management (Table 4 contains examples of wheel-running induced brain alterations).

Table 4.

Summary of neuroanatomical changes secondary to wheel running in rodents.

Species Sex Duration of access to wheel Effect on brain Reference
Rats (spontaneously hypertensive) Males 5–6 weeks Wheel running increased CSF beta-endorphin Hoffmann et al. (1990)
Rat (Sprague–Dawley) Male 2–7 nights of running Increased FGF2 mRNA and IR in hippocampus following 2 and 4 nights of running Gomez-Pinilla et al. (1997)
Rat (Sprague–Dawley) Not reported 20 days Increased BDNF mRNA in hippocampus Russo-Neustadt et al. (1999)
Rat (Sprague–Dawley) Male 1–28 days unrestricted Increased BDNF protein and mRNA (4 weeks of running was required) Adlard et al. (2004)
Rat (Sprague–Dawley) Male 2–90 days (continuous or alternating days) Increased hippocampal BDNF with alternating and continuous running Berchtold et al. (2005)
Rat (Sprague–Dawley) Male 3–48 days Neurogenesis in socially isolated rats (chronic wheel running was required) Stranahan et al. (2006)
Mouse (C57BL/6J) Male 7 days Increased VGF mRNA and protein Hunsberger et al. (2007)
Mouse (C57BL/6) Male 3 weeks unrestricted Protected against stress-induced decreases in hippocampal BDNF Adlard and Cotman (2004)
Mouse (C57BL/6) Not reported 2 weeks Increased hippocampal neurogenesis which correlated with regional cerebral blood flow as measured by fMRI Pereira et al. (2007)
Mouse (C57BL/6) Female 12 days Increased neurogenesis and neuronal survival in the hippocampus van Praag et al. (1999)
Mouse (C57BL/6, beta-endorphin deficient) Male 10 days or 39 days Beta-endorphin was necessary for wheel-induced hippocampal cell proliferation Koehl et al. (2008)
Mouse (C57BL/6) Male 1, 3, or 10 days Increased vascularization and neurogenesis in hippocampus Van der Borght et al. (2009)
Mouse (CF1) Not reported 4 weeks Improved insulin receptor signaling and glucose oxidation in hippocampu Muller et al. (2011)

A variety of environmental conditions and behaviors, including wheel running, have the ability to influence plasticity and neurogenesis through their actions on growth factor systems; examples are given in Table 4. The BDNF and fibroblast growth factor (FGF) families function in hippocampal plasticity; both BDNF and FGF are responsive to voluntary wheel running. Wheel running increases growth factor expression (Berchtold et al., 2005; Cotman and Berchtold, 2002; Hunsberger et al., 2007; Russo-Neustadt et al., 1999) and neurogenesis (van Praag et al., 1999) in the rodent hippocampus. It is noteworthy that interventions that alter growth factor expression, including antidepressant administration (Berton and Nestler, 2006; Mallei et al., 2002) and environmental enrichment (Ickes et al., 2000; Turner and Lewis, 2003), also bring about concurrent improvements in behavior. Moreover, wheel-running activity, but not other types of activity, significantly improved the ability of rats to learn a classical conditioning task (Green et al., 2011). It is clear, therefore, that wheel running produces differences in hippocampal growth factor expression that relate to behavioral endpoints.

Alterations in growth factor expression may be required for behavioral changes in response to exercise and antidepressants (Adachi et al., 2008; Russo-Neustadt et al., 1999). Interestingly, transient changes in several growth factor families occur after only a couple days of exercise. While growth factor alterations and behavior appear not to have a direct relationship (Bjornebekk et al., 2008; Greenwood et al., 2007), the relatively acute effect of wheel running on growth factor expression (Gomez-Pinilla et al., 1997) could potentially be a necessary precursor to functional and structural alterations in the brain that take several weeks to produce behavioral adaptations (Adlard et al., 2004; Burghardt et al., 2004, 2006b; Greenwood et al., 2005; Van Hoomissen et al., 2004). It should also be noted that there is a distinct time course for changes in mRNA levels compared to protein levels for BDNF following exercise (Adlard et al., 2004). Further emphasizing the importance of timing, a recent report has shown that the processes of vascularization and neurogenesis begin relatively quickly (within 10 days) after the introduction of running wheels (Van der Borght et al., 2009). Intriguingly, these changes are very labile and the increases in vascularization and neurogenesis found in the hippocampus after several days of wheel running returned to baseline levels 24 h after the cessation of wheel running (Van der Borght et al., 2009). Therefore, the temporal effects of wheel running on neuroplasticity systems and behavior must be taken into account when animals are given access to running wheels.

Wheel running alters the neurochemical systems involved in plasticity and neurogenesis in the hippocampus (Bjornebekk et al., 2005; Hoffmann et al., 1990; Koehl et al., 2008; Pereira et al., 2007; Persson et al., 2004), as listed in Table 4, but how do these actions translate to changes in energy balance? Many of the growth factors and neuropeptide systems altered by running wheels impact the brain's control of energy balance. Specific examples are emerging that suggest wheel-running-induced changes in the hippocampus could impinge on hypothalamic circuitry controlling metabolism and behavior, including a potential role of BDNF (Adachi et al., 2008). First, BDNF and its receptor, trkB, impact food intake and energy expenditure through their actions in the hypothalamus (Wang et al., 2007a,b,c), and have recently been implicated as downstream regulators of melanocortin-system control of food intake (Bariohay et al., 2009). Second, the FGF family appears to be involved in regulating feeding (Oomura et al., 1992) as levels of FGF increase in cerebral spinal fluid postprandially or following a glucose infusion. Beyond changes in growth factor families, receptors for neuropeptides and hormones involved in energy management (e.g., leptin and insulin) are expressed in the hippocampus. This provides an opportunity for energy-management signals from the periphery to directly influence the function of the hippocampus, suggesting a role for the hippocampus in regulating food intake (Davidson et al., 2007). Indeed, leptin influences behavior via hippocampal circuitry (Liu et al., 2010; Lu et al., 2006) and has been shown to be neuroprotective (Tang, 2008). Further, improvements in hippocampal-dependent learning and memory, as well as in insulin receptor signaling and glucose oxidation in the hippocampus, are seen after voluntary wheel running in mice (Muller et al., 2011). Additionally, increases in NPY and NPY-1 receptor mRNA, a neuropeptide family long known to play a role in feeding, have been reported in the hippocampus following wheel running (Bjornebekk et al., 2010).

The overlap of processes underlying energy balance, wheel running, and neurogenesis is further illustrated by the report that the endogenous endocannabinoid system (eCB) is involved in hippocampal neurogenesis in response to wheel running (Hill et al., 2010); the eCBsystem is also known to modulate feeding and stress response (Fride, 2005). Additionally, eCB regulation of exercise may be related to metabolic status as the motivation for wheel running, and the wheel-running behavior itself, is differentially responsive to cannabinoid ligands in lean compared to obese Zucker rats (Smith and Rasmussen, 2010). Taken together, the apparent role for the hippocampus in energy management, coupled with the robust effect of exercise on hippocampal structure and function, indicate that the influence of wheel running on the hippocampus should not be discounted in studies of food intake and energy expenditure. It is well known that wheel running results in alterations in neuropeptide levels in brain areas more commonly associated with energy management (e.g., hypothalamus) (Bi et al., 2005; de Rijke et al., 2005; Lewis et al., 1993). It is becoming more apparent that wheel running has the potential to impact energy balance indirectly through alterations in brain regions such as the hippocampus and, as described above, brain reward systems.

8. Models for wheel running: mice and rats selectively bred for wheel running

Because of the varied effects of running wheels on physiology and behavior of rodents, and the variability of responses between different species and individual animals within species, a cohesive picture of rodent wheel running is lacking (Sherwin, 1998). One method that may clarify the underlying mechanisms behind these phenomena is selective breeding for high and low levels of wheel running. Such breeding programs have been undertaken for rats and mice. The results of these studies will be described in turn, as well as physical activity and wheel-running activity in rats selectively bred for other traits.

Mice have been selectively bred for high levels of wheel running for over 35 generations that display levels of wheel running 170% greater than control lines (Rezende et al., 2006; Swallow et al., 1998). Four replicate selected and control lines were created, and several variables have been measured in these mice compared to control-line counterparts. Selection for high wheel-running activity yields mice with lower body weight and fat content (Swallow et al., 1999, 2001). But does high wheel-running activity predetermine high activity in general? Using short (3 min) tests of open field activity (Bronikowski et al., 2001), or behavioral observation (Koteja et al., 1999), no differences were detected between high-wheel-running and control lines of mice (Bronikowski et al., 2001). Open-field activity is mostly used as a measure of stress, anxiety, or locomotor response to novelty rather than overall spontaneous activity levels (Stead et al., 2006). Indeed, when daily activity was measured, robust differences in cage activity were found between selected and control lines of mice (Malisch et al., 2008, 2009). It was suggested that wheel running motivation may be the key factor differentiating the running and non-running strains, and that this is not necessarily linked to performance capacity (Rezende et al., 2005). Mice selected for high levels of wheel running also display behaviors such as increased exploration and risk-taking, in addition to increased anxiety (Jonas et al., 2010). Together, these studies highlight the importance of standardized daily measures of physical activity, and illustrate the likelihood of overlap between the biological and behavioral systems governing daily physical activity and wheel running in rodents. Dopaminergic brain reward pathways are also heightened in mice bred for high wheel-running activity (Mathes et al., 2010), again supporting the hypothesis that wheel running is associated with changes in brain reward system activation, differentiating the neural systems underlying wheel running versus general physical activity.

Lines of rats showing high levels of voluntary wheel-running activity have also been established (Morishima-Yamato et al., 2005), with male Spontaneously-Running-Tokushima-Shikoku (SPORTS) rats running roughly six times longer than controls. Male SPORTS rats had lower body weight, circulating insulin, and visceral fat than controls, as well as increased skeletal muscle oxidative capacity (Morishima-Yamato et al., 2005). Here again, the authors concluded that the selected rats have a heightened “running motivation.” General activity in the absence of a wheel was not examined, so we do not know whether or not running motivation correlates with spontaneous daily activity in these rats.

Examining wheel running and spontaneous physical activity in other selectively bred rodents also gives insight into the behavioral and physiological traits underlying wheel-running and spontaneous activity. Rats bred for high intrinsic aerobic capacity (HCRs) have high levels of both activity on running wheels (Burghardt et al., 2006a; Vaanholt et al., 2008; Waters et al., 2008) and general locomotor activity (Novak et al., 2010, 2009). Within each selected line, the level of variability between rats' general locomotor activity is surprisingly low (Novak et al., 2010, 2009), whereas there is a large amount of variance between the amount of activity on a wheel between rats, especially HCRs (Burghardt et al., 2006a; Waters et al., 2008). Interestingly, locomotor response to novelty does not differ between HCR and LCR rats (Burghardt et al., 2011). This hints that intrinsic aerobic capacity and physical activity may share underlying causes, and that some of these mechanisms—but not all—overlap with wheel-running activity.

Comparisons of different selected lines may also help us to understand the effects of diet on spontaneous activity. When rats or mice are fed a high-fat diet, physical activity typically declines (Bjursell et al., 2008; Novak et al., 2006) [though see (Simoncic et al., 2008)]. We have also recently found that both HCR and LCR rats share a similar decrease in physical activity after high-fat feeding (Novak et al., 2010). This effect is relatively consistent across most rodents studied, with the exception of rats bred specifically for resistance to weight gain on a high-fat diet (Novak et al., 2006). Moreover, there is an acute decrease in physical activity levels in mice, seen as soon as three days after giving the high-fat Western diet (Bjursell et al., 2008), making it less likely to be secondary to increases in body weight; a high-fat diet does not appear to affect wheel running in the same way (Vaanholt et al., 2008). In mice selected for high levels of wheel running, a high-fat diet increases spontaneous daily activity, and this effect was specific to female mice (Vaanholt et al., 2008). This further underlines how perturbations of energy balance differentially affect activity on versus off of a wheel, and how selected lines of rats and mice can be utilized to determine how these properties overlap.

9. Conclusions

An animal's locomotion about its environment is critical to its survival on many levels. More recently, the association of high intrinsic physical activity levels with a lean phenotype has been a topic of interest due to the ever-increasing prevalence of obesity and the need to find useful, effective obesity treatments. While it is not uncommon to see a positive association between high general physical activity levels and high wheel-running activity, it is unlikely that wheel running accurately represents inherent tendencies to be generally physically active in rodents. On the contrary, access to a running wheel affects several aspects of normal energy balance, most notably through the amplification of activity and energy expenditure, though other routes cannot be discounted. The myriad factors that differentially affect general cage activity and wheel-running activity are outlined in Fig. 1. Though many more recent investigations of physical activity utilize measures of general activity in the absence of or in addition to wheel-running activity (Pistilli et al., 2011), it is still not uncommon to refer to these two methods interchangeably. To more accurately and precisely characterize general physical activity in laboratory studies, the following guidelines are suggested:

Fig. 1.

Fig. 1

Several factors affect the amount of wheel running a rodent will display. These same traits and factors will also influence general physical activity, but often not in the same manner or to the same extent. Whereas levels of general physical activity positively correlate with wheel-running activity, the presence of a wheel will increase overall physical activity while decreasing non-wheel home-cage activity. Lastly, wheel running alters an array of behavioral and physiological variables in ways that general home-cage activity does not, and wheel-running activity has more profound effects on several aspects of metabolism than does general physical activity.

Apparent contradictions and confusion can be avoided by following these guidelines when designing studies and considering the literature. Time and effort can be saved if investigators are aware of how running wheels will affect rodents in their studies.

Acknowledgements

We appreciate the assistance of Antonio Nunez, Lydia Heemstra, Peter Bodary, Maria Waselus, and Kate Dykhuis for comments on the manuscript during preparation.

References

  1. Adachi M, Barrot M, Autry AE, Theobald D, Monteggia LM. Selective loss of brain-derived neurotrophic factor in the dentate gyrus attenuates antidepressant efficacy. Biol. Psychiatry. 2008;63(7):642–649. doi: 10.1016/j.biopsych.2007.09.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Adlard PA, Cotman CW. Voluntary exercise protects against stress-induced decreases in brain-derived neurotrophic factor protein expression. Neuroscience. 2004;124(4):985–992. doi: 10.1016/j.neuroscience.2003.12.039. [DOI] [PubMed] [Google Scholar]
  3. Adlard PA, Perreau VM, Engesser-Cesar C, Cotman CW. The timecourse of induction of brain-derived neurotrophic factor mRNA and protein in the rat hippocampus following voluntary exercise. Neurosci. Lett. 2004;363(1):43–48. doi: 10.1016/j.neulet.2004.03.058. [DOI] [PubMed] [Google Scholar]
  4. Bariohay B, Roux J, Tardivel C, Trouslard J, Jean A, Lebrun B. Brain-derived neurotrophic factor/tropomyosin-related kinase receptor type B signaling is a downstream effector of the brainstem melanocortin system in food intake control. Endocrinology. 2009;150(6):2646–2653. doi: 10.1210/en.2008-1184. [DOI] [PubMed] [Google Scholar]
  5. Bassett DR, Schneider PL, Huntington GE. Physical activity in an Old Order Amish community. Med. Sci. Sports Exerc. 2004;36(1):79–85. doi: 10.1249/01.MSS.0000106184.71258.32. [DOI] [PubMed] [Google Scholar]
  6. Belke TW. Running and responding reinforced by the opportunity to run: effect of reinforcer duration. J. Exp. Anal. Behav. 1997;67(3):337–351. doi: 10.1901/jeab.1997.67-337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Belke TW, Wagner JP. The reinforcing property and the rewarding aftereffect of wheel running in rats: a combination of two paradigms. Behav. Processes. 2005;68(2):165–172. doi: 10.1016/j.beproc.2004.12.006. [DOI] [PubMed] [Google Scholar]
  8. Berchtold NC, Chinn G, Chou M, Kesslak JP, Cotman CW. Exercise primes a molecular memory for brain-derived neurotrophic factor protein induction in the rat hippocampus. Neuroscience. 2005;133(3):853–861. doi: 10.1016/j.neuroscience.2005.03.026. [DOI] [PubMed] [Google Scholar]
  9. Berton O, Nestler EJ. New approaches to antidepressant drug discovery: beyond monoamines. Nat. Rev. Neurosci. 2006;7(2):137–151. doi: 10.1038/nrn1846. [DOI] [PubMed] [Google Scholar]
  10. Berton O, Ramos A, Chaouloff F, Mormde P. Behavioral reactivity to social and nonsocial stimulations: a multivariate analysis of six inbred rat strains. Behav. Genet. 1997;27(2):155–166. doi: 10.1023/a:1025641509809. [DOI] [PubMed] [Google Scholar]
  11. Bi S, Scott KA, Hyun J, Ladenheim EE, Moran TH. Running wheel activity prevents hyperphagia and obesity in Otsuka Long–Evans Tokushima Fatty rats: role of hypothalamic signaling. Endocrinology. 2005;146(4):1676–1685. doi: 10.1210/en.2004-1441. [DOI] [PubMed] [Google Scholar]
  12. Binder E, Droste SK, Ohl F, Reul JM. Regular voluntary exercise reduces anxiety-related behaviour and impulsiveness in mice. Behav. Brain Res. 2004;155(2):197–206. doi: 10.1016/j.bbr.2004.04.017. [DOI] [PubMed] [Google Scholar]
  13. Bjornebekk A, Mathe AA, Brene S. The antidepressant effect of running is associated with increased hippocampal cell proliferation. Int. J. Neuropsychopharmacol. 2005;8(3):357–368. doi: 10.1017/S1461145705005122. [DOI] [PubMed] [Google Scholar]
  14. Bjornebekk A, Mathe AA, Brene S. Running has differential effects on NPY, opiates, and cell proliferation in an animal model of depression and controls. Neuropsychopharmacology. 2006;31(2):256–264. doi: 10.1038/sj.npp.1300820. [DOI] [PubMed] [Google Scholar]
  15. Bjornebekk A, Mathe AA, Brene S. The antidepressant effects of running and escitalopram are associated with levels of hippocampal NPY and Y1 receptor but not cell proliferation in a rat model of depression. Hippocampus. 2010;20(7):820–828. doi: 10.1002/hipo.20683. [DOI] [PubMed] [Google Scholar]
  16. Bjornebekk A, Mathe AA, Gruber SH, Brene S. Housing conditions modulate escitalopram effects on antidepressive-like behaviour and brain neurochemistry. Int. J. Neuropsychopharmacol. 2008;11(8):1135–1147. doi: 10.1017/S1461145708008912. [DOI] [PubMed] [Google Scholar]
  17. Bjursell M, Gerdin AK, Lelliott CJ, Egecioglu E, Elmgren A, Tornell J, Oscarsson J, Bohlooly YM. Acutely reduced locomotor activity is a major contributor to Western diet-induced obesity in mice. Am. J. Physiol. Endocrinol. Metab. 2008;294(2):E251–E260. doi: 10.1152/ajpendo.00401.2007. [DOI] [PubMed] [Google Scholar]
  18. Blanchong JA, McElhinny TL, Mahoney MM, Smale L. Nocturnal and diurnal rhythms in the unstriped Nile rat, Arvicanthis niloticus. J. Biol. Rhythms. 1999;14(5):364–377. doi: 10.1177/074873099129000777. [DOI] [PubMed] [Google Scholar]
  19. Bliss EL, Ailion J. Relationship of stress and activity to brain dopamine and homovanillic acid. Life. Sci. I. 1971;10(20):1161–1169. doi: 10.1016/0024-3205(71)90276-1. [DOI] [PubMed] [Google Scholar]
  20. Booth FW, Laye MJ, Lees SJ, Rector RS, Thyfault JP. Reduced physical activity and risk of chronic disease: the biology behind the consequences. Eur. J. Appl. Physiol. 2008;102(4):381–390. doi: 10.1007/s00421-007-0606-5. [DOI] [PubMed] [Google Scholar]
  21. Bronikowski AM, Carter PA, Swallow JG, Girard IA, Rhodes JS, Garland T., Jr. Open-field behavior of house mice selectively bred for high voluntary wheel-running. Behav. Genet. 2001;31(3):309–316. doi: 10.1023/a:1012283426530. [DOI] [PubMed] [Google Scholar]
  22. Burghardt PR, Britton SL, Koch LG, Watson SJ, Akil H. Paper Presented at the Society for Neuroscience. Atlanta, GA: 2006a. Selective breeding for intrinsic aerobic capacity alters levels of neuropeptides that modulate energy management. [Google Scholar]
  23. Burghardt PR, Flagel SB, Burghardt KJ, Britton SL, Gerard-Koch L, Watson SJ, Akil H. Risk-assessment and coping strategies segregate with divergent intrinsic aerobic capacity in rats. Neuropsychopharmacology. 2011;36(2):390–401. doi: 10.1038/npp.2010.144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Burghardt PR, Fulk LJ, Hand GA, Wilson MA. The effects of chronic tread-mill and wheel running on behavior in rats. Brain Res. 2004;1019(1–2):84–96. doi: 10.1016/j.brainres.2004.05.086. [DOI] [PubMed] [Google Scholar]
  25. Burghardt PR, Pasumarthi RK, Wilson MA, Fadel J. Alterations in fear conditioning and amygdalar activation following chronic wheel running in rats. Pharmacol. Biochem. Behav. 2006b;84(2):306–312. doi: 10.1016/j.pbb.2006.05.015. [DOI] [PubMed] [Google Scholar]
  26. Cabeza de Vaca S, Kannan P, Pan Y, Jiang N, Sun Y, Carr KD. The adenosine A2A receptor agonist, CGS-21680, blocks excessive rearing, acquisition of wheel running, and increases nucleus accumbens CREB phosphorylation in chronically food-restricted rats. Brain Res. 2007;1142:100–109. doi: 10.1016/j.brainres.2007.01.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Carroll ME. Food deprivation produces persistent increases in self-administration behavior during cocaine extinction. NIDA Res. Monogr. 1984;55:125–131. [PubMed] [Google Scholar]
  28. Castillo-Ruiz A, Nixon JP, Smale L, Nunez AA. Neural activation in arousal and reward areas of the brain in day-active and night-active grass rats. Neuroscience. 2010;165(2):337–349. doi: 10.1016/j.neuroscience.2009.10.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Chomistek AK, Chiuve SE, Jensen MK, Cook NR, Rimm EB. Vigorous physical activity, mediating biomarkers, and risk of myocardial infarction. Med. Sci. Sports Exerc. 2011;43(10):1884–1890. doi: 10.1249/MSS.0b013e31821b4d0a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Church TS, Earnest CP, Skinner JS, Blair SN. Effects of different doses of physical activity on cardiorespiratory fitness among sedentary, overweight or obese postmenopausal women with elevated blood pressure: a randomized controlled trial. JAMA. 2007;297(19):2081–2091. doi: 10.1001/jama.297.19.2081. [DOI] [PubMed] [Google Scholar]
  31. Clifton PG, Vickers SP, Somerville EM. Little and often: ingestive behavior patterns following hippocampal lesions in rats. Behav. Neurosci. 1998;112(3):502–511. doi: 10.1037//0735-7044.112.3.502. [DOI] [PubMed] [Google Scholar]
  32. Collier G, Hirsch E. Reinforcing properties of spontaneous activity in the rat. J. Comp. Physiol. Psychol. 1971;77(1):155–160. doi: 10.1037/h0031588. [DOI] [PubMed] [Google Scholar]
  33. Collins A, Hill LE, Chandramohan Y, Whitcomb D, Droste SK, Reul JM. Exercise improves cognitive responses to psychological stress through enhancement of epigenetic mechanisms and gene expression in the dentate gyrus. PLoS ONE. 2009;4(1):e4330. doi: 10.1371/journal.pone.0004330. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Cosgrove KP, Hunter RG, Carroll ME. Wheel-running attenuates intravenous cocaine self-administration in rats: sex differences. Pharmacol. Biochem. Behav. 2002;73(3):663–671. doi: 10.1016/s0091-3057(02)00853-5. [DOI] [PubMed] [Google Scholar]
  35. Cotman CW, Berchtold NC. Exercise: a behavioral intervention to enhance brain health and plasticity. Trends Neurosci. 2002;25(6):295–301. doi: 10.1016/s0166-2236(02)02143-4. [DOI] [PubMed] [Google Scholar]
  36. Cotman CW, Engesser-Cesar C. Exercise enhances and protects brain function. Exerc. Sport Sci. Rev. 2002;30(2):75–79. doi: 10.1097/00003677-200204000-00006. [DOI] [PubMed] [Google Scholar]
  37. Coutinho AE, Campbell JE, Fediuc S, Riddell MC. Effect of voluntary exercise on peripheral tissue glucocorticoid receptor content and the expression and activity of 11beta-HSD1 in the Syrian hamster. J. Appl. Physiol. 2006a;100(5):1483–1488. doi: 10.1152/japplphysiol.01236.2005. [DOI] [PubMed] [Google Scholar]
  38. Coutinho AE, Fediuc S, Campbell JE, Riddell MC. Metabolic effects of voluntary wheel running in young and old Syrian golden hamsters. Physiol. Behav. 2006b;87(2):360–367. doi: 10.1016/j.physbeh.2005.10.006. [DOI] [PubMed] [Google Scholar]
  39. Coyle CA, Strand SC, Good DJ. Reduced activity without hyperphagia contributes to obesity in Tubby mutant mice. Physiol. Behav. 2008;95(1–2):168–175. doi: 10.1016/j.physbeh.2008.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Cryan JF, Valentino RJ, Lucki I. Assessing substrates underlying the behavioral effects of antidepressants using the modified rat forced swimming test. Neurosci. Biobehav. Rev. 2005;29(4–5):547–569. doi: 10.1016/j.neubiorev.2005.03.008. [DOI] [PubMed] [Google Scholar]
  41. Cushing BS, Cawthorn JM. Species differences in activity patterns during oestrus. Can. J. Zool.-Rev. Can. Zool. 1996;74(3):473–479. [Google Scholar]
  42. Cushing BS, Marhenke S, McClure PA. Estradiol concentration and the regulation of locomotor activity. Physiol. Behav. 1995;58(5):953–957. doi: 10.1016/0031-9384(95)00158-f. [DOI] [PubMed] [Google Scholar]
  43. D'Anci KE, Gerstein AV, Kanarek RB. Long-term voluntary access to running wheels decreases kappa-opioid antinociception. Pharmacol. Biochem. Behav. 2000;66(2):343–346. doi: 10.1016/s0091-3057(00)00245-8. [DOI] [PubMed] [Google Scholar]
  44. Danaei G, Ding EL, Mozaffarian D, Taylor B, Rehm J, Murray CJ, Ezzati M. The preventable causes of death in the United States: comparative risk assessment of dietary, lifestyle, and metabolic risk factors. PLoS Med. 2009;6(4):e1000058. doi: 10.1371/journal.pmed.1000058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Davidson SR, Burnett M, Hoffman-Goetz L. Training effects in mice after long-term voluntary exercise. Med. Sci. Sports Exerc. 2006;38(2):250–255. doi: 10.1249/01.mss.0000183179.86594.4f. [DOI] [PubMed] [Google Scholar]
  46. Davidson TL, Kanoski SE, Schier LA, Clegg DJ, Benoit SC. A potential role for the hippocampus in energy intake and body weight regulation. Curr. Opin. Pharmacol. 2007;7(6):613–616. doi: 10.1016/j.coph.2007.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Day DE, Bartness TJ. Effects of foraging effort on body fat and food hoarding in Siberian hamsters. J. Exp. Zool. 2001;289(3):162–171. [PubMed] [Google Scholar]
  48. de Castro JM, Duncan G. Operantly conditioned running: effects on brain catecholamine concentrations and receptor densities in the rat. Pharmacol. Biochem. Behav. 1985;23(4):495–500. doi: 10.1016/0091-3057(85)90407-1. [DOI] [PubMed] [Google Scholar]
  49. de Castro JM, Hill JO. Exercise and brain catecholamine relationships with brown adipose tissue and whole-body oxygen consumption in rats. Physiol. Behav. 1988;43(1):9–12. doi: 10.1016/0031-9384(88)90090-x. [DOI] [PubMed] [Google Scholar]
  50. de Rijke CE, Hillebrand JJ, Verhagen LA, Roeling TA, Adan RA. Hypothalamic neuropeptide expression following chronic food restriction in sedentary and wheel-running rats. J. Mol. Endocrinol. 2005;35(2):381–390. doi: 10.1677/jme.1.01808. [DOI] [PubMed] [Google Scholar]
  51. de Visser L, van den Bos R, Kuurman WW, Kas MJ, Spruijt BM. Novel approach to the behavioural characterization of inbred mice: automated home cage observations. Genes Brain Behav. 2006;5(6):458–466. doi: 10.1111/j.1601-183X.2005.00181.x. [DOI] [PubMed] [Google Scholar]
  52. de Visser L, van den Bos R, Spruijt BM. Automated home cage observations as a tool to measure the effects of wheel running on cage floor locomotion. Behav. Brain Res. 2005;160(2):382–388. doi: 10.1016/j.bbr.2004.12.004. [DOI] [PubMed] [Google Scholar]
  53. Della Maggiore V, Ralph MR. The effect of amphetamine on locomotion depends on the motor device utilized. The open field vs. the running wheel. Pharmacol. Biochem. Behav. 2000;65(4):585–590. doi: 10.1016/s0091-3057(99)00260-9. [DOI] [PubMed] [Google Scholar]
  54. Dichter GS, Brunelli SA, Hofer MA. Elevated plus-maze behavior in adult offspring of selectively bred rats. Physiol. Behav. 1996;60(1):299–304. doi: 10.1016/0031-9384(95)02222-8. [DOI] [PubMed] [Google Scholar]
  55. Dishman RK. Brain monoamines, exercise, and behavioral stress: animal models. Med. Sci. Sports Exerc. 1997;29(1):63–74. doi: 10.1097/00005768-199701000-00010. [DOI] [PubMed] [Google Scholar]
  56. Dixon DP, Ackert AM, Eckel LA. Development of, and recovery from, activity-based anorexia in female rats. Physiol. Behav. 2003;80(2–3):273–279. doi: 10.1016/j.physbeh.2003.08.008. [DOI] [PubMed] [Google Scholar]
  57. Droste SK, Chandramohan Y, Hill LE, Linthorst AC, Reul JM. Voluntary exercise impacts on the rat hypothalamic–pituitary–adrenocortical axis mainly at the adrenal level. Neuroendocrinology. 2007;86(1):26–37. doi: 10.1159/000104770. [DOI] [PubMed] [Google Scholar]
  58. Droste SK, Gesing A, Ulbricht S, Muller MB, Linthorst AC, Reul JM. Effects of long-term voluntary exercise on the mouse hypothalamic–pituitary–adrenocortical axis. Endocrinology. 2003;144(7):3012–3023. doi: 10.1210/en.2003-0097. [DOI] [PubMed] [Google Scholar]
  59. Dubreucq S, Marsicano G, Chaouloff F. Emotional consequences of wheel running in mice: which is the appropriate control? [Research Support, Non-U.S. Gov't] Hippocampus. 2011;21(3):239–242. doi: 10.1002/hipo.20778. [DOI] [PubMed] [Google Scholar]
  60. Duman CH, Schlesinger L, Russell DS, Duman RS. Voluntary exercise produces antidepressant and anxiolytic behavioral effects in mice. Brain Res. 2008;1199:148–158. doi: 10.1016/j.brainres.2007.12.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Engel SR, Creson TK, Hao Y, Shen Y, Maeng S, Nekrasova T, Landreth GE, Manji HK, Chen G. The extracellular signal-regulated kinase pathway contributes to the control of behavioral excitement. Mol. Psychiatry. 2009;14(4):448–461. doi: 10.1038/sj.mp.4002135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Exner C, Hebebrand J, Remschmidt H, Wewetzer C, Ziegler A, Herpertz S, Schweiger U, Blum WF, Preibisch G, Heldmaier G, Klingenspor M. Leptin suppresses semi-starvation induced hyperactivity in rats: implications for anorexia nervosa. Mol. Psychiatry. 2000;5(5):476–481. doi: 10.1038/sj.mp.4000771. [DOI] [PubMed] [Google Scholar]
  63. Feltenstein MW, See RE. The neurocircuitry of addiction: an overview. Br. J. Pharmacol. 2008;154(2):261–274. doi: 10.1038/bjp.2008.51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Ferreira A, Lamarque S, Boyer P, Perez-Diaz F, Jouvent R, Cohen-Salmon C. Spontaneous appetence for wheel-running: a model of dependency on physical activity in rat. Eur. Psychiatry. 2006;21(8):580–588. doi: 10.1016/j.eurpsy.2005.02.003. [DOI] [PubMed] [Google Scholar]
  65. Finger FW. The effect of food deprivation and subsequent satiation upon general activity in the rat. J. Comp. Physiol. Psychol. 1951;44(6):557–564. doi: 10.1037/h0055692. [DOI] [PubMed] [Google Scholar]
  66. Fride E. Endocannabinoids in the central nervous system: from neuronal networks to behavior. Curr. Drug Targets CNS Neurol. Disord. 2005;4(6):633–642. doi: 10.2174/156800705774933069. [DOI] [PubMed] [Google Scholar]
  67. Friedman WA, Garland T, Jr., Dohm MR. Individual variation in locomotor behavior and maximal oxygen consumption in mice. Physiol. Behav. 1992;52(1):97–104. doi: 10.1016/0031-9384(92)90438-8. [DOI] [PubMed] [Google Scholar]
  68. Fulton S. Appetite and reward. Front. Neuroendocrinol. 2010;31(1):85–103. doi: 10.1016/j.yfrne.2009.10.003. [DOI] [PubMed] [Google Scholar]
  69. Gammie SC, Hasen NS, Rhodes JS, Girard I, Garland T., Jr. Predatory aggression, but not maternal or intermale aggression, is associated with high voluntary wheel-running behavior in mice. Horm. Behav. 2003;44(3):209–221. doi: 10.1016/s0018-506x(03)00140-5. [DOI] [PubMed] [Google Scholar]
  70. Garland T, Jr., Schutz H, Chappell MA, Keeney BK, Meek TH, Copes LE, Acosta W, Drenowatz C, Maciel RC, van Dijk G, Kotz CM, Eisenmann JC. The biological control of voluntary exercise, spontaneous physical activity and daily energy expenditure in relation to obesity: human and rodent perspectives. J. Exp. Biol. 2011;214(Pt 2):206–229. doi: 10.1242/jeb.048397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Gomez-Pinilla F, Dao L, So V. Physical exercise induces FGF-2 and its mRNA in the hippocampus. Brain Res. 1997;764(1–2):1–8. doi: 10.1016/s0006-8993(97)00375-2. [DOI] [PubMed] [Google Scholar]
  72. Green JT, Chess AC, Burns M, Schachinger KM, Thanellou A. The effects of two forms of physical activity on eyeblink classical conditioning. Behav. Brain Res. 2011;219(1):165–174. doi: 10.1016/j.bbr.2011.01.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Greenwood BN, Fleshner M. Exercise, learned helplessness, and the stress-resistant brain. Neuromol. Med. 2008;10(2):81–98. doi: 10.1007/s12017-008-8029-y. [DOI] [PubMed] [Google Scholar]
  74. Greenwood BN, Foley TE, Burhans D, Maier SF, Fleshner M. The consequences of uncontrollable stress are sensitive to duration of prior wheel running. Brain Res. 2005;1033(2):164–178. doi: 10.1016/j.brainres.2004.11.037. [DOI] [PubMed] [Google Scholar]
  75. Greenwood BN, Foley TE, Day HE, Campisi J, Hammack SH, Campeau S, Maier SF, Fleshner M. Freewheel running prevents learned helplessness/behavioral depression: role of dorsal raphe serotonergic neurons. J. Neurosci. 2003;23(7):2889–2898. doi: 10.1523/JNEUROSCI.23-07-02889.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Greenwood BN, Foley TE, Le TV, Strong PV, Loughridge AB, Day HE, Fleshner M. Long-term voluntary wheel running is rewarding and produces plasticity in the mesolimbic reward pathway. Behav. Brain Res. 2011;217(2):354–362. doi: 10.1016/j.bbr.2010.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Greenwood BN, Strong PV, Foley TE, Thompson RS, Fleshner M. Learned helplessness is independent of levels of brain-derived neurotrophic factor in the hippocampus. Neuroscience. 2007;144(4):1193–1208. doi: 10.1016/j.neuroscience.2006.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Gutierrez E, Vazquez R, Boakes RA. Activity-based anorexia: ambient temperature has been a neglected factor. Psychon. Bull. Rev. 2002;9(2):239–249. doi: 10.3758/bf03196278. [DOI] [PubMed] [Google Scholar]
  79. Hamilton MT, Hamilton DG, Zderic TW. Role of low energy expenditure and sitting in obesity, metabolic syndrome, type 2 diabetes, and cardiovascular disease. Diabetes. 2007;56(11):2655–2667. doi: 10.2337/db07-0882. [DOI] [PubMed] [Google Scholar]
  80. Haskell-Luevano C, Schaub JW, Andreasen A, Haskell KR, Moore MC, Koerper LM, Rouzaud F, Baker HV, Millard WJ, Walter G, Litherland SA, Xiang Z. Voluntary exercise prevents the obese and diabetic metabolic syndrome of the melanocortin-4 receptor knockout mouse. FASEB J. 2009;23(2):642–655. doi: 10.1096/fj.08-109686. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Hebben N, Corkin S, Eichenbaum H, Shedlack K. Diminished ability to interpret and report internal states after bilateral medial temporal resection: case H.M. Behav. Neurosci. 1985;99(6):1031–1039. doi: 10.1037//0735-7044.99.6.1031. [DOI] [PubMed] [Google Scholar]
  82. Hebebrand J, Exner C, Hebebrand K, Holtkamp C, Casper RC, Remschmidt H, Herpertz-Dahlmann B, Klingenspor M. Hyperactivity in patients with anorexia nervosa and in semistarved rats: evidence for a pivotal role of hypoleptinemia. Physiol. Behav. 2003;79(1):25–37. doi: 10.1016/s0031-9384(03)00102-1. [DOI] [PubMed] [Google Scholar]
  83. Heiderstadt KM, McLaughlin RM, Wright DC, Walker SE, Gomez-Sanchez CE. The effect of chronic food and water restriction on open-field behaviour and serum corticosterone levels in rats. Lab. Anim. 2000;34(1):20–28. doi: 10.1258/002367700780578028. [DOI] [PubMed] [Google Scholar]
  84. Henriksen EJ, Halseth AE. Adaptive responses of GLUT-4 and citrate synthase in fast-twitch muscle of voluntary running rats. Am. J. Physiol. 1995;268(1 (Pt 2)):R130–R134. doi: 10.1152/ajpregu.1995.268.1.R130. [DOI] [PubMed] [Google Scholar]
  85. Heth CD, Inglis P, Russell JC, Pierce WD. Conditioned taste aversion induced by wheel running is not due to novelty of the wheel. Physiol. Behav. 2001;74(1–2):53–56. doi: 10.1016/s0031-9384(01)00553-4. [DOI] [PubMed] [Google Scholar]
  86. Hickey MA, Kosmalska A, Enayati J, Cohen R, Zeitlin S, Levine MS, Chesselet MF. Extensive early motor and non-motor behavioral deficits are followed by striatal neuronal loss in knock-in Huntington's disease mice. Neuroscience. 2008;157(1):280–295. doi: 10.1016/j.neuroscience.2008.08.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Hill MN, Titterness AK, Morrish AC, Carrier EJ, Lee TT, Gil-Mohapel J, Gorzalka BB, Hillard CJ, Christie BR. Endogenous cannabinoid signaling is required for voluntary exercise-induced enhancement of progenitor cell proliferation in the hippocampus. Hippocampus. 2010;20(4):513–523. doi: 10.1002/hipo.20647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Hoffmann P, Terenius L, Thoren P. Cerebrospinal fluid immunoreactive beta-endorphin concentration is increased by voluntary exercise in the spontaneously hypertensive rat. Regul. Pept. 1990;28(2):233–239. doi: 10.1016/0167-0115(90)90021-n. [DOI] [PubMed] [Google Scholar]
  89. Hoffmann P, Thoren P, Ely D. Effect of voluntary exercise on open-field behavior and on aggression in the spontaneously hypertensive rat (SHR) Behav. Neural. Biol. 1987;47(3):346–355. doi: 10.1016/s0163-1047(87)90461-4. [DOI] [PubMed] [Google Scholar]
  90. Holmes PV. Rodent models of depression: reexamining validity without anthropomorphic inference. Crit. Rev. Neurobiol. 2003;15(2):143–174. doi: 10.1615/critrevneurobiol.v15.i2.30. [DOI] [PubMed] [Google Scholar]
  91. Hope BT, Nye HE, Kelz MB, Self DW, Iadarola MJ, Nakabeppu Y, Duman RS, Nestler EJ. Induction of a long-lasting AP-1 complex composed of altered Fos-like proteins in brain by chronic cocaine and other chronic treatments. Neuron. 1994;13(5):1235–1244. doi: 10.1016/0896-6273(94)90061-2. [DOI] [PubMed] [Google Scholar]
  92. Hopkins ME, Bucci DJ. Interpreting the effects of exercise on fear conditioning: the influence of time of day. Behav. Neurosci. 2010;124(6):868–872. doi: 10.1037/a0021200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Hunsberger JG, Newton SS, Bennett AH, Duman CH, Russell DS, Salton SR, Duman RS. Antidepressant actions of the exercise-regulated gene VGF. Nat. Med. 2007;13(12):1476–1482. doi: 10.1038/nm1669. [DOI] [PubMed] [Google Scholar]
  94. Ickes BR, Pham TM, Sanders LA, Albeck DS, Mohammed AH, Granholm AC. Long-term environmental enrichment leads to regional increases in neurotrophin levels in rat brain. Exp. Neurol. 2000;164(1):45–52. doi: 10.1006/exnr.2000.7415. [DOI] [PubMed] [Google Scholar]
  95. Inoue K, Zorrilla EP, Tabarin A, Valdez GR, Iwasaki S, Kiriike N, Koob GF. Reduction of anxiety after restricted feeding in the rat: implication for eating disorders. Biol. Psychiatry. 2004;55(11):1075–1081. doi: 10.1016/j.biopsych.2004.01.026. [DOI] [PubMed] [Google Scholar]
  96. Iversen IH. Techniques for establishing schedules with wheel running as reinforcement in rats. J. Exp. Anal. Behav. 1993;60(1):219–238. doi: 10.1901/jeab.1993.60-219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Jahng JW, Kim JG, Kim HJ, Kim BT, Kang DW, Lee JH. Chronic food restriction in young rats results in depression- and anxiety-like behaviors with decreased expression of serotonin reuptake transporter. Brain Res. 2007;1150:100–107. doi: 10.1016/j.brainres.2007.02.080. [DOI] [PubMed] [Google Scholar]
  98. Jappe LM, Frank GK, Shott ME, Rollin MD, Pryor T, Hagman JO, Yang TT, Davis E. Heightened sensitivity to reward and punishment in anorexia nervosa. Int. J. Eat. Disord. 2011;44(4):317–324. doi: 10.1002/eat.20815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Jonas I, Schubert KA, Reijne AC, Scholte J, Garland T, Jr., Gerkema MP, Scheurink AJ, Nyakas C, van Dijk G. Behavioral traits are affected by selective breeding for increased wheel-running behavior in mice. Behav. Genet. 2010;40(4):542–550. doi: 10.1007/s10519-010-9359-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Joosen AM, Gielen M, Vlietinck R, Westerterp KR. Genetic analysis of physical activity in twins. Am. J. Clin. Nutr. 2005;82(6):1253–1259. doi: 10.1093/ajcn/82.6.1253. [DOI] [PubMed] [Google Scholar]
  101. Kabbaj M, Devine DP, Savage VR, Akil H. Neurobiological correlates of individual differences in novelty-seeking behavior in the rat: differential expression of stress-related molecules. J. Neurosci. 2000;20(18):6983–6988. doi: 10.1523/JNEUROSCI.20-18-06983.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Kagan J, Berkun M. The reward value of running activity. J. Comp. Physiol. Psychol. 1954;47(2):108. doi: 10.1037/h0058877. [DOI] [PubMed] [Google Scholar]
  103. Kanarek RB, Gerstein AV, Wildman RP, Mathes WF, D'Anci KE. Chronic running-wheel activity decreases sensitivity to morphine-induced analgesia in male and female rats. Pharmacol. Biochem. Behav. 1998;61(1):19–27. doi: 10.1016/s0091-3057(98)00059-8. [DOI] [PubMed] [Google Scholar]
  104. Kanarek RB, Marks-Kaufman R, D'Anci KE, Przypek J. Exercise attenuates oral intake of amphetamine in rats. Pharmacol. Biochem. Behav. 1995;51(4):725–729. doi: 10.1016/0091-3057(95)00022-o. [DOI] [PubMed] [Google Scholar]
  105. Kaprio J, Koskenvuo M, Sarna S. Cigarette smoking, use of alcohol, and leisure-time physical activity among same-sexed adult male twins. Prog. Clin. Biol. Res. 1981;69(Pt C):37–46. [PubMed] [Google Scholar]
  106. Kas MJ, Edgar DM. A nonphotic stimulus inverts the diurnal-nocturnal phase preference in Octodon degus. J. Neurosci. 1999;19(1):328–333. doi: 10.1523/JNEUROSCI.19-01-00328.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Katona C, Smale L. Wheel-running rhythms in Arvicanthis niloticus. Physiol. Behav. 1997;61(3):365–372. doi: 10.1016/s0031-9384(96)00407-6. [DOI] [PubMed] [Google Scholar]
  108. Koehl M, Meerlo P, Gonzales D, Rontal A, Turek FW, Abrous DN. Exercise-induced promotion of hippocampal cell proliferation requires beta-endorphin. FASEB J. 2008;22(7):2253–2262. doi: 10.1096/fj.07-099101. [DOI] [PubMed] [Google Scholar]
  109. Koteja P, Garland T, Jr., Sax JK, Swallow JG, Carter PA. Behaviour of house mice artificially selected for high levels of voluntary wheel running. Anim. Behav. 1999;58(6):1307–1318. doi: 10.1006/anbe.1999.1270. [DOI] [PubMed] [Google Scholar]
  110. Lanningham-Foster L, Nysse LJ, Levine JA. Labor saved, calories lost: the energetic impact of domestic labor-saving devices. Obes. Res. 2003;11(10):1178–1181. doi: 10.1038/oby.2003.162. [DOI] [PubMed] [Google Scholar]
  111. Lantova P, Zub K, Koskela E, Sichova K, Borowski Z. Is there a linkage between metabolism and personality in small mammals? The root vole (Microtus oeconomus) example. Physiol. Behav. 2011;104(3):378–383. doi: 10.1016/j.physbeh.2011.04.017. [DOI] [PubMed] [Google Scholar]
  112. Lathe R. Hormones and the hippocampus. J. Endocrinol. 2001;169(2):205–231. doi: 10.1677/joe.0.1690205. [DOI] [PubMed] [Google Scholar]
  113. Lattanzio SB, Eikelboom R. Wheel access duration in rats: I. Effects on feeding and running. Behav. Neurosci. 2003;117(3):496–504. doi: 10.1037/0735-7044.117.3.496. [DOI] [PubMed] [Google Scholar]
  114. Lawrence VJ, Kopelman PG. Medical consequences of obesity. Clin. Dermatol. 2004;22(4):296–302. doi: 10.1016/j.clindermatol.2004.01.012. [DOI] [PubMed] [Google Scholar]
  115. Lerman I, Harrison BC, Freeman K, Hewett TE, Allen DL, Robbins J, Leinwand LA. Genetic variability in forced and voluntary endurance exercise performance in seven inbred mouse strains. J. Appl. Physiol. 2002;92(6):2245–2255. doi: 10.1152/japplphysiol.01045.2001. [DOI] [PubMed] [Google Scholar]
  116. Lett BT, Grant VL, Koh MT. Naloxone attenuates the conditioned place preference induced by wheel running in rats. Physiol. Behav. 2001;72(3):355–358. doi: 10.1016/s0031-9384(00)00427-3. [DOI] [PubMed] [Google Scholar]
  117. Lett BT, Grant VL, Koh MT, Flynn G. Prior experience with wheel running produces cross-tolerance to the rewarding effect of morphine. Pharmacol. Biochem. Behav. 2002;72(1–2):101–105. doi: 10.1016/s0091-3057(01)00722-5. [DOI] [PubMed] [Google Scholar]
  118. Levin BE. Spontaneous motor activity during the development and maintenance of diet-induced obesity in the rat. Physiol. Behav. 1991;50(3):573–581. doi: 10.1016/0031-9384(91)90548-3. [DOI] [PubMed] [Google Scholar]
  119. Levine JA, Eberhardt NL, Jensen MD. Role of nonexercise activity thermogenesis in resistance to fat gain in humans. Science. 1999;283(5399):212–214. doi: 10.1126/science.283.5399.212. [DOI] [PubMed] [Google Scholar]
  120. Levine JA, Lanningham-Foster LM, McCrady SK, Krizan AC, Olson LR, Kane PH, Jensen MD, Clark MM. Interindividual variation in posture allocation: possible role in human obesity. Science. 2005;307(5709):584–586. doi: 10.1126/science.1106561. [DOI] [PubMed] [Google Scholar]
  121. Lewis DE, Shellard L, Koeslag DG, Boer DE, McCarthy HD, McKibbin PE, Russell JC, Williams G. Intense exercise and food restriction cause similar hypothalamic neuropeptide Y increases in rats. Am. J. Physiol. 1993;264(2 (Pt 1)):E279–E284. doi: 10.1152/ajpendo.1993.264.2.E279. [DOI] [PubMed] [Google Scholar]
  122. Liu J, Garza JC, Bronner J, Kim CS, Zhang W, Lu XY. Acute administration of leptin produces anxiolytic-like effects: a comparison with fluoxetine. Psychopharmacology (Berl.) 2010;207(4):535–545. doi: 10.1007/s00213-009-1684-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Lou SJ, Liu JY, Chang H, Chen PJ. Hippocampal neurogenesis and gene expression depend on exercise intensity in juvenile rats. Brain Res. 2008;1210:48–55. doi: 10.1016/j.brainres.2008.02.080. [DOI] [PubMed] [Google Scholar]
  124. Lu XY, Kim CS, Frazer A, Zhang W. Leptin: a potential novel antidepressant. Proc. Natl. Acad. Sci. U.S.A. 2006;103(5):1593–1598. doi: 10.1073/pnas.0508901103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Lutter M, Krishnan V, Russo SJ, Jung S, McClung CA, Nestler EJ. Orexin signaling mediates the antidepressant-like effect of calorie restriction. J. Neurosci. 2008a;28(12):3071–3075. doi: 10.1523/JNEUROSCI.5584-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Lutter M, Sakata I, Osborne-Lawrence S, Rovinsky SA, Anderson JG, Jung S, Birnbaum S, Yanagisawa M, Elmquist JK, Nestler EJ, Zigman JM. The orexigenic hormone ghrelin defends against depressive symptoms of chronic stress. Nat. Neurosci. 2008b;11(7):752–753. doi: 10.1038/nn.2139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Lynn SE, Breuner CW, Wingfield JC. Short-term fasting affects locomotor activity, corticosterone, and corticosterone binding globulin in a migratory songbird. Horm. Behav. 2003;43(1):150–157. doi: 10.1016/s0018-506x(02)00023-5. [DOI] [PubMed] [Google Scholar]
  128. Malisch JL, Breuner CW, Gomes FR, Chappell MA, Garland T., Jr. Circadian pattern of total and free corticosterone concentrations, corticosteroid-binding globulin, and physical activity in mice selectively bred for high voluntary wheel-running behavior. Gen. Comp. Endocrinol. 2008;156(2):210–217. doi: 10.1016/j.ygcen.2008.01.020. [DOI] [PubMed] [Google Scholar]
  129. Malisch JL, Breuner CW, Kolb EM, Wada H, Hannon RM, Chappell MA, Middleton KM, Garland T., Jr. Behavioral despair and home-cage activity in mice with chronically elevated baseline corticosterone concentrations. Behav. Genet. 2009;39(2):192–201. doi: 10.1007/s10519-008-9246-8. [DOI] [PubMed] [Google Scholar]
  130. Mallei A, Shi B, Mocchetti I. Antidepressant treatments induce the expression of basic fibroblast growth factor in cortical and hippocampal neurons. Mol. Pharmacol. 2002;61(5):1017–1024. doi: 10.1124/mol.61.5.1017. [DOI] [PubMed] [Google Scholar]
  131. Mathes WF, Kanarek RB. Wheel running attenuates the antinociceptive properties of morphine and its metabolite, morphine-6-glucuronide, in rats. Physiol. Behav. 2001;74(1–2):245–251. doi: 10.1016/s0031-9384(01)00577-7. [DOI] [PubMed] [Google Scholar]
  132. Mathes WF, Kanarek RB. Chronic running wheel activity attenuates the antinociceptive actions of morphine and morphine-6-glucouronide administration into the periaqueductal gray in rats. Pharmacol. Biochem. Behav. 2006;83(4):578–584. doi: 10.1016/j.pbb.2006.03.020. [DOI] [PubMed] [Google Scholar]
  133. Mathes WF, Nehrenberg DL, Gordon R, Hua K, Garland T, Jr., Pomp D. Dopaminergic dysregulation in mice selectively bred for excessive exercise or obesity. Behav. Brain Res. 2010;210(2):155–163. doi: 10.1016/j.bbr.2010.02.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. McDermott C, Kelly JP. Comparison of the behavioural pharmacology of the Lister-Hooded with 2 commonly utilised albino rat strains. Prog. Neuropsychopharmacol. Biol. Psychiatry. 2008;32(8):1816–1823. doi: 10.1016/j.pnpbp.2008.08.004. [DOI] [PubMed] [Google Scholar]
  135. Meeusen R. Exercise and the brain: insight in new therapeutic modalities. Ann. Transplant. 2005;10(4):49–51. [PubMed] [Google Scholar]
  136. Meeusen R, De Meirleir K. Exercise and brain neurotransmission. Sports Med. 1995;20(3):160–188. doi: 10.2165/00007256-199520030-00004. [DOI] [PubMed] [Google Scholar]
  137. Moraska A, Fleshner M. Voluntary physical activity prevents stress-induced behavioral depression and anti-KLH antibody suppression. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2001;281(2):R484–R489. doi: 10.1152/ajpregu.2001.281.2.R484. [DOI] [PubMed] [Google Scholar]
  138. Morgan MA, Pfaff DW. Effects of estrogen on activity and fear-related behaviors in mice. Horm. Behav. 2001;40(4):472–482. doi: 10.1006/hbeh.2001.1716. [DOI] [PubMed] [Google Scholar]
  139. Morgan MA, Pfaff DW. Estrogen's effects on activity, anxiety, and fear in two mouse strains. Behav. Brain Res. 2002;132(1):85–93. doi: 10.1016/s0166-4328(01)00398-9. [DOI] [PubMed] [Google Scholar]
  140. Morishima-Yamato M, Hisaoka F, Shinomiya S, Harada N, Matoba H, Takahashi A, Nakaya Y. Cloning and establishment of a line of rats for high levels of voluntary wheel running. Life. Sci. 2005;77(5):551–561. doi: 10.1016/j.lfs.2004.10.074. [DOI] [PubMed] [Google Scholar]
  141. Morse AD, Russell JC, Hunt TW, Wood GO, Epling WF, Pierce WD. Diurnal variation of intensive running in food-deprived rats. Can. J. Physiol. Pharmacol. 1995;73(10):1519–1523. doi: 10.1139/y95-210. [DOI] [PubMed] [Google Scholar]
  142. Muller AP, Gnoatto J, Moreira JD, Zimmer ER, Haas CB, Lulhier F, Perry ML, Souza DO, Torres-Aleman I, Portela LV. Exercise increases insulin signaling in the hippocampus: physiological effects and pharmacological impact of intracerebroventricular insulin administration in mice. Hippocampus. 2011:1082–1092. doi: 10.1002/hipo.20822. [DOI] [PubMed] [Google Scholar]
  143. Murray PS, Groves JL, Pettett BJ, Britton SL, Koch LG, Dishman RK, Holmes PV. Locus coeruleus galanin expression is enhanced after exercise in rats selectively bred for high capacity for aerobic activity. Peptides. 2010:2264–2268. doi: 10.1016/j.peptides.2010.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Nestler EJ, Barrot M, Self DW. DeltaFosB: a sustained molecular switch for addiction. Proc. Natl. Acad. Sci. U.S.A. 2001;98(20):11042–11046. doi: 10.1073/pnas.191352698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Novak CM, Escande C, Burghardt PR, Zhang M, Barbosa MT, Chini EN, Britton SL, Koch LG, Akil H, Levine JA. Spontaneous activity, economy of activity, and resistance to diet-induced obesity in rats bred for high intrinsic aerobic capacity. Horm. Behav. 2010;58(3):355–367. doi: 10.1016/j.yhbeh.2010.03.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Novak CM, Escande C, Gerber SM, Chini EN, Zhang M, Britton SL, Koch LG, Levine JA. Endurance capacity, not body size, determines physical activity levels: role of skeletal muscle PEPCK. PLoS ONE. 2009;4(6):e5869. doi: 10.1371/journal.pone.0005869. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Novak CM, Jiang X, Wang C, Teske JA, Kotz CM, Levine JA. Caloric restriction and physical activity in zebrafish (Danio rerio) Neurosci. Lett. 2005;383(1–2):99–104. doi: 10.1016/j.neulet.2005.03.048. [DOI] [PubMed] [Google Scholar]
  148. Novak CM, Kotz CM, Levine JA. Central orexin sensitivity, physical activity, and obesity in diet-induced obese and diet-resistant rats. Am. J. Physiol. Endocrinol. Metab. 2006;290(2):E396–E403. doi: 10.1152/ajpendo.00293.2005. [DOI] [PubMed] [Google Scholar]
  149. Novak CM, Levine JA. Central neural and endocrine mechanisms of non-exercise activity thermogenesis and their potential impact on obesity. J. Neuroendocrinol. 2007;19(12):923–940. doi: 10.1111/j.1365-2826.2007.01606.x. [DOI] [PubMed] [Google Scholar]
  150. Ogawa S, Chan J, Gustafsson JA, Korach KS, Pfaff DW. Estrogen increases locomotor activity in mice through estrogen receptor alpha: specificity for the type of activity. Endocrinology. 2003;144(1):230–239. doi: 10.1210/en.2002-220519. [DOI] [PubMed] [Google Scholar]
  151. Oomura Y, Sasaki K, Suzuki K, Muto T, Li AJ, Ogita Z, Hanai K, Tooyama I, Kimura H, Yanaihara N. A new brain glucosensor and its physiological significance. Am. J. Clin. Nutr. 1992;55(Suppl. 1):278S–282S. doi: 10.1093/ajcn/55.1.278s. [DOI] [PubMed] [Google Scholar]
  152. Overstreet DH, Rezvani AH. Behavioral differences between two inbred strains of Fawn-Hooded rat: a model of serotonin dysfunction. Psychopharmacology (Berl.) 1996;128(3):328–330. doi: 10.1007/s002130050141. [DOI] [PubMed] [Google Scholar]
  153. Overton JM, Williams TD. Behavioral and physiologic responses to caloric restriction in mice. Physiol. Behav. 2004;81(5):749–754. doi: 10.1016/j.physbeh.2004.04.025. [DOI] [PubMed] [Google Scholar]
  154. Ozburn AR, Harris RA, Blednov YA. Wheel running, voluntary ethanol consumption, and hedonic substitution. Alcohol. 2008;42(5):417–424. doi: 10.1016/j.alcohol.2008.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Patterson CM, Levin BE. Role of exercise in the central regulation of energy homeostasis and in the prevention of obesity. Neuroendocrinology. 2008;87(2):65–70. doi: 10.1159/000100982. [DOI] [PubMed] [Google Scholar]
  156. Pereira AC, Huddleston DE, Brickman AM, Sosunov AA, Hen R, McKhann GM, Sloan R, Gage FH, Brown TR, Small SA. An in vivo correlate of exercise-induced neurogenesis in the adult dentate gyrus. Proc. Natl. Acad. Sci. U.S.A. 2007;104(13):5638–5643. doi: 10.1073/pnas.0611721104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Persson AI, Naylor AS, Jonsdottir IH, Nyberg F, Eriksson PS, Thorlin T. Differential regulation of hippocampal progenitor proliferation by opioid receptor antagonists in running and non-running spontaneously hypertensive rats. Eur. J. Neurosci. 2004;19(7):1847–1855. doi: 10.1111/j.1460-9568.2004.03268.x. [DOI] [PubMed] [Google Scholar]
  158. Pham TM, Brene S, Baumans V. Behavioral assessment of intermittent wheel running and individual housing in mice in the laboratory. J. Appl. Anim. Welf. Sci. 2005;8(3):157–173. doi: 10.1207/s15327604jaws0803_1. [DOI] [PubMed] [Google Scholar]
  159. Pierce WD, Epling WF, Boer DP. Deprivation and satiation: the interrelations between food and wheel running. J. Exp. Anal. Behav. 1986;46(2):199–210. doi: 10.1901/jeab.1986.46-199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Pirke KM, Broocks A, Wilckens T, Marquard R, Schweiger U. Starvation-induced hyperactivity in the rat: the role of endocrine and neurotransmitter changes. Neurosci. Biobehav. Rev. 1993;17(3):287–294. doi: 10.1016/s0149-7634(05)80012-0. [DOI] [PubMed] [Google Scholar]
  161. Pistilli EE, Bogdanovich S, Garton F, Yang N, Gulbin JP, Conner JD, Anderson BG, Quinn LS, North K, Ahima RS, Khurana TS. Loss of IL-15 receptor alpha alters the endurance, fatigability, and metabolic characteristics of mouse fast skeletal muscles. J. Clin. Invest. 2011;121(8):3120–3132. doi: 10.1172/JCI44945. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Redlin U, Mrosovsky N. Nocturnal activity in a diurnal rodent (Arvicanthis niloticus): the importance of masking. J. Biol. Rhythms. 2004;19(1):58–67. doi: 10.1177/0748730403260371. [DOI] [PubMed] [Google Scholar]
  163. Rezende EL, Chappell MA, Gomes FR, Malisch JL, Garland T., Jr. Maximal metabolic rates during voluntary exercise, forced exercise, and cold exposure in house mice selectively bred for high wheel-running. J. Exp. Biol. 2005;208(Pt 12):2447–2458. doi: 10.1242/jeb.01631. [DOI] [PubMed] [Google Scholar]
  164. Rezende EL, Kelly SA, Gomes FR, Chappell MA, Garland T., Jr. Effects of size, sex, and voluntary running speeds on costs of locomotion in lines of laboratory mice selectively bred for high wheel-running activity. Physiol. Biochem. Zool. 2006;79(1):83–99. doi: 10.1086/498187. [DOI] [PubMed] [Google Scholar]
  165. Richter H, Ambree O, Lewejohann L, Herring A, Keyvani K, Paulus W, Palme R, Touma C, Schabitz WR, Sachser N. Wheel-running in a transgenic mouse model of Alzheimer's disease: protection or symptom? Behav. Brain Res. 2008;190(1):74–84. doi: 10.1016/j.bbr.2008.02.005. [DOI] [PubMed] [Google Scholar]
  166. Rikke BA, Yerg JE, 3rd, Battaglia ME, Nagy TR, Allison DB, Johnson TE. Strain variation in the response of body temperature to dietary restriction. Mech. Ageing Dev. 2003;124(5):663–678. doi: 10.1016/s0047-6374(03)00003-4. [DOI] [PubMed] [Google Scholar]
  167. Robinson TE. Neuroscience. Addicted rats. Science. 2004;305(5686):951–953. doi: 10.1126/science.1102496. [DOI] [PubMed] [Google Scholar]
  168. Routtenberg A, Kuznesof AW. Self-starvation of rats living in activity wheels on a restricted feeding schedule. J. Comp. Physiol. Psychol. 1967;64(3):414–421. doi: 10.1037/h0025205. [DOI] [PubMed] [Google Scholar]
  169. Rozin P, Dow S. What causes humans to begin and end a meal? A role for memory for what has been eaten, as evidenced by a study of multiple meal eating in amnesic patients. Psychol. Sci. 1998;9(5):392–396. [Google Scholar]
  170. Russell JC, Epling WF, Pierce D, Amy RM, Boer DP. Induction of voluntary prolonged running by rats. J. Appl. Physiol. 1987;63(6):2549–2553. doi: 10.1152/jappl.1987.63.6.2549. [DOI] [PubMed] [Google Scholar]
  171. Russo-Neustadt A, Beard RC, Cotman CW. Exercise, antidepressant medications, and enhanced brain derived neurotrophic factor expression. Neuropsychopharmacology. 1999;21(5):679–682. doi: 10.1016/S0893-133X(99)00059-7. [DOI] [PubMed] [Google Scholar]
  172. Satvat E, Eikelboom R. Dissociation of conditioned and unconditioned factors in the running-induced feeding suppression. Physiol. Behav. 2006;89(3):428–437. doi: 10.1016/j.physbeh.2006.07.005. [DOI] [PubMed] [Google Scholar]
  173. Scheurink AJ, Boersma GJ, Nergardh R, Sodersten P. Neurobiology of hyperactivity and reward: agreeable restlessness in anorexia nervosa. Physiol. Behav. 2010;100(5):490–495. doi: 10.1016/j.physbeh.2010.03.016. [DOI] [PubMed] [Google Scholar]
  174. Schmelzeis MC, Mittleman G. The hippocampus and reward: effects of hippocampal lesions on progressive-ratio responding. Behav. Neurosci. 1996;110(5):1049–1066. doi: 10.1037//0735-7044.110.5.1049. [DOI] [PubMed] [Google Scholar]
  175. Schmidt S, Gawlik V, Holter SM, Augustin R, Scheepers A, Behrens M, Wurst W, Gailus-Durner V, Fuchs H, Hrabe de Angelis M, Kluge R, Joost HG, Schurmann A. Deletion of glucose transporter GLUT8 in mice increases locomotor activity. Behav. Genet. 2008;38(4):396–406. doi: 10.1007/s10519-008-9208-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Self DW. Neural substrates of drug craving and relapse in drug addiction. Ann. Med. 1998;30(4):379–389. doi: 10.3109/07853899809029938. [DOI] [PubMed] [Google Scholar]
  177. Severinsen T, Munch IC. Body core temperature during food restriction in rats. Acta Physiol. Scand. 1999;165(3):299–305. doi: 10.1046/j.1365-201x.1999.00488.x. [DOI] [PubMed] [Google Scholar]
  178. Sherwin CM. Voluntary wheel running: a review and novel interpretation. Anim. Behav. 1998;56(1):11–27. doi: 10.1006/anbe.1998.0836. [DOI] [PubMed] [Google Scholar]
  179. Simoncic M, Horvat S, Stevenson PL, Bunger L, Holmes MC, Kenyon CJ, Speak-man JR, Morton NM. Divergent physical activity and novel alternative responses to high fat feeding in polygenic fat and lean mice. Behav. Genet. 2008;38(3):292–300. doi: 10.1007/s10519-008-9199-y. [DOI] [PubMed] [Google Scholar]
  180. Sisson SB, Camhi SM, Church TS, Martin CK, Tudor-Locke C, Bouchard C, Earnest CP, Smith SR, Newton RL, Jr, Rankinen T, Katzmarzyk PT. Leisure time sedentary behavior, occupational/domestic physical activity, and metabolic syndrome in U.S. men and women. Metab. Syndr. Relat. Disord. 2009;7(6):529–536. doi: 10.1089/met.2009.0023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Smith MA, Schmidt KT, Iordanou JC, Mustroph ML. Aerobic exercise decreases the positive-reinforcing effects of cocaine. Drug Alcohol Depend. 2008;98(1–2):129–135. doi: 10.1016/j.drugalcdep.2008.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Smith SL, Rasmussen EB. Effects of 2-AG on the reinforcing properties of wheel activity in obese and lean Zucker rats. Behav. Pharmacol. 2010;21(4):292–300. doi: 10.1097/FBP.0b013e32833aec4d. [DOI] [PubMed] [Google Scholar]
  183. Solberg LC, Horton TH, Turek FW. Circadian rhythms and depression: effects of exercise in an animal model. Am. J. Physiol. 1999;276(1 (Pt 2))):R152–R161. doi: 10.1152/ajpregu.1999.276.1.R152. [DOI] [PubMed] [Google Scholar]
  184. Sothmann MS, Buckworth J, Claytor RP, Cox RH, White-Welkley JE, Dishman RK. Exercise training and the cross-stressor adaptation hypothesis. Exerc. Sport Sci. Rev. 1996;24:267–287. [PubMed] [Google Scholar]
  185. Spigelman MN, McLeod WS, Rockman GE. Caloric vs. pharmacologic effects of ethanol consumption on activity anorexia in rats. Pharmacol. Biochem. Behav. 1991;39(1):85–90. doi: 10.1016/0091-3057(91)90401-m. [DOI] [PubMed] [Google Scholar]
  186. Stamatakis E, Hamer M, Dunstan DW. Screen-based entertainment time, all-cause mortality, and cardiovascular events: population-based study with ongoing mortality and hospital events follow-up. J. Am. Coll. Cardiol. 2011;57(3):292–299. doi: 10.1016/j.jacc.2010.05.065. [DOI] [PubMed] [Google Scholar]
  187. Stead JD, Clinton S, Neal C, Schneider J, Jama A, Miller S, Vazquez DM, Watson SJ, Akil H. Selective breeding for divergence in novelty-seeking traits: heritability and enrichment in spontaneous anxiety-related behaviors. Behav. Genet. 2006;36(5):697–712. doi: 10.1007/s10519-006-9058-7. [DOI] [PubMed] [Google Scholar]
  188. Stephan FK. The other circadian system: food as a Zeitgeber. J. Biol. Rhythms. 2002;17(4):284–292. doi: 10.1177/074873040201700402. [DOI] [PubMed] [Google Scholar]
  189. Stephens BR, Granados K, Zderic TW, Hamilton MT, Braun B. Effects of 1 day of inactivity on insulin action in healthy men and women: interaction with energy intake. Metabolism. 2011;60(7):941–949. doi: 10.1016/j.metabol.2010.08.014. [DOI] [PubMed] [Google Scholar]
  190. Stranahan AM, Khalil D, Gould E. Social isolation delays the positive effects of running on adult neurogenesis. Nat. Neurosci. 2006;9(4):526–533. doi: 10.1038/nn1668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Stranahan AM, Lee K, Mattson MP. Central mechanisms of HPA axis regulation by voluntary exercise. Neuromol. Med. 2008;10(2):118–127. doi: 10.1007/s12017-008-8027-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Sullivan M, Karlsson J, Sjostrom L, Backman L, Bengtsson C, Bouchard C, Dahlgren S, Jonsson E, Larsson B, Lindstedt S, et al. Swedish obese subjects (SOS) – an intervention study of obesity. Baseline evaluation of health and psychosocial functioning in the first 1743 subjects examined. Int. J. Obes. Relat. Metab. Disord. 1993;17(9):503–512. [PubMed] [Google Scholar]
  193. Swallow JG, Carter PA, Garland T., Jr. Artificial selection for increased wheel-running behavior in house mice. Behav. Genet. 1998;28(3):227–237. doi: 10.1023/a:1021479331779. [DOI] [PubMed] [Google Scholar]
  194. Swallow JG, Koteja P, Carter PA, Garland T. Artificial selection for increased wheel-running activity in house mice results in decreased body mass at maturity. J. Exp. Biol. 1999;202(Pt 18):2513L–2520. doi: 10.1242/jeb.202.18.2513. [DOI] [PubMed] [Google Scholar]
  195. Swallow JG, Koteja P, Carter PA, Garland T., Jr. Food consumption and body composition in mice selected for high wheel-running activity. J. Comp. Physiol. [B] 2001;171(8):651–659. doi: 10.1007/s003600100216. [DOI] [PubMed] [Google Scholar]
  196. Tang BL. Leptin as a neuroprotective agent. Biochem. Biophys. Res. Commun. 2008;368(2):181–185. doi: 10.1016/j.bbrc.2008.01.063. [DOI] [PubMed] [Google Scholar]
  197. Teegarden SL, Nestler EJ, Bale TL. Delta FosB-mediated alterations in dopamine signaling are normalized by a palatable high-fat diet. Biol. Psychiatry. 2008;64(11):941–950. doi: 10.1016/j.biopsych.2008.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Tejani-Butt S, Kluczynski J, Pare WP. Strain-dependent modification of behavior following antidepressant treatment. Prog. Neuropsychopharmacol. Biol. Psychiatry. 2003;27(1):7–14. doi: 10.1016/s0278-5846(02)00308-1. [DOI] [PubMed] [Google Scholar]
  199. Teske JA, Levine AS, Kuskowski M, Levine JA, Kotz CM. Elevated hypothalamic orexin signaling, sensitivity to orexin A and spontaneous physical activity in obesity resistant rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2006;294(4):R889–R899. doi: 10.1152/ajpregu.00536.2005. [DOI] [PubMed] [Google Scholar]
  200. Turner CA, Lewis MH. Environmental enrichment: effects on stereotyped behavior and neurotrophin levels. Physiol. Behav. 2003;80(2–3):259–266. doi: 10.1016/j.physbeh.2003.07.008. [DOI] [PubMed] [Google Scholar]
  201. Vaanholt LM, Jonas I, Doornbos M, Schubert KA, Nyakas C, Garland T, Jr., Visser GH, van Dijk G. Metabolic and behavioral responses to high-fat feeding in mice selectively bred for high wheel-running activity. Int. J. Obes. (Lond.) 2008;32(10):1566–1575. doi: 10.1038/ijo.2008.136. [DOI] [PubMed] [Google Scholar]
  202. van Baak MA, van Mil E, Astrup AV, Finer N, Van Gaal LF, Hilsted J, Kopelman PG, Rossner S, James WP, Saris WH. Leisure-time activity is an important determinant of long-term weight maintenance after weight loss in the Sibutramine Trial on Obesity Reduction and Maintenance (STORM trial) Am. J. Clin. Nutr. 2003;78(2):209–214. doi: 10.1093/ajcn/78.2.209. [DOI] [PubMed] [Google Scholar]
  203. Van der Borght K, Kobor-Nyakas DE, Klauke K, Eggen BJ, Nyakas C, Van der Zee EA, Meerlo P. Physical exercise leads to rapid adaptations in hippocampal vasculature: temporal dynamics and relationship to cell proliferation and neurogenesis. Hippocampus. 2009;19(10):928–936. doi: 10.1002/hipo.20545. [DOI] [PubMed] [Google Scholar]
  204. Van Hoomissen JD, Holmes PV, Zellner AS, Poudevigne A, Dishman RK. Effects of beta-adrenoreceptor blockade during chronic exercise on contextual fear conditioning and mRNA for galanin and brain-derived neurotrophic factor. Behav. Neurosci. 2004;118(6):1378–1390. doi: 10.1037/0735-7044.118.6.1378. [DOI] [PubMed] [Google Scholar]
  205. van Praag H, Kempermann G, Gage FH. Running increases cell proliferation and neurogenesis in the adult mouse dentate gyrus. Nat. Neurosci. 1999;2(3):266–270. doi: 10.1038/6368. [DOI] [PubMed] [Google Scholar]
  206. Vargas-Perez H, Mena-Segovia J, Giordano M, Diaz JL. Induction of c-fos in nucleus accumbens in naive male Balb/c mice after wheel running. Neurosci. Lett. 2003;352(2):81–84. doi: 10.1016/j.neulet.2003.08.073. [DOI] [PubMed] [Google Scholar]
  207. Vargas-Perez H, Borrelli E, Diaz JL. Wheel running use in dopamine D2L receptor knockout mice. Neurosci. Lett. 2004;366(2):172–175. doi: 10.1016/j.neulet.2004.05.029. [DOI] [PubMed] [Google Scholar]
  208. Vargas-Perez H, Sellings LH, Paredes RG, Prado-Alcala RA, Diaz JL. Reinforcement of wheel running in BALB/c mice: role of motor activity and endogenous opioids. J. Mot. Behav. 2008;40(6):587–593. doi: 10.3200/JMBR.40.6.587-593. [DOI] [PubMed] [Google Scholar]
  209. Wang C, Bomberg E, Billington C, Levine A, Kotz CM. Brain-derived neurotrophic factor in the hypothalamic paraventricular nucleus increases energy expenditure by elevating metabolic rate. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2007a;293(3):R992–R1002. doi: 10.1152/ajpregu.00516.2006. [DOI] [PubMed] [Google Scholar]
  210. Wang C, Bomberg E, Billington C, Levine A, Kotz CM. Brain-derived neurotrophic factor in the hypothalamic paraventricular nucleus reduces energy intake. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2007b;293(3):R1003–R1012. doi: 10.1152/ajpregu.00011.2007. [DOI] [PubMed] [Google Scholar]
  211. Wang C, Bomberg E, Levine A, Billington C, Kotz CM. Brain-derived neurotrophic factor in the ventromedial nucleus of the hypothalamus reduces energy intake. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2007c;293(3):R1037–R1045. doi: 10.1152/ajpregu.00125.2007. [DOI] [PubMed] [Google Scholar]
  212. Waters RP, Renner KJ, Pringle RB, Summers CH, Britton SL, Koch LG, Swallow JG. Selection for aerobic capacity affects corticosterone, monoamines and wheel-running activity. Physiol. Behav. 2008;93(4–5):1044–1054. doi: 10.1016/j.physbeh.2008.01.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Weinsier RL, Hunter GR, Heini AF, Goran MI, Sell SM. The etiology of obesity: relative contribution of metabolic factors, diet, and physical activity. Am. J. Med. 1998;105(2):145–150. doi: 10.1016/s0002-9343(98)00190-9. [DOI] [PubMed] [Google Scholar]
  214. Werme M, Thoren P, Olson L, Brene S. Addiction-prone Lewis but not Fischer rats develop compulsive running that coincides with downregulation of nerve growth factor inducible-B and neuron-derived orphan receptor 1. J. Neurosci. 1999;19(14):6169–6174. doi: 10.1523/JNEUROSCI.19-14-06169.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Werme M, Thoren P, Olson L, Brene S. Running and cocaine both upregulate dynorphin mRNA in medial caudate putamen. Eur. J. Neurosci. 2000;12(8):2967–2974. doi: 10.1046/j.1460-9568.2000.00147.x. [DOI] [PubMed] [Google Scholar]
  216. Williams TD, Chambers JB, Henderson RP, Rashotte ME, Overton JM. Cardiovascular responses to caloric restriction and thermoneutrality in C57BL/6J mice. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2002;282(5):R1459–R1467. doi: 10.1152/ajpregu.00612.2001. [DOI] [PubMed] [Google Scholar]
  217. Wood RI. Oral testosterone self-administration in male hamsters: dose-response, voluntary exercise, and individual differences. Horm. Behav. 2002;41(3):247–258. doi: 10.1006/hbeh.2002.1769. [DOI] [PubMed] [Google Scholar]
  218. Zheng H, Lenard NR, Shin AC, Berthoud HR. Appetite control and energy balance regulation in the modern world: reward-driven brain overrides repletion signals. Int. J. Obes. (Lond.) 2009;33(Suppl. 2):S8–S13. doi: 10.1038/ijo.2009.65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Zheng H, Liu Y, Li W, Yang B, Chen D, Wang X, Jiang Z, Wang H, Wang Z, Cornelisson G, Halberg F. Beneficial effects of exercise and its molecular mechanisms on depression in rats. Behav. Brain Res. 2006;168(1):47–55. doi: 10.1016/j.bbr.2005.10.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Zombeck JA, Deyoung EK, Brzezinska WJ, Rhodes JS. Selective breeding for increased home cage physical activity in collaborative cross and Hsd:ICR mice. Behav. Genet. 2011;41(4):571–582. doi: 10.1007/s10519-010-9425-2. [DOI] [PubMed] [Google Scholar]

RESOURCES