Top Banner
Review Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance Jere Lindén a,b , Sanna Lensu c,d , Jouko Tuomisto c , Raimo Pohjanvirta a, * a Department of Food Hygiene and Environmental Health, Faculty of Veterinary Medicine, University of Helsinki, P.O. Box 66, FI-00014 University of Helsinki, Finland b Department of Veterinary Biosciences, Faculty of Veterinary Medicine, University of Helsinki, P.O. Box 66, FI-00014 University of Helsinki, Finland c Department of Environmental Health, National Institute for Health and Welfare, P.O. Box 95, FI-70701 Kuopio, Finland d Faculty of Health Sciences, The School of Pharmacy, University of Eastern Finland, P.O. Box 1611, FI-70211 Kuopio, Finland article info Article history: Available online 17 July 2010 Keywords: Dioxins TCDD 2,3,7,8-tetrachlorodibenzo-p-dioxin Aryl hydrocarbon receptor Food intake Energy balance Body weight Wasting syndrome Central nervous system abstract Dioxins are ubiquitous environmental contaminants that have attracted toxicological interest not only for the potential risk they pose to human health but also because of their unique mechanism of action. This mechanism involves a specific, phylogenetically old intracellular receptor (the aryl hydrocarbon receptor, AHR) which has recently proven to have an integral regulatory role in a number of physiological pro- cesses, but whose endogenous ligand is still elusive. A major acute impact of dioxins in laboratory animals is the wasting syndrome, which represents a puzzling and dramatic perturbation of the regulatory systems for energy balance. A single dose of the most potent dioxin, TCDD, can permanently readjust the defended body weight set-point level thus providing a potentially useful tool and model for physiological research. Recent evidence of response-selective modulation of AHR action by alternative ligands suggests further that even therapeutic implications might be possible in the future. Ó 2010 Elsevier Inc. All rights reserved. 1. Introduction Dioxins constitute a toxicologically and endocrinologically intriguing group of chemicals because they affect most endocrine organs in laboratory animals. An especially tantalizing and scientif- ically challenging impact highly characteristic of these compounds – but a notably rare outcome of chemical exposure in general – is severely and perpetually derailed regulation of energy homeostasis and body weight whose manifestations range from inhibition of further growth to a dramatic body weight loss (dubbed the wasting syndrome). In this review, we will present new behavioral and bio- chemical findings from dioxin-exposed animals in the context of current concepts on the central regulatory systems for energy bal- ance. We will first introduce dioxins and briefly outline their major toxic impacts in laboratory animals and humans (Section 2). We will then deal with their molecular mechanisms of action and the crucial role played therein by a specific receptor protein (aryl hydrocarbon receptor; AHR or ‘‘dioxin receptor”) and its main sig- naling pathway (Section 3). Evidence demonstrating the existence of this signaling pathway in a fully functional composition in the central nervous system (CNS) will be provided next (Section 4). The final Section 5 will focus on alterations caused by the most po- tent dioxin, TCDD, in consummatory behaviors and energy balance with special reference to the wasting syndrome. The biochemical basis of these alterations will be pursued by elaborating on the interference of TCDD with the networks, circuits and neurochemi- cals underlying the central regulation of energy balance, of which a succinct account will also be given. 2. Dioxins as environmental and food contaminants ‘‘Dioxins” are perhaps the most feared group among the persis- tent organic pollutants (POPs). In fact, it is an inaccurate term for relatively loosely related group of various chemicals, including polychlorinated dibenzo-p-dioxins (PCDDs), polychlorinated dib- enzofurans (PCDFs), and dioxin-like polychlorinated biphenyls (PCBs). Toxicologically relevant congeners of these have been given a toxicity equivalence factor (TEF value) [484] related to the most toxic dioxin, 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD or simply ‘‘dioxin”). In addition, respective polybrominated compounds exert dioxin-like activity, as well as some more distant chemicals such as polychlorinated azoxy-benzenes. The potency of these compounds varies widely, as does the sensitivity of animal species and even strains or substrains to their toxicity. Furthermore, their toxicity is qualitatively species-dependent [338]. This makes their risk assessment very challenging. There have been a few dramatic cases of human exposure due to occupational exposure, accidents, or deliberate poisoning. The most recent ones are illustrative of the challenges these com- pounds represent to risk assessment. In 1998 two adult women were poisoned at their workplace in Vienna by huge doses of TCDD 0091-3022/$ - see front matter Ó 2010 Elsevier Inc. All rights reserved. doi:10.1016/j.yfrne.2010.07.002 * Corresponding author. Fax: +358 9 19157170. E-mail address: raimo.pohjanvirta@helsinki.fi (R. Pohjanvirta). Frontiers in Neuroendocrinology 31 (2010) 452–478 Contents lists available at ScienceDirect Frontiers in Neuroendocrinology journal homepage: www.elsevier.com/locate/yfrne
27

Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

Apr 21, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

Frontiers in Neuroendocrinology 31 (2010) 452–478

Contents lists available at ScienceDirect

Frontiers in Neuroendocrinology

journal homepage: www.elsevier .com/locate /yfrne

Review

Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

Jere Lindén a,b, Sanna Lensu c,d, Jouko Tuomisto c, Raimo Pohjanvirta a,*

a Department of Food Hygiene and Environmental Health, Faculty of Veterinary Medicine, University of Helsinki, P.O. Box 66, FI-00014 University of Helsinki, Finlandb Department of Veterinary Biosciences, Faculty of Veterinary Medicine, University of Helsinki, P.O. Box 66, FI-00014 University of Helsinki, Finlandc Department of Environmental Health, National Institute for Health and Welfare, P.O. Box 95, FI-70701 Kuopio, Finlandd Faculty of Health Sciences, The School of Pharmacy, University of Eastern Finland, P.O. Box 1611, FI-70211 Kuopio, Finland

a r t i c l e i n f o

Article history:Available online 17 July 2010

Keywords:DioxinsTCDD2,3,7,8-tetrachlorodibenzo-p-dioxinAryl hydrocarbon receptorFood intakeEnergy balanceBody weightWasting syndromeCentral nervous system

0091-3022/$ - see front matter � 2010 Elsevier Inc. Adoi:10.1016/j.yfrne.2010.07.002

* Corresponding author. Fax: +358 9 19157170.E-mail address: [email protected] (R. P

a b s t r a c t

Dioxins are ubiquitous environmental contaminants that have attracted toxicological interest not only forthe potential risk they pose to human health but also because of their unique mechanism of action. Thismechanism involves a specific, phylogenetically old intracellular receptor (the aryl hydrocarbon receptor,AHR) which has recently proven to have an integral regulatory role in a number of physiological pro-cesses, but whose endogenous ligand is still elusive. A major acute impact of dioxins in laboratoryanimals is the wasting syndrome, which represents a puzzling and dramatic perturbation of theregulatory systems for energy balance. A single dose of the most potent dioxin, TCDD, can permanentlyreadjust the defended body weight set-point level thus providing a potentially useful tool and model forphysiological research. Recent evidence of response-selective modulation of AHR action by alternativeligands suggests further that even therapeutic implications might be possible in the future.

� 2010 Elsevier Inc. All rights reserved.

1. Introduction

Dioxins constitute a toxicologically and endocrinologicallyintriguing group of chemicals because they affect most endocrineorgans in laboratory animals. An especially tantalizing and scientif-ically challenging impact highly characteristic of these compounds– but a notably rare outcome of chemical exposure in general – isseverely and perpetually derailed regulation of energy homeostasisand body weight whose manifestations range from inhibition offurther growth to a dramatic body weight loss (dubbed the wastingsyndrome). In this review, we will present new behavioral and bio-chemical findings from dioxin-exposed animals in the context ofcurrent concepts on the central regulatory systems for energy bal-ance. We will first introduce dioxins and briefly outline their majortoxic impacts in laboratory animals and humans (Section 2). Wewill then deal with their molecular mechanisms of action andthe crucial role played therein by a specific receptor protein (arylhydrocarbon receptor; AHR or ‘‘dioxin receptor”) and its main sig-naling pathway (Section 3). Evidence demonstrating the existenceof this signaling pathway in a fully functional composition in thecentral nervous system (CNS) will be provided next (Section 4).The final Section 5 will focus on alterations caused by the most po-tent dioxin, TCDD, in consummatory behaviors and energy balancewith special reference to the wasting syndrome. The biochemical

ll rights reserved.

ohjanvirta).

basis of these alterations will be pursued by elaborating on theinterference of TCDD with the networks, circuits and neurochemi-cals underlying the central regulation of energy balance, of which asuccinct account will also be given.

2. Dioxins as environmental and food contaminants

‘‘Dioxins” are perhaps the most feared group among the persis-tent organic pollutants (POPs). In fact, it is an inaccurate term forrelatively loosely related group of various chemicals, includingpolychlorinated dibenzo-p-dioxins (PCDDs), polychlorinated dib-enzofurans (PCDFs), and dioxin-like polychlorinated biphenyls(PCBs). Toxicologically relevant congeners of these have been givena toxicity equivalence factor (TEF value) [484] related to the mosttoxic dioxin, 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD or simply‘‘dioxin”). In addition, respective polybrominated compounds exertdioxin-like activity, as well as some more distant chemicals such aspolychlorinated azoxy-benzenes. The potency of these compoundsvaries widely, as does the sensitivity of animal species and evenstrains or substrains to their toxicity. Furthermore, their toxicityis qualitatively species-dependent [338]. This makes their riskassessment very challenging.

There have been a few dramatic cases of human exposure dueto occupational exposure, accidents, or deliberate poisoning. Themost recent ones are illustrative of the challenges these com-pounds represent to risk assessment. In 1998 two adult womenwere poisoned at their workplace in Vienna by huge doses of TCDD

Page 2: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 453

[143,144]. TCDD concentrations especially in one of them were ex-tremely high, in fact the highest ever measured, 144,000 pg/g inserum fat (equal to a single dose of about 25 lg/kg). She survived,but suffered from severe chloracne lasting for years. Other symp-toms included relatively mild gastrointestinal symptoms and sub-sequent amenorrhea, but abnormal laboratory findings were veryfew [143].

In 2004, then presidential candidate of Ukraine Victor Yush-chenko was deliberately poisoned with a huge dose of TCDD[436]. His TCDD concentration was 108,000 pg/g fat, and after ini-tial malaise and stomach pain the most remarkable symptom wasagain severe chloracne. These two cases indicate that humans arenot likely to be as sensitive as the most sensitive animal species,and that acute toxicity is not the most essential feature of dioxinadverse effects; rather the long-term effects may be of concern.This supports the notion after an industrial accident in Seveso,Italy, in 1976 where thousands of inhabitants were exposed toTCDD after a release of kilograms of TCDD from a pressure tank.High TCDD concentrations up to 56,000 pg/g fat were noted espe-cially in children obviously playing outside and eating local food.The acute effects were limited to about 200 cases of chloracne[280] (the human impacts of dioxins will be addressed in more de-tail in Section 2.6). Because of the unique mechanism of action ofTCDD, however, all aspects of dioxin toxicity are highly interestingin a scientific sense, as a tool to understand animal physiology andbiochemistry.

2.1. Chemistry

There are 75 possible congeners of polychlorinated dibenzo-p-dioxins (PCDD) and 135 possible congeners of polychlorinated dib-enzofurans (PCDF). However, only compounds with so called lat-eral chlorine substitutions at the positions 2,3,7, and 8 (Fig. 1)are specifically toxic, because others do not bind to the aryl hydro-carbon receptor (AHR) and/or are metabolized much faster thanthe 2,3,7,8-congeners. TEF values have been estimated for 17 ofthem (seven dibenzo-p-dioxins and 10 dibenzofurans) having 4–8 chlorine substitutes. Each chlorine substitute in excess of the four

Dibenzo-p-dioxin

10 9

8

7

5 64

3

2

1O

O

2,3,7,8-TCDD

O

O

ClCl

Cl Cl

2,3,4,7,8-PeCDF

O

Cl

Cl

Cl

ClCl

Fig. 1. Structures of dibenzo-p-dioxin, 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)and 2,3,4,7,8-pentachlorodibenzofuran (2,3,4,7,8-PeCDF).

(2,3,7, and 8) decreases the potency, but by and large the toxic ef-fects remain the same [484].

The 209 PCB-compounds can be divided to three groups (Fig. 2).The four non-ortho compounds having no chlorine substitute inany o-position to the inter-ring C–C-bridge (2,20,6 or 60) are the mostpotent as a group. Among them 3,30,4,40,5-penta-CB (PCB126) isclose to the most potent dioxins in its toxicity [484]. The eightmono-ortho PCBs have some activity (roughly thirty thousand timesweaker than TCDD), and all other PCBs are devoid of a noticeabledioxin-like effect. This is due to the requirement of binding to theAHR: only compounds able to assume a planar (flat) position are ableto bind to the receptor. Only non-ortho compounds are freely rotat-ing along the C–C-bridge, and each o-chlorine makes it more difficultfor the molecule to assume a planar conformation (Fig. 2).

Other groups of planar halogenated molecules have not been gi-ven TEF values, although e.g. some brominated compounds mightdeserve it [484].

2.2. Sources

The sources of PCBs are different from those of PCDD/PCDFs.PCB-compounds were technically excellent oils because they wereresistant to pressure, chemically stable, non-flammable, and elec-trically non-conductive. Therefore they were used from 1930s to1980s for multiple purposes, e.g. in heavy-load hydraulic equip-ment, heat exchangers, electrical equipment, and as componentsin plastic ware. Even after their adverse properties were notedand their production discontinued in most countries in the1980s, these compounds linger on in many products such as elec-trical transformers and plastic materials, and some of the residualPCB ends up to waste incineration or dumps, and then to the gen-eral environment.

PCDD and PCDF compounds were never synthesized except forscientific research, though they are unwanted synthesis side prod-ucts of many chemicals such as PCBs, chlorophenol fungicides andphenoxy acid herbicides. Their major sources are many, mostlydealing with burning processes [374,440]. Any burning will pro-duce dioxins, if chlorine is available, and especially so, if metal cat-alysts such as copper are present. Therefore poorly controlledurban waste incineration used to be one of the most importantsources. This can be technically solved by ensuring high incinera-tion temperature (1000 �C or more) and effective flue gas filtration,and in modern good-quality incinerators dioxins are not a problem[274]. In fact dumpsites as an alternative may be worse, becausesome dumpsite fires are almost unavoidable, and in poor burningconditions the production of dioxins may be high [104,247,390].

Chlorine bleaching of pulp is another example of dioxin produc-tion that has been successfully abated by improving technology.Likewise, most chemical syntheses prone to produce dioxin

Biphenyl

5’6’65

4’4

3’2’3 2

Cl

Cl

Cl

Cl

Cl

3,3’,4,4’,5-PCB

Fig. 2. Structures of biphenyl and 3,30 ,4,40 ,5-pentachlorinated biphenyl (PCB 126).

Page 3: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

454 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

impurities (production of PCBs, chlorophenols, phenoxy acids, etc.)have been discontinued or technology improved. Examples of theremaining dioxin sources aside from the poor incinerator plantsare steel industries and local burning of solid fuels [18,233,369]as well as recycling of electrical devices [262,500].

2.3. Fate in the environment

Both PCBs and PCDD/PCDFs are highly resistant compounds inall aspects: physical, chemical and biological. They are very slowlybroken down in the environment by ultraviolet radiation and bio-degradation [428]. In the dark they are stable for decades, maybeeven millennia. Because they are extremely poorly soluble to water[8,269], they tend to bioaccumulate to all organic materials, espe-cially to lipids, due to their high lipid–water partition coefficients[268]. There has been an active search for micro-organisms ableto break down PCBs and PCDD/PCDFs, but the success has as yetbeen limited [56,512]. Mammals metabolize these compoundsvery ineffectively and slowly [483]. Therefore they also biomagnifyin the food chain meaning that the higher an organism is in thefood chain, the higher is the concentration. This is clearly seen infish-eating birds and seals which easily collect harmful amountsto their bodies [220]. Humans are at less risk because their dietis not from a single source, even if they are at the top of the foodchain.

2.4. Human intake and levels

Almost all of PCB and dioxin intake is from food of animal ori-gin: milk, meat or fish predominate depending on the countryand the culture [251]. In many countries the daily intake of dioxinsand dioxin-like PCBs is of the order of 100 pg/day, i.e. 1–2 pg/kg/day, in a few countries higher (given as TEQ as are other valuesin this paragraph, see below). Since the half-lives are very long(for e.g. TCDD 7–8 years), the body burden will increase almostover the whole lifetime. Therefore the concentrations may increase5- to 10-fold from age 20 to age 60 [216,322]. In Western Europe,the highest body burdens were found in the 1970s and early 1980s[309], and the trends have been similar in the US [405].

One of the most useful measures of time trends of body burdens isconcentration in breast milk measured over decades [52,77,251,485].In many countries with relatively high exposures the concentrationshave decreased to about one tenth of those in 1970s, and the total TEQconcentrations (TCDD/F + PCB) are now of the order of 10–30 pg/g fat[1,130,252,309,408,460]. The decrease is due to strict emissioncontrols and also to the control of concentrations in food (e.g.[130,252]). In the US young adult female population (age group20–39), the concentration was 9.7 pg/g lipid in 2001–2002 (geomet-ric mean) [322].

2.5. Toxicity equivalents

Because we are dealing with a number of chemicals of differentpotencies, simply adding the amounts or concentrations does notgive useful information on total toxicity. Therefore toxicity equiv-alence factors (TEF) are used. This is mostly but not exclusivelybased on potency (see [484]), and the starting point is TCDD giventhe value of 1. Other compounds are designated with TEF valuesfrom 0.00003 to 1. The amount or concentration of a given com-pound is multiplied by its TEF, and the result is the amount or con-centration equivalent to that of TCDD. These partial equivalents ofcongeners are then added up to make sum toxic equivalent (TEQ)of the mixture. This can be used as a proxy of the total toxicamount of dioxin-like compounds while appreciating that it is aconsensus value based on several assumptions rather than a scien-tific fact [484].

2.6. Toxic effects in humans

This subsection will only deal with toxic effects believed to bedue to binding to the AHR. Non-dioxin-like PCBs may have othereffects based on different mechanisms.

Toxic effects confirmed in humans are, fortunately, relativelylimited. For example, the hallmark effect of acute dioxin toxicityin animals, the wasting syndrome (see Section 5), has apparentlynot been detected in humans to date, although one of the two wo-men exposed to high levels of TCDD in Vienna described above suf-fered from a notable body weight loss [143]. However, this wasassociated with nausea and vomiting, which are atypical of thewasting syndrome in animals [342]. In humans, a classical hall-mark of dioxin toxicity is chloracne. It has manifested after occupa-tional exposure [447], as well as after high acute doses [143,436].Experience from the Seveso accident in 1976 seems to indicate thatchildren are more sensitive to the effects of TCDD, with chloracneat concentrations of less than 1000 pg/g fat [280] while someadults with concentrations of up to 10,000 pg/g fat had no chlor-acne. Even in children very few clinical symptoms and almost nopathological laboratory findings were noted regardless of a maxi-mum concentration of 56,000 pg/g TCDD in serum lipid [280].

The most sensitive effects in humans as well as in experimentalanimals are believed to be developmental effects [79]. In animalsthese occur after single doses of 30–100 ng/kg to the dam[151,201,266]. There are very few data on these at the dioxin levelsprevalent in the general human population. Some mineralizationdisturbances of teeth were noted in 7-year-old children (born in1987), related to their dioxin exposure via breastfeeding [13,14].Enamel defects were found in those Seveso victims who were ex-posed under the age of five [11].

Most other findings are from accidents or occupational settingswith very high intakes (cf. [113,447]). The most dramatic effectswere seen after the so called Yusho and Yu-cheng PCB accidents inJapan 1968 and Taiwan 1979, respectively. The developmental ef-fects included intrauterine growth retardation, hyperpigmentation,neurological dysfunctions, natal teeth, and alterations in sexualdevelopment (for reference see [113]). An interesting finding afterthe Seveso accident was a change in sex ratio towards more girlsborn to those fathers who were exposed to TCDD as children [279].

Other possible health effects include liver problems, changes inlipid metabolism or thyroid function, diabetes, other endocrinedisturbances, immunological effects and cardiac mortality hypoth-esized by the authors to be due to stress responses [447]. There islittle evidence of these at the present exposure levels, and in manycases the evidence is ambiguous even at high exposure levels. Oneof the problems is simultaneous exposure to other chemicals, oftenthe main chemicals containing dioxins as minor impurities.

Cancer is perhaps the most debated issue. While it is clear thatdioxins have carcinogenic potential both in animals and in man, itis unclear whether this is only a high-dose phenomenon basedmostly on tumor-promoting effects of dioxins [98]. The Interna-tional Agency for Research on Cancer classified TCDD as a humancarcinogen in 1997 [186]. A large case-control study found no evi-dence of the correlation of dioxins with soft-tissue sarcoma, one ofthe implied ‘‘dioxin cancers” [466]. In a recent study the mortalityin Baltic Sea fishermen with high dioxin levels was found to belower than in controls [475], and this was not only true of low car-diac mortality which may be due to high intake of omega-3 fattyacids in fishermen, but also of cancer mortality. Only mortalityfrom lymphoid/hematopoietic malignancies was non-significantlyincreased. This might be in line with the Seveso cancer studieswhich have implied an excess of lymphatic/hematopoietic malig-nances, among which excesses of multiple myeloma and myeloidleukemia were also significant in the second most contaminatedzone [78,324].

Page 4: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 455

2.7. Toxicity in experimental animals

In animal studies, a myriad of various toxic effects have beenshown mostly by using TCDD (cf. [40,42,338,462,501]. In addition,there are many biochemical effects such as enzyme induction thatmay not necessarily mean toxic insults but could rather representadaptive responses. Animal toxicity described in hundreds of pa-pers will not be exhaustively dealt with here; instead, the readeris referred to the reviews above.

There are several conspicuous features in the toxicity of TCDD.Even high doses of TCDD or other dioxins do not kill animalsimmediately but in one to several weeks, after substantially re-duced feed intake and severe weight loss of even more than 50%– dubbed the wasting syndrome (see Section 5 for details). Lethaldose levels vary widely among species and strains. The LD50 value(dose killing 50% of animals) for guinea pigs is 1–2 lg/kg and formost rat strains 20–50 lg/kg, but the most resistant Han/Wistar(Kuopio) (H/W) rat strain may tolerate more than 10,000 lg/kg.In the case of mice, the LD50 for most strains ranges from 150 to300 lg/kg, but for a resistant strain, DBA/2, it is about 10 timeshigher. The most resistant mammalian species is the hamster withan LD50 of 1000–5000 lg/kg [338].

The mechanisms of intra-species differences vary, but seem to in-volve the AHR. The AHR is a ligand-activated transcription factorwhich regulates the transcriptional activities of numerous genesincluding CYP1A1, CYP1A2 and certain other drug-metabolizing en-zymes. It will be discussed in detail in Section 3. In the DBA/2 mousestrain, the affinity of dioxins to the AHR is decreased due to a muta-tion in the ligand-binding domain of this receptor (cf. Fig. 3), easilyexplaining the sensitivity difference as compared with most otherstrains [348]. In contrast, in H/W rats the binding of TCDD to theAHR as well as the binding of the receptor complex to the DNA occurnormally (see Section 3.1), but alternative splicing in the transacti-vation domain of the receptor protein probably makes the receptorineffective at initiating transcription of some genes but not all[345]. This causes a very peculiar difference as compared with thesensitive-resistant mouse pairs. In the resistant mouse all responsesare hampered about to the same extent, but in H/W rat someresponses are not changed at all (typically induction of CYP1A1and CYP1A2 enzymes) while other responses are drastically weak-ened (e.g. lethality, increase of bilirubin levels) [338,426,472]. Thisencouraged our group to divide dioxin responses to two types, dioxinI responses not sensitive to alterations in the transactivation domain(Fig. 3), and dioxin II responses highly sensitive to them [426].

The mechanism for the different sensitivity to TCDD lethalityand other type II responses between the strains is not clear, butit may be envisioned that ligand-bound AHR, acting as a transcrip-tion factor, up- or down-regulates many genes, and some of these

Fig. 3. Major functional domains of the AHR. The approximate locations of theregions responsible for DNA binding, nuclear translocation and export (NLS andNES, respectively), ligand binding (depicted with red TCDD molecule), HSP90binding, heterodimerization as well as transactivation are shown. The PAS domaincomprises two subdomains, A and B. The transactivation domain has at least threesubdomains, but only one of them, Q-rich, is illustrated.

are sensitive to the changes in the transactivation domain, someare not. Indeed, rat AHR seems to be involved in the regulationof hundreds of genes, and transcriptomic studies reveal numerousdifferences in up- or down-regulation between the strains ([459]).Furthermore, the number of hepatic genes affected by TCDD issmaller in H/W rats than in TCDD-sensitive strains [124].

The wasting syndrome is associated with a variety of changes inmetabolism, but it has been difficult to explain lethality or wastingby these changes (cf. [338]). Acute toxicity is also qualitatively dif-ferent among species, e.g. there is severe liver necrosis in the rabbitbut not in most other species, and there seem to be differences inendotoxin responses, edema, cytokines and inflammation betweenrats and mice [338].

Various pleiotropic responses are typical of dioxin toxicity:there may be both hyperplastic or proliferative responses andhypoplastic or atrophic responses [338]. Hyperplasia of the seba-ceous glands is seen in some species but not in most, even if chlor-acne is typical in human poisoning [42]. Thymic atrophy is aconsistent finding. Some effects on the immune system are ob-served at remarkably low dose levels. It is not fully clear if the mostsensitive changes are related to toxicity or if they are adaptive, butincreased susceptibility towards infections has been noted at quitelow dose levels [42].

The most sensitive effects may be developmental effects, andthese are also the basis of the most recent risk assessment [79].In mice typical malformations are cleft palate and hydronephrosisat doses not toxic to the dam. In rats various adverse effects of sex-ual development [151,266] are more typical, as well as distur-bances of tooth development [12,201]. As mentioned above,many of these developmental effects have in some studies been re-ported to occur at single doses as low as 30–100 ng/kg to the ratdam. However, other studies have found higher thresholds (re-cently reviewed in [31]).

Dioxins are clear multisite carcinogens at doses low in absolutesense but relatively high as compared with general toxicity [98].Much of the cancer risk assessment has been based on a singlerat study [219], demonstrating liver tumors in female rats at lowerdoses (10 ng/kg/day TCDD for 2 years) than most other studies.Toxic hepatitis was also found in animals with tumors, and be-cause there is no evidence of direct genotoxicity of TCDD, non-genotoxic or promoting mechanisms are favored [98]. When differ-ently sensitive Long-Evans (Turku/AB) (L-E) and H/W rat substrainswere compared in a 3-month tumor promotion study, there was adifference of almost two orders of magnitude, and in both strainstumor promotion was associated with signs of liver toxicity[486]. Such findings suggest that carcinogenicity is secondary totoxicity, possibly due to oxygen radical formation or inflammation.

3. Intracellular messenger of dioxins: the AHR

The biological effects of dioxins in the body critically depend on acytosolic protein, the AHR. The evidence supporting this conclusionis multilevel. For example, the toxicity of dioxin congeners correlateswith their binding affinity to the AHR [347,394]. In inbred C57BL/6Jand DBA/2J mouse strains as well as in C57BL/6J mice congenic at theAh locus, a 10-fold difference in binding affinity of the AHR to TCDD isassociated with a sensitivity difference of a similar magnitude inmost of the toxic and biochemical impacts caused by TCDD[41,67]. A similar correlation exists among bird species [198]. Asdescribed above, a substrain of rat, H/W, is up to 1000-fold moreresistant to TCDD lethality than sensitive rat (sub)strains, and thisresistance is mainly attributable to a mutation in its AHR structure[332,338,345,472]. The most definitive evidence comes from studieswith AHR knockout mice, which have proved to be unresponsive toall major effects of TCDD [115,278,491].

Page 5: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

456 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

The AHR belongs to basic Helix–Loop–Helix/Periodic, AHR nucle-ar translocator, Single-minded (bHLH/PAS) proteins which haveimportant roles in protection against hypoxia, in regulation of neuraldevelopment, in generation and maintenance of circadian rhythmsand as transcriptional partners and co-activators [129,211]. Func-tionally, the AHR resembles the nuclear receptors (in which it wasinitially categorized) acting as a ligand-activated transcriptionfactor. In response to activation by dioxins, the AHR signaling path-way modifies the expression levels of numerous genes. The bestcharacterized of these at the molecular level is the induction of thegene for a Phase I drug-metabolizing enzyme, CYP1A1 (for recentreviews, see [28,129,314]).

3.1. The canonical AHR signaling pathway as elucidated for theCYP1A1 gene (Fig. 4)

In the dormant state, the AHR resides in the cytosol in a proteincomplex also containing a dimer of Heat Shock Protein 90 (HSP90),AHR-associated protein-9; also known as XAP2 or AIP (ARA9) andp23 [28,89,265,302]. These chaperone proteins stabilize the AHRmaintaining it in a conformation which is unable to enter the nu-cleus but optimal for ligand binding [203–205,329]. Upon ligandbinding, the AHR undergoes a transformation in which the N-ter-minal nuclear localization signal is exposed and recognized byimportin-b, and the protein complex translocates into the nucleus[187,328,379]. Recent evidence suggests that CaMKs (Ca2+/calmod-ulin-dependent protein kinases) and phosphodiesterase type 2Amay be required for the translocation step [87,282]. The AHR thendissociates from the chaperones and heterodimerizes with anotherbHLH/PAS protein, ARNT (AHR nuclear translocator) [134,377]. TheAHR/ARNT dimer binds to the DNA within the major groove of theDNA helix at specific sites, dioxin response elements (DREs, alsoknown as xenobiotic response elements or AHR elements) locatedupstream of the proximal promoter region of the CYP1A1 gene andharboring the minimal consensus core sequence 50-A/TNGCGTG-30

[90,126,230,444,495]. In these enhancers, ARNT binds to the E-boxhalf-site GTG and AHR to 50 sequence immediately adjacent to it[23,445]. DNA binding is followed by interactions of the AHR/ARNTdimer with transcriptional coregulators such as steroid receptor co-activator 1 (SRC-1) [30,234], steroid receptor co-activator 2 (SRC-2)[450], nuclear receptor co-activator 2 (NCoA2) [30,174], nuclearreceptor co-activator 3 (NCoA3) [30], CREB binding protein (CBP)/p300 [30,217,461], receptor-interacting protein 140 (RIP140)[236], Brahma/Switch 2-related gene 1 (BRG-1) [450,493], thyroidhormone receptor/retinoblastoma interacting protein 230 (TRIP230)[29] and silencing mediator for retinoic acid and thyroid hormonereceptors (SMRT) [304,391,505]. Which coregulators are recruiteddepends on cellular context. Especially important appear to be thosecofactors that possess intrinsic histone acetyltransferase activitysuch as p300, CBP, SRC-1 and SRC-2 since they can relax the chroma-tin structure [30,174,421]. H4 or H3 histone acetylation at theCYP1A1 promoter is then associated with recruitment of RNA poly-merase II to the proximal promoter and launching of CYP1A1 mRNAtranscription [30,273]. AHR activity is terminated by nuclear exportof the receptor [86,187] and by its ubiquitin-mediated degradationby the 26S proteasome [86,263,384].

3.2. Other genes regulated by the AHR

In addition to CYP1A1, dioxin-activated AHR induces otherPhases I and II drug-metabolizing enzymes in liver includingCYP1A2, CYP1B1, CYP2S1, CYP2A5, ALDH3, GSTA1, UGT1A1,UGT1A6, UGT1A7 and NQO1, presumably in most cases by thesame general mechanism [19,27,28,55,229,231,311,448,525,526].However, for the CYP1A2 gene there may be remarkable species-specific variation in DRE location [316] as well as in DRE structure

and in the mode the AHR-ARNT heterodimer binds to it [433]. Fur-thermore, the contribution of Nrf2 (erythroid 2-related factor 2)appears to be required for induction of UGT1A6, NQO1 and GSTA1[264,523], and ALDH3 may be also be regulated by mechanismsdifferent from the canonical AHR signaling pathway [101,227]. Tis-sue-dependency of the induction phenomenon varies among theenzymes: While CYP1A1 responds to dioxins in a wide variety oftissues, induction of CYP1A2 is mainly confined to liver and nasalepithelium [146,259,503]. Due to its ability to bind dioxins, in-duced CYP1A2 can modulate their kinetics by accumulating themin the liver [207,349,350,490]. There is a nice correlation amongdioxins and related compounds between their potency to elicit he-patic CYP1A1 induction and to cause major toxicities such as im-mune suppression, body weight loss and thymic atrophy [392].However, differences among species or strains in their sensitivitiesto the acute toxicity of dioxins are generally not reflected in theirsensitivities to enzyme induction, and thus enhanced activity ofcytochrome P450-associated mono-oxygenases is not believed tobe causally related to, or play any major direct role in, dioxin tox-icity ([138,311,333], although there are some contradictory datafrom mice [431,482].

Apart from drug-metabolizing enzymes, TCDD exposure modi-fies the expression of a large number of other genes, presumablyby a similar mechanism. For example, in adult mouse or rat liver,hundreds or even thousands of genes are affected [49,123,124,401,459]. A somewhat surprising finding has been that there ap-pears to be conspicuously little overlap in the patterns of targetgenes across species or among tissues within a given species [48–50,63,212,361]. Frequently, gene expression is repressed by TCDDinstead of being induced. Next to nothing is known about the mech-anism(s) of this gene silencing [381] but at least in adult rats, it doesnot appear to involve micro-RNAs [281]. Epigenetic regulationthrough DNA methylation or histone modification may, however,be at play [375].

One of the genes induced by TCDD plays a role in AHR regulationby forming a feedback loop: AHR repressor (AHRR) [277]. It inhibitsthe transcriptional activity of dioxin-activated AHR by a mechanismwhich is incompletely understood at present but which may involveSUMOylation of certain lysine residues in the C-terminal repressiondomain of the AHRR [318] (reviewed in [159]). The effect of the AHRRis not fully specific to the AHR but extends to a structurally relatedprotein, hypoxia-inducible factor signaling [199].

3.3. Non-canonical signaling pathways

The AHR has been shown to interact with other signal transduc-tion pathways including nuclear factor kappa B (NF-jB) [206,214,321,455,488], protein tyrosine kinases and EGFR (epidermal growthfactor receptor) [21,73,110,443], p53 and pRb (retinoblastoma)[108,140,319], HIF-1a (hypoxia-inducible factor 1a) [65,177], TGF-b (transforming growth factor b) [131,190,382], Nrf2 (nuclear factorerythroid 2-related factor 2) [423,523], cAMP/PKA (cyclo-AMP/pro-tein kinase A) [88,95,312,489] and MAPKs (mitogen-activated pro-tein kinases) [239,320,449,499] (for reviews, see [62,156,359,362,458]). Importantly, this crosstalk appears to often occur by mecha-nisms distinct from the canonical pathway such that DRE bindingor heterodimerization with ARNT is not involved. Both the classicaland alternative pathways can also jointly contribute to the final out-come as in the case of one of the most exhaustively examined pro-teins interacting with the AHR, the estrogen receptor (ER)(reviewed in [272,446]). Dioxin-activated AHR interferes with ERsignaling at multiple steps: It causes transcriptional repression ofER [456,457], may reduce ER ligand levels [438], interferes with ERsignaling through inhibitory DREs [393], recruits ER to establishedAHR-regulated genes (where ER may enhance AHR function butAHR diverts it from estrogen target genes) [273,507], competes with

Page 6: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 457

ER for ARNT as a partner or co-activator [389], and can target ER forproteasomal degradation by acting as an E3 ubiquitin ligase [313].

One of the earliest changes caused by TCDD in vitro is an in-crease in intracellular Ca2+ concentration [43,164,360]. This ef-fect may have important sequelae such as activation of CaMKswhich appears to converge on the canonical pathway at theAHR nuclear translocation phase [282]. Furthermore, it has beenproposed to lead to elevated levels of Cox2 mRNA and thus to aninflammatory response by two distinct alternative nongenomicpathways whose selection depends on cellular context [94–96,271,413,414].

Although it has been well-established that the AHR itself isindispensable for virtually all aspects of dioxin toxicity examined[15,185,255,256,278,306,454,491], the contribution of the alterna-tive AHR signaling pathways in the diverse biochemical and toxiceffects of dioxins is as yet poorly explored. Two characteristic ter-atogenic impacts of TCDD in mice are hydronephrosis and cleft pal-ate. It was recently shown that hydronephrosis induced in mousepups by lactational exposure to TCDD appears to be criticallydependent on Cox2 [305] suggesting a key role for the alternativepathways. However, a systematic approach by means of transgenicmouse models carried out in Christopher Bradfield’s laboratory hasconvincingly demonstrated the supremacy of the canonical path-way in the mediation of the major outcomes of dioxin exposurein vivo. Mice in which either the nuclear translocation or DRE bind-ing is defective are refractory to both adaptive and toxic effects ofTCDD (CYP1A1 induction, hepatotoxicity, thymic atrophy and ter-atogenesis) [57,58]. An especially noteworthy observation in thesestudies was that hydronephrosis elicited by in utero exposure toTCDD was absent in mice whose AHR is unable to bind to the DREs.Either this lesion evolves by different mechanisms depending ondevelopmental stage or in this case induction of Cox2 occurs viathe classical pathway. Furthermore, mice which express reducedlevels of ARNT exhibit CYP1A1 induction but not thymic atrophyor hepatotoxicity [492]. Interestingly, all these mouse models withimpaired canonical AHR signaling also manifest the developmentalphenotype of AHR-deficient mice, foremost persistent ductusvenosus and a small liver (see below). This implies that both theeffects triggered by binding of a high-affinity ligand (TCDD) tothe AHR and the major endogenous functions of the AHR are med-iated by one and the same, canonical, signaling pathway.

3.4. Phylogeny of the AHR

Evolutionarily the AHR appears to be an over 500-million-year-old protein occurring in all vertebrates [59,160]. Ancient homologsof the AHR have even been discovered in invertebrates such asnematodes (C. elegans) [356] and insects (spineless in D. melano-gaster) [100]. However, a distinctive feature of these primitiveAHR forms is their inability to bind dioxin or other ligands of themammalian AHR [59,161,356]. In spite of this incapability, theyhave been shown to play important developmental roles in neuro-nal differentiation and regulation of feeding-related aggregationbehavior (C. elegans) [183,365,366] or in regulation of normal mor-phogenesis of the leg or antenna and bristles (D. melanogaster)[100,109]. A recent study provided direct experimental evidencethat the drosophila AHR homolog is constitutively active [232].

While mammals have only a single AHR gene, fish and birds areendowed with at least two distinct AHR genes (for reviews, see[157,158,160]). Nevertheless, splicing variants of both the AHRand ARNT have been found to exist in rats [226,345].

3.5. Molecular structure of the AHR in relation to its function

The AHR has a modular structure (Fig. 3). The ligand-bindingdomain is located in the middle region and exhibits variation in

primary structure that is functionally reflected in inter- and intra-species differences in sensitivity to dioxin toxicity. For example, sen-sitivity differences between C57BL/6 and DBA/2 mouse strains aswell as those among bird species derive from polymorphisms atone or two amino acids in the ligand-binding domain resulting inaffinity differences in dioxin binding [53,167,198,348,398]. Lowaffinity of amphibian AHR to TCDD may also account for the insensi-tivity of frogs to dioxin toxicity [26,245]. At the N-terminal end, thebasic region is responsible for DNA binding [22,127] while the HLHmotif constitutes the primary interface for heterodimerization towhich the PAS domain confers specificity [257,351]. The N-terminalend further contains structures required for HSP90 binding as well asfor nuclear entry and exit [80,127,187,323]. At the C-terminus, theAHR carries a large transactivation domain consisting of interactingsubunits one of which is a glutamine-rich subdomain [188,192,235,388,432,496,502]. The transactivation domain is intimately in-volved in the interplay with co-regulatory proteins and general tran-scription factors [234,388,493,496], and influences the intracellularlocalization of the receptor [371]. The C-terminal structure of theAHR can also be a key determinant of dioxin sensitivity being sub-stantially altered in the most TCDD-resistant laboratory animals,hamsters and H/W rats [225,314,332,472,486]. However, in contrastto alterations in the ligand-binding domain, remodeling of the trans-activation domain seems to result in a response-selective modula-tion of dioxin sensitivity [314,332].

3.6. Activation of the AHR

The best-understood way of AHR activation is by ligand binding.In addition to dioxins and other halogenated aromatic hydrocar-bons, polycyclic aromatic hydrocarbons (PAH compounds) such asbenzo[a]pyrene and 3-methylcholanthrene constitute another largegroup of toxicologically important environmental contaminantswhich can bind to the AHR [39,47,373]. Although the intrinsic affin-ity of PAHs to the receptor does not necessarily lag far behind that ofTCDD, their potency in bringing about the adaptive enzyme induc-tion response is several orders of magnitude lower, and they donot elicit an identical pattern of toxicities (e.g. wasting syndrome)to that of dioxins [346,347,380]. One of the main reasons for this dis-crepancy is believed to be more persistent receptor binding by TCDDdue to its slower metabolic inactivation compared with PAHs[47,380]. The same explanation also applies to a high-affinity ligandof dietary origin, indolo(3,2-b)carbazole, which appears to bemetabolized too rapidly to cause AHR-mediated toxicity [334].

The ligand-binding domain of the AHR is not highly selectiveregarding the molecular properties of the ligands (for review, see[91]). Both exogenous and endogenous compounds as diverse asflavonoids, 7-ketocholesterol, tryptophan and tryptophan metabo-lites, bilirubin, indirubin, indigo, fused mesoionic heterocycles, andarachidonic acid metabolites such as lipoxin A4 can bind to thereceptor (albeit at varying avidities) and function as either agonists,partial agonists or antagonists [5,7,44,121,136,168,307,325,330,402,403,427,528]. The potency of a compound as a modulatorof AHR action may vary among species [121,171,530]. There is alsoevidence that the pattern of effects mediated by the AHR may de-pend on the ligand, on species from which the receptor originates,or on ligand � receptor interaction, thus providing scientific groundfor the development of selective modulators of AHR action in the fu-ture [122,174,295–297,315,529].

It is possible that ligand is not a prerequisite for AHR activation.For example, certain carotenoids and benzimidazole derivativessuch as omeprazole can induce CYP1A1 in some species in anAHR-dependent manner, but apparently fail to bind to the AHR[85,105,150,248,400]. Similarly, in cell cultures CYP1A1 has beenfound to be inducible by hydrodynamic shear stress [290,291]and by certain heavy metals [107]. Moreover, the AHR has been

Page 7: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

1.

2.

Ca2+Cox2 mRNA

Phase I & II DMEs

AHRR

Toxicity genes

mRNA for:

3.

APP mRNA6.

5.

Ubiquitin-26Sproteasome

4.

E3 ubiquitin ligase complex

ER β-catenin

7.

Fig. 4. A schematic and simplified diagram of some salient features of AHR signaling pathways. The canonical pathway is depicted with solid black arrows, alternativepathways with dashed black arrows, and an intersection of these two distinct but linked signal transduction routes with a solid red arrow. The blue bars represent the AHR,light red bars ARNT, green bars ARA9, yellowish brown bars HSP90 and the bright red ovals p23. Dioxin binding to the AHR (1.) leads to its translocation into the nucleus (2.),heterodimerization with ARNT and binding to the DNA at DREs with ultimately modulating expression levels of target genes (3.). One of the gene products elevated by thismechanism is AHRR which forms a feedback loop by inhibiting AHR action. The AHR is finally degraded by the ubiquitin–proteasome system (4.). AHR activation can alsorapidly increase intracellular Ca2+ concentration (5.) which in turn may (through intermediate steps not shown) ultimately result in augmented Cox2 gene expression.Elevation of Ca2+ activates CaMKs which appear to have a critical role in the translocation of the AHR. Two further examples of effects mediated by the AHR via non-canonicalpathways are suppression acute-phase proteins (6.) which does not involve DNA binding, and degradation of e.g. ER and b-catenin [202] by acting as an atypical E3 ubiquitinligase (7.).

458 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

reported to be activated by a high glucose concentration [82].However, it is also possible that the current methods employedfor analysis of AHR ligand binding are not sensitive enough to re-veal a very weak association between the AHR and either the com-pound studied or some endogenous ligand (e.g. arachidonic acidmetabolite or low-density lipoprotein in the case of the shearstress [275,292]). Indeed, involvement of the AHR in the regulationof a number of essential physiological phenomena in the body (seebelow) strongly suggests the existence of a high-affinity endoge-nous ligand (or ligands) for the receptor. Although tryptophanmetabolites in particular have attracted attention in this respect[169,294,303,509], the issue remains unresolved at present.

3.7. Physiological functions of the AHR

AHR-deficient mice have proved to be a highly fruitful model inassessing the role of the AHR in the development of organs and inphysiological functions outside of, and beyond, xenobiotic bio-transformation. These knockouts have independently been gener-ated by 3 laboratories [114,278,407], and their phenotypes haveboth common and distinctive features [240]. One of the most con-sistent findings has been a 25–50% reduction in liver size whichmay result from persistent ductus venosus and thereby from a por-tocaval shunt [165,241,242], or from augmented apoptosis due toelevated levels of TGF-b [147,527]. Relevant to this, angiogenesisin the AHR-deficient mice has been reported to be impaired by amechanism that seems to involve TGF-b [385]. The AHR-null micealso display a slightly slower growth rate compared with wild-typemice for the first few weeks of life [114,278,407]. In addition to theliver, the development of a number of other organs and tissues isinfluenced by the presence or absence of the AHR including heart,spleen, thymus, kidneys, lungs, salivary glands, skin, testes, ante-rior and dorsolateral prostate, seminal vesicles, epididymides, ova-

ries, and, evidently, ventral telencephalon [116,172,255,256]. Inadult mice, the AHR status modulates the expression of severalhundred genes both in liver and in kidney [48,459]. AHR knockoutmice are either hypertensive or hypotensive depending on the alti-tude they are housed [261], and they may be prematurely affectedby age-related degenerative changes [116] leading to a shortenedlife-span [176]. Furthermore, they exhibit reduced fertility [4] atleast in part due to impaired conversion of testosterone to estradiolby ovarian P450 aromatase (Cyp19) [20]. They are also hypersensi-tive to bacterial endotoxin (lipopolysaccharide) [20,214,419,452]as well as having some other less pronounced immune alterations(reviewed in [111,441]).

4. AHR and the CNS

In vitro studies have implied that the degree of AHR expressionmay influence neuronal function. Reduction of AHR levels bymeans of small interfering RNA has been shown to reduce N-methyl-D-aspartate (NMDA) excitotoxicity while augmentingNMDA-induced brain-derived neurotrophic factor expression incortical cells [253]. Over-expression of AHR in the undifferentiatedmurine neuroblastoma Neuro2a cell line, in turn, was reported toinduce catecholaminergic differentiation in them [10].

4.1. Expression of principal molecules of the AHR signaling cascade inthe CNS

The AHR itself as well as other key proteins of the canonicalAHR signaling pathway are expressed from early stages on in themammalian CNS. For example, in mouse embryos both AHR mRNAand protein were present in high concentrations in the neuroepi-thelium of the developing brain at gestational days (GD) 10–11but then started to diminish with only regional, low levels being

Page 8: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 459

expressed after GD12 [2]. Regarding ARNT in mice, in situ hybrid-ization revealed its mRNA to be highly expressed in the neuroepi-thelium of neural tube at GD9 and especially in neocortical andtectal neuroepithelium thereafter, but overall its levels tended todecline with differentiation of neuronal structures; in hypothala-mus, a strong signal was recorded at GD11 and GD15 [9,189]. ARNTprotein on the other hand was highly expressed in neuroepithe-lium between GD10 and GD12.5 but decreased thereafter in neuralstructures by GD15 [3,434]. Both primary cortical astrocytes frompost-natal day 1 and primary cortical endothelial cells from wean-ling C57BL/6 mice yielded a positive signal in immunocytochemi-cal analysis of AHR protein [120]. When cerebellar AHR andARNT protein expressions were monitored by Western blotting inC57BL/6J mice from birth to adulthood, AHR proved to be presentat all time-points and ARNT from post-natal day 3 on [508].

The perinatal expression profiles of hypothalamic AHR andARNT mRNAs were studied by semi-quantitative RT-PCR in Spra-gue–Dawley rats. AHR mRNA was found to rise steadily from thestart of the study (GD16) to birth and then retain that level tothe last time-point of analysis (post-natal day 15). ARNT mRNA,in contrast, stayed at a constant concentration throughout theobservation period [357]. In another study on 3-day-old rat pups,AHR proved to be expressed in virtually all GABAergic neurons inthe preoptic area [166].

In adult C57BL/6J and DBA/2J mice, AHR mRNA expression levelin brain was similar to that in liver and about 10-fold lower than inthe tissue displaying the highest expression, lung [250]. In the caseof AHRR mRNA, the highest concentrations among the nine tissuesexamined in C57BL/6 mice were found in the brain and heart; inAHR-deficient mice these concentrations were two orders of mag-nitude lower implying a key regulatory role for the AHR [34]. In animmunohistochemical study of the adult Balb/cX129SV mousebrain, AHRR-immunoreactivity was recorded in nuclei of neuronsall over the hypothalamus with the greatest intensity in the arcu-ate nucleus (ARC); outside of the hypothalamus, strong stainingwas seen in the hippocampus and cortex [117].

AHR and ARNT mRNAs are also widely expressed in adult ratbrain [182,195,326]. The presence of their proteins has furtherbeen confirmed by immunoblotting [237]. Although an early studyfailed to detect AHR mRNA in the rat hypothalamus, subsequentstudies have convincingly demonstrated that AHR, ARNT and AHRRare all expressed at this site [182,224] with especially high levels ofAHR and ARNT occurring in caudal ARC [326]. In the suprachias-matic nucleus (SCN), AHR mRNA proved to exhibit circadian fluctu-ation with only slightly smaller amplitude than that for mRNA ofthe clock gene Per1 [293]. The hypothalamus exhibited pro-nounced DRE binding in the rat being second only to cerebellumof the seven brain regions examined [237]. AHR or ARNT mRNAlevels in rat brain do not seem to be influenced by TCDD treatmentin contrast to those of the AHRR which are rapidly (within hoursafter exposure) elevated [182,224].

In human brain, the expression level of AHR mRNA was lowcompared with heart, lung, liver and especially placenta [93,103].An RT-qPCR analysis of adult human samples revealed AHR mRNAexpression to be 2.7 times higher in brain microvessels than in thecortex suggesting preferential expression in the blood–brain bar-rier [84]. In support of this view, CYP1B1 was also highly expressedin the blood–brain barrier. CYP1B1 mRNA represented over 80% ofall the CYP mRNAs detected and its levels were 14 times as high inthe microvessels as in the cortex. The presence of CYP1B1 proteinwas further confirmed by Western blotting [84].

4.2. Induction of drug-metabolizing enzymes in the CNS by TCDD

As functional evidence of a fully integrated AHR signal trans-duction system in the CNS, the classic adaptive response to TCDD

exposure, induction of Phases I and II drug-metabolizing enzymes,has been well documented to occur in various regions of the rodentbrain at mRNA, protein and enzyme activity levels [66,179,181,182,224,476]. In Sprague–Dawley rats, a non-lethal dose of10 lg/kg TCDD brought about a fairly uniform induction of CYP1A1mRNA throughout the brain 1–28 days after exposure. The basallevels amounted to 15–90% of those in liver, but the inductionresponse was much weaker in the CNS [182]. On the other hand,10 days after 5 or 50 lg/kg TCDD to dioxin-sensitive L-E and diox-in–resistant H/W rats, certain regions of the brain (olfactory bulband mid-brain + thalamus) exhibited induction in CYP1A1 activity(measured as EROD [ethoxyresorufin O-deethylase] activity)whose magnitude was almost comparable to that measured inthe liver [476]. Interestingly, in the hypothalamus of these tworat strains not only CYP1A1 but also CYP1A2 mRNA (whose induc-tion exhibits high tissue specificity) was found to be rapidly andpersistently induced by 50 lg/kg TCDD with the magnitude ofCYP1A2 (but not of CYP1A1) induction correlating with TCDD sen-sitivity of the strains [224]. Treatment of Sprague–Dawley ratswith another AHR agonist, b-naphtoflavone, also led to inductionof both CYP1A1 and CYP1A2 mRNAs in various regions of the brain[406]. After b-naphtoflavone treatment, CYP1A1 immunoreactivityand catalytic activity appeared to largely localize in choroid andarachnoid membranes [288].

As to individual cell types in the CNS, CYP1A1 and CYP1B1induction response to TCDD has been reported in mouse primarycerebellar granule neuroblasts [508] and in mouse vascular endo-thelial cells; in mouse astrocytes, CYP1B1 but not CYP1A1 was in-duced [120].

Thus, there is little doubt that a fully functional AHR signalingcascade is present in the CNS, evidently also in humans, and ableto mount an induction response to dioxin exposure.

5. Alterations in consummatory behaviors and in centralregulatory mechanisms of energy balance in dioxin-exposedlaboratory animals

In experimental animals, the acute toxicity of dioxins is typifiedby a peculiarly delayed occurrence of mortality preceded by a dra-matic body weight loss. Although this phenomenon has beenknown for over three decades, its biochemical basis has remainedelusive. However, the potency of TCDD and some other dioxins toirreversibly and fatally suppress the numerous and redundant sys-tems existing in the body to secure one of the most vital functions,adequate food intake, as well as to perpetually derail the controlmechanisms for body weight maintenance implies that resolutionof the pathogenesis of the wasting syndrome might cast new lighton the key factors of energy balance more generally. We have pre-viously reviewed the early data on the influence of TCDD on theCNS, feeding behavior and energy balance [338,478]. In the presentreview, we will mainly focus on more recent findings on alterationscaused by dioxins in consummatory behaviors and on the currentunderstanding of their biochemical basis, but will also outline thesalient facets established previously to help the reader composethe full picture.

5.1. The wasting syndrome

Dioxin exposure causes a dramatic reduction in feeding and aconsequent decline in body weight in exposed animals. This phe-nomenon is known as the wasting syndrome and it is a ratheruncommon manifestation of chemical toxicity in general. A charac-teristic feature of acute dioxin toxicity is that lethality is delayedand preceded by a progressive body weight loss which occurs evenafter a single administration. Death does not ensue until 1–8 weeks

Page 9: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

460 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

after exposure with the time course depending on dose as well ason sensitivity of the animal [42,347]. At lethal doses of dioxin, thehypophagic animals can lose over 50% of their body weight beforedeath [338,416,417]. In addition to mice and rats, also other animalspecies such as hamsters and guinea pigs have been reported tosuffer from hypophagia after dioxin exposure although at widelydifferent doses [137,170,209,317]. Despite the fact that TCDD isan extremely potent compound in reshifting body weight, thewasting syndrome has attracted surprisingly little research inter-est. Because of the extensive research efforts put on elucidatingthe mechanisms of obesity and the physiological regulation of en-ergy homeostasis, the resolution of the pathogenesis of the wastingsyndrome would be important not only toxicologically but alsophysiologically.

5.1.1. Relationship with TCDD lethalityThe reduction of body weight has been shown to primarily re-

sult from hypophagia in rats, mice and guinea pigs [209]. Never-theless, TCDD-treated rats do not appear to suffer from nausea[342]. It is also important to note that neither gross malabsorptionnor increased energy expenditure seem to contribute substantiallyto the wasting [354,416,418]. Therefore, our focus in this sectionwill be centered in factors regulating energy intake and bodyweight. Depletion of energy stores in consequence of hypophagiais a major cause of death in dioxin-exposed rats as evidenced bya similar life-span in pair-fed controls [75,498]. However, the factthat force-fed rats [135] and guinea pigs [180] do not lose weightbut still succumb at approximately the same time after TCDDadministration indicates that other mechanisms must also be atplay.

5.1.2. Dietary modulation of the wasting syndromeWhether nutritional obesity, forced feeding or different diets

could diminish or postpone the wasting syndrome was studied inour laboratory using TCDD-sensitive L-E and TCDD-resistant H/Wrats of both sexes [468]. L-E rats having free access to a high-en-ergy diet survived longer after TCDD exposure than rats on a regu-lar diet. The type of diet or feeding protocol was found to affectsurvival: rats on a liquid, high-fat diet (force-fed by gavage) suc-cumbed earlier than rats fed the regular feed (either a fixedamount or ad libitum) or a balanced liquid feed. In fact, it has beenshown by several groups that fat seems to be an unfavorablesource of energy after TCDD exposure [298,299,473]. A balanced li-quid diet (force-fed) prevented weight loss in L-E males following alethal TCDD dose, 50 lg/kg, but still none of the exposed rats sur-vived and the time to death was similar to that in their free-feedingpartners [468]. Obesity, high-energy diet or forced feeding did notprevent weight loss in resistant H/W rats, treated with a high butsublethal dose of TCDD. Although the body weights of H/W ratswere higher on a high-energy diet than on the regular diet,TCDD-treated rats still maintained their weights at a lower levelthan their controls [468]. Taken together, it seems that whereasTCDD lethality cannot be prevented by parenteral nutrition orforce-feeding in rats [135,468] or in guinea pigs [180], the typeof diet affects the severity of TCDD-induced body weight loss inexperimental animals [298,299,468].

A puzzling and totally unexpected outcome appeared in a re-cent study in female C3H/HeN mice that harbor a high-affinityAHR for TCDD. In mice kept on a high-fat diet, repeated exposureto 100 lg/kg TCDD every 2 weeks for a period of 8 weeks resultedin a significantly accelerated body weight gain compared with vehi-cle-treated mice on the same diet. By the end of the experimentalperiod, the mice treated with this dose of TCDD weighed 46% morethan controls on the high-fat diet. In lower exposure groups (1 or10 lg/kg), body weight gain did not deviate from that in controlsand the same was true of all TCDD-dosed groups fed a standard

diet [531]. Although extremely rare, this response opposite towasting is not entirely without precedent: 3,4,30,40-tetrachlorobi-phenyl, a co-planar PCB congener (i.e. an AHR agonist), was foundto cause obesity in Skh:HR-1 mice [363]. However, no cases ofobesity provoked by TCDD or related compounds have so far beenreported in species other than the mouse.

5.1.3. The wasting syndrome and reduced body weight set-pointOnly a single administration of TCDD is needed to elicit the

wasting syndrome because of the sluggish biotransformation andelimination of TCDD in laboratory animals [137,162,344,497]. Dur-ing the first few days after exposure, food intake typically takes aprogressive downhill course leading to weight loss, but the ex-posed animals do not usually show a total refusal of feeding. Atsublethal doses, the animals resume feeding and gaining weightin 1–2 weeks [338,416,417]. Supralethal doses do not much short-en the time to death. In both rodents and nonhuman primates alethal dose leads to death by 8 weeks [42]. Sublethal doses maycause a permanently stunted growth, decreased feeding [340,416,417] and alterations in feeding behavior [340,473].

Even though it is known that TCDD affects several physiologicalmechanisms involved in the maintenance of body weight balanceand energy homeostasis (previously reviewed e.g. in [338,46,267]),the biochemical factors underlying the wasting syndrome and lead-ing to death are still elusive. Lethally TCDD-exposed animals losetheir body fat stores and even muscle mass. However, rats exposedto high but sublethal doses are able to gain weight (although theirbody weight lags behind that of control rats), increase their feed in-take after fasting, and defend their lowered body weight levelagainst feeding challenges [337,338,416–418]. These factors areconsistent with lowered body weight set-point in dioxin-treated rats[336,337,416–418]. This hypothesis is further supported by findingsfrom an experimental model of simultaneous hyperphagia andunderweight: diabetic rats. Rats rendered diabetic with streptozoto-cin continued to exhibit hyperphagia relative to non-diabetic controlrats after TCDD administration, thus revealing that hypophagia is asecondary rather than the primary effect of TCDD [343].

5.1.4. The role of set-point in body weight regulationEnergy homeostasis, body weight and adiposity are tightly reg-

ulated in animals, albeit the reference point of this homeostasis isbiased to shift upwards to favor excess adiposity resulting in a newbalance that is then defended [289,397,422,511]. This unbalancedregulation – attested in the human population by the current rapidincrease in the prevalence of obesity – may have been evolution-arily beneficial until quite recently [358] thus resulting in a ‘‘thriftygenotype”, or it may represent a genetic drift of the upper inter-vention limit of the body weight after the release of predatorypressure some 2 million years ago [437]. An opposite shift in en-ergy balance towards reduced body fat seems less favored [412]and is seldom observed in a physiological context (except for sea-sonal animals) without external manipulation.

The CNS orchestrates long-term energy balance regulation andimmediate-to-short-term energy repositories (mainly plasma glu-cose and FFA and liver glycogen) on the whole organism level byintegrating the information of the amount of adipose tissue, bloodnutrient levels and environment [76,154,410]. The integratedinformation is further converted into changes in energy intake,expenditure and release from storage [243,411,422]. As early as1969 Hervey proposed that there is a set-point towards whichbody mass (or its correlate) is regulated [173]. Subsequently, thishypothesis has been contested by some researchers contendingthat the body weight would just represent a ‘‘settling point” forall the integrated hormonal, metabolic and behavioral actions thatinfluence energy homeostasis [510]. The critical issue discriminat-ing between the two models is the nature of the response to a

Page 10: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 461

perturbation: an active defense to oppose the change is indicativeof a set-point [60,208]. Numerous studies in laboratory animalshave demonstrated that body weight is indeed defended [208].Nevertheless, gradual caloric excess can result in a fixed upregula-tion of the defended body weight level, relevant to the obesity con-dition of millions of humans today [249]. In seasonal animalspecies such as the Siberian hamster exhibiting regular rhythmsof body weight that depend on the length of the photoperiods,the point of energy balance is continuously re-adjusted duringthe transition in body weight reflecting an apparent ‘‘sliding set-point” [284] or a dynamic, defended steady state [36]. The compel-ling evidence of the avidity of dioxin-treated rats to defend theirlowered body weight level cited above shows that the wasting syn-drome is accompanied by, and presumably due to, down-shiftedset-point, and thus it can provide a useful model of chemicallyreset ponderostat (a sensor of body weight or its correlate withthe associated physiological machinery to counteract disturbances;analogous to a thermostat) for studies on the regulatory mecha-nisms of energy balance maintenance.

5.1.5. The primary target of energy balance regulation and the wastingsyndrome

For the body weight set-point, the regulated primary variable isstill not established despite the fact that the importance of the sizeof fat deposits in body weight regulation has been known for over50 years [210]. Initially, blood steroids were suggested as the crit-ical correlate of body weight [173]. More recently, this was slightlymodified by hypothesizing that hypothalamic corticotropin-releas-ing factor (CRF) concentration would serve this function [60]. Onthe other hand, studies in seasonal mammals have yielded strongevidence for thyroid hormone availability as a cause for the rhyth-mic alterations in energy balance. Tri-iodothyronin (T3)-releasingimplants placed in the hypothalami of Siberian hamsters were ableto totally block their short-day-induced weight loss [106]. The ulti-mate regulator in this case might conceivably be the thyrotropin-releasing hormone neurons [246]. In the hypothalamus, the regu-lation of thyroid hormone signaling is very complex involving bothglial cell-expressed thyroxin-activating deiodinase D2 and neuro-nal triiodothyronine (T3)-inactivating deiodinase D3 [141]. Finally,synaptic or neuronal plasticity with a perpetual struggle betweenorexigenic and anorexigenic tones [6,92] might form a structuralbasis for a body weight set-point.

According to the CRF hypothesis, a decrease in hypothalamicCRF concentration would stimulate food intake and vice versa. Inrats treated with a dose of TCDD capable of down-regulating bodyweight, plasma ACTH and corticosterone levels are elevated[38,283], which is suggestive of increased hypothalamic CRF. In-deed, a recent study detected increased mRNA levels of CRF inhypothalamic paraventricular (PVN) and ARC nuclei 7 and 14 daysafter TCDD administration, although there was a transient reduc-tion in CRF expression in the central amygdala and bed nucleusof stria terminalis (BNST) at 2 days [283]. PVN has been shown tobe a critical site for the effects of CRF on food intake [228]. Thus,hypothalamic (especially PVN) CRF concentrations are a potentialmediator of dioxins’ impacts on energy balance and worth furtherresearch (see also Section 5.4.3).

As to thyroid hormone balance, plasma thyroxin (T4) levels areseverely diminished after TCDD exposure in TCDD-sensitive rats(substantially more than in pair-fed controls) but elevated in ham-sters which represent a TCDD-resistant species [149,170,335,355].However, this promising correlation is shattered by TCDD-resistantrats which also display significantly decreased T4 concentrations inplasma [335]. The type II deiodinase (deiodinase D2) which con-verts T4 to the biologically more active T3 and is also expressedin the brain (concentrated in a population of specialized glial cells,tanycytes, located in the base and infralateral walls of the third

ventricle) showed a tendency to increased activity in response toTCDD [370,409], plausibly as a compensatory mechanism.

5.1.6. Effect of TCDD on body weight in developing ratsAge and developmental stage of the animal at dioxin exposure

are important factors for dioxin toxicity [424,425]. In adult rats,the sensitivity to the acute lethality of TCDD is reflected in theseverity of the wasting syndrome [338]. Body weight gain is alsoaffected in prenatally dioxin-exposed rats and intimately associ-ated with the general toxicity of dioxin, but at this stage the adultpatterns of sensitivity do not seem to hold. When rat dams of dif-ferently sensitive lines A, B and C (LD50 > 10,000, 830 and 40 lg/kg,respectively) were treated with 1 lg/kg TCDD on gestation day 15,male progeny of all three lines showed lower body weights com-pared with their controls [424]. On the other hand, when male ratsof lines A and B were exposed to a single dose of dioxin betweenpost-natal days 2–56, their growth was immediately retarded withthe effect being more severe with early exposures. The rats dosedyounger did not reach the body weights of their counterparts trea-ted at an older age during the 3-month observation period [425].This finding is in agreement with the hypothesis of altered bodyweight set-point in TCDD-treated rats [336,337,417]. Nevertheless,it still remains to be determined whether the mechanism of inhib-ited body weight gain is the same for adult and developing rats.

5.2. Biochemical signals involved in long-term regulation of energybalance (Fig. 5A)

5.2.1. Signals inducing satietyAnimals and humans adjust the amount of food eaten according

to the caloric content of the food, their energy needs and long-termenergy balance with an inherently robust system to favor a positiveenergy balance. When ample, uniform food is constantly availableand attainable, regular eating patterns integrated with otherbehaviors are established – the major ‘‘zeitgeber” being the light–dark cycle. This also enables the use of distal cues (smell, taste,texture, etc.) in regulating the size of meals and preparatory physio-logical alterations that allow the consumption of larger meals[442,513,515]. Even in nature, with more variation in both availabil-ity and choices of food, the maintenance of energy homeostasis isprimarily based on adjusting the amount of food eaten per meal(meal size), even though this regulation is mixed with or even dom-inated by environmental influences and internal extrahypothalamicnon-homeostatic functions and processes of food reward, satisfac-tion, learning and cognition [289,422].

5.2.1.1. Leptin and insulin. According to the generally accepted ‘‘adi-posity-feedback model” the long-term adjustment of body weightand regulation of energy balance in mammals is based on keepingthe body fat stores constant with the aid of adiposity signals (forrecent reviews, see e.g. [289,397,514]). Two proteins, leptin origi-nating from adipocytes and insulin produced by the pancreas, aresecreted in a direct proportion to the amount of adipose tissueand thus convey information on body adiposity to the CNS by a sat-urable transport system across the blood–brain barrier [24,25].Insulin also has a second, equally important and partly overlap-ping, role in constant regulation of plasma glucose [142,243].

Depletion of leptin or insulin, or their receptors, results in in-creased food intake and obesity mediated by the CNS (it shouldbe noted that type-I diabetes represents a non-physiological stateof low insulin where peripheral glucose utilization is severely im-paired leading to weight loss) [142,397]. Intracerebroventricularadministration of exogenous leptin or insulin induces markedreductions in food intake and body weight [153,163,223,396]. Ifleptin administration continues for about 2 weeks, food intake nor-malizes while body weight does not [163,395,396]. Resistance to

Page 11: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

A

B

Fig. 5. A highly simplified, schematic representation of some major centers in the CNS involved in food intake regulation and in the reception of peripheral signals mediatinginformation related to adiposity, nutrients and satiation. (A) Nutrients as well as signals of adiposity and satiation are sensed in multiple nuclei and subregions both withinand outside the hypothalamus with ARC occupying a central position within it. Hindbrain NTS has an essential role in satiety perception and is an autonomous integrator ofthe vagal and hormonal signals. (B) Neuronal pathways involved in satiation-based food intake regulation and non-homeostatic ingestion (food reward) are closely related.LHA is the central convergence point of the dopaminergic reward circuit and the hypothalamic energy balance regulating circuit, and PVN is the major relay site inanorexigenic signaling. GABAergic transmission from the ARC AgRP/NPY orexigenic neurons to PBN is important for non-homeostatic anorexia. The CNS nuclei and areas aredesignated anorexigenic (red), orexigenic (green) or integrating by their primary role based on e.g. lesion experiments. Directly or indirectly excitatory connections aredepicted by bright green links and directly or indirectly inhibitory connections by bright red links. PVN paraventricular nucleus, DMH dorsomedial hypothalamus, VMHventromedial hypothalamic nucleus, LHA lateral hypothalamic area, VTA ventral tegmental area, ARC arcuate nucleus, NTS nucleus tractus solitarii, PBN parabrachial nucleus,Nac nucleus accumbens. Modified from [289,397].

462 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

leptin and insulin occurs in obesity and type-II diabetes[118,163,287] as well as in inflammation in the CNS [352,451].

5.2.1.2. Impacts of TCDD on insulin and leptin. At doses high enoughto cause body weight loss, TCDD reduces circulating insulin levelsin rats [148,149]. At lethal doses, the change is similar to thatoccurring in pair-fed control rats and therefore seems to largelyrepresent a secondary effect [148]. At a sublethal dose, however,diminished feed intake apparently cannot account for the decrease

in plasma insulin [148]; indeed, a significant impairment of glu-cose-stimulated insulin secretion was reported in islets isolatedfrom TCDD-treated rats [310]. On the other hand, the function ofinsulin may be enhanced by TCDD. A recent study in 12-h fastedmice challenged with glucose revealed that already 24 h after a rel-atively low dose (10 lg/kg) of TCDD, plasma insulin levels in-creased less than in control animals, although plasma glucoseconcentrations behaved in the same way in both groups [238]. Inrats made diabetic with streptozotocin and a high-fat diet, TCDD

Page 12: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 463

ameliorated the induced hyperglycemia [125]. Likewise, in theonly study applying the ‘‘golden standard” method for the analysisof insulin sensitivity, euglycemic insulin clamp [270], glucose dis-posal rate was found to be increased by TCDD in rats, indicatingaugmented insulin sensitivity [308] (published in Japanese). Thesefindings thus imply that TCDD improves insulin sensitivity at thewhole organism level, at least in rats and mice. Unfortunately,there are no data available regarding the transport of insulin fromblood to the CNS in TCDD-exposed animals. Should that also be en-hanced, insulin could well be involved in the molecular pathogen-esis of the wasting syndrome despite the decrease in its circulatinglevels.

As to the possible influence of dioxin on leptin, there is a paucityof data in the literature. Unchanged plasma leptin levels were re-ported 24 h after a low dose of TCDD not expected to influencebody weight in rats (1 lg/kg) [310]. In a small-scale experimentat higher doses, we noted that plasma leptin concentration was ini-tially elevated on day 1 after TCDD exposure but then, three to se-ven days later, took an identical downhill course in both TCDD-treated L-E rats and their pair-fed controls [465]. However, in thosedays a rat-specific leptin antibody was not yet available and wehad to resort to an antibody raised for mouse leptin. Further stud-ies are clearly warranted to verify and extend these preliminaryfindings.

5.2.2. Signals inducing initiation of feeding5.2.2.1. Ghrelin. The best-characterized peripheral signal related tofood intake initiation is ghrelin, either considered to be a true hungersignal [364,516] or emphasizing its role as a feeding stimulatory sig-nal and a longer-term augmenting factor of feeding [37,92,191].Peripherally, ghrelin is mainly produced by endocrine X/A-like cellsin the gastric mucosa [83,221], and centrally by specialized nervecells in ARC. It is a highly potent stimulator of food intake in humansand experimental animals when given either peripherally or into theCNS, and its plasma levels rise before meals and fall shortly aftereating [37,364,516]. The relationship of the plasma ghrelin to foodintake stimulation may be direct [37,92,364] or conveyed via thevagus nerve with a possible involvement of de novo hypothalamicproduction of ghrelin [194].

5.2.2.2. Nutrient levels and feeding initiation. Neurons in the hypo-thalamus, elsewhere in the CNS and peripherally are capable ofsensing the levels of glucose, fatty acids and amino acids (leucine),although glucose through astrocyte-produced lactate is their al-most exclusive source of energy except for in a prolonged fast[243,397,453]. In addition, the CNS receives information of (atleast) peripheral glucose level through vagus, ghrelin and insulin.This information is primarily processed in the hypothalamus[243,397,453].

Direct nutrient sensing of foremost glucose is important inmaintaining the short-term energy stores at physiological levels,but it is also intricately intertwined with the long-term energy bal-ance regulation. A decrease in blood glucose below the euglygemiclevel triggers an immediate peripheral response starting with thesecretion of glucagon and elicitation of feeding, and this responsecan be prevented or induced by hypothalamic glucose administra-tion or depletion, respectively [243,453]. These temporary or long-er-lasting downward and upward alterations in plasma glucoselevels are sensed by glucose-inhibited and -excited neurons inthe hypothalamus and elsewhere in the CNS, and also relayed tothe CNS by at least insulin, ghrelin and vagal afferents [243,435,453,521]. These signals cause activity changes in the glucose-responsive neurons. The induced changes are most likely basedon KATP channels and type II glucose transporters (GLUT2), coupledintracellularly with mitochondrial uncoupling-protein 2 (UCP2),glucokinase and AMP-activated protein kinase (AMPK) [397,453].

The glucostatic/glucoprivic feeding initiation (possibly involvingghrelin [435]) is assumed to come into play in environmentally-induced non-physiological, acute energy depletion states [102]. Itis initiated by glucoreceptors in the brainstem (hindbrain), but itis mediated by several other brain regions including hypothalamicnuclei [383]. Although a slight drop in plasma glucose precedes thespontaneous initiation of feeding in rats and humans [61], it is con-sidered to be a feeding initiation-related phenomenon but not aprimary initiation signal per se.

5.2.2.3. Interference of TCDD with feeding initiation. No studies havehitherto addressed the possible role of ghrelin in the wasting syn-drome. On the other hand, glucoprivic challenges, brought abouteither with the competitive inhibitor of glycolysis at the glucose-6-phosphate isomerase step, 2-deoxy-D-glucose (2DG) [504], orwith a non-physiologically high acute dose of insulin, have beenimposed on TCDD-treated rats. A high but sublethal dose of TCDD(1000 lg/kg), which entirely prevented further growth of TCDD-resistant H/W rats, rapidly abolished the feeding response to gluco-privation whether induced by 2DG or insulin [340]. This changealso proved to be persistent as it was still complete for 2DG andpartial for insulin at 2 months after exposure [337]. The inabilityof TCDD-exposed rats to respond by eating to an acute energeticcrisis may be the key reason for their reported sensitivity to insulinlethality [149,337,340].

In C57BL/6 mice, TCDD at a dose (116 lg/kg) that did not affectthe overall body weight gain of the animals caused a swift (within24 h) and long-lasting (30 days) reduction (20–30%) in glucosetransport in the brain. GLUT1 protein expression was similarlydampened (GLUT2 was not analyzed) [260]. A total subdiaphrag-matic vagotomy failed to appreciably modify the effect of TCDDon feed intake or body weight in L-E or H/W rats [463], arguingagainst a decisive role for vagal signals in the wasting syndrome.It is of interest, however, that vagotomy had an additive diminish-ing impact on body weight which resembled that of a dorsomedialhypothalamus (DMH) lesion (see Section 5.4.1.3). The fact thatvagotomy results in body weight loss has been known for a longtime; exactly why this happens is still obscure [415].

5.3. Short-term (meal-to-meal) regulation of feeding: peripheralpeptides

When feeding has commenced, a number of messengers emergeto suppress appetite in response to food ingestion and act as sati-ation signals. These satiation signals are small peptides producedin the intestine (e.g. cholecystokinin [CCK], glucagon-like peptide1 [GLP-1], oxyntomodulin, peptide-YY [PYY], enterostatin) or inthe pancreas (amylin and pancreatic polypeptide). Most of themact primarily or partially via vagal afferent fibers [33,54,68,99,200,222,387]. Experimentally, they are defined as substances thatresult in a reduction of the meal size when given at the start offeeding. They also work at physiological concentrations and with-out generating malaise or incapacitation [514]. Traditionally, thesatiation signals were considered to be short-acting and only rele-vant to a single meal, but at least some of them are currentlythought to have a role in longer-term energy balance regulationas well [37,516]. In contrast to the long-term satiety signals insulinand leptin, at least PYY and GLP-1 are able to cross the blood–brainbarrier via non-saturable mechanisms [519].

5.3.1. Contribution of the short-term signals to TCDD-induced wastingFrom some 2 weeks on after exposure to 1000 lg/kg TCDD, H/W

rats started to exhibit a feeding behavioral change which wasinterpreted as augmented sensitivity to post-ingestive satiety (orsatiation) signals. It manifested as reduced consumption of glucoseor sucrose solution but increased consumption of the energetically

Page 13: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

464 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

valueless saccharine solution, and diminished feeding in pre-load-ing experiments if the pre-loads yielded energy [336,341]. Thesealterations persisted for at least several months. However, con-comitantly the feeding response to peripherally administeredCCK remained unaffected [337,341]. As noted earlier, vagotomyfailed to modulate the effects of TCDD on feed intake or bodyweight in either L-E or H/W rats; thus it is unlikely that these ef-fects were mediated by vagal afferents. Likewise, portocaval anas-tomosis did not appreciably interfere with the effect of TCDD ondaily feed consumption or body weight gain in these rat strains[463].

5.4. Regulatory centers and neurotransmitters in the CNS involved inenergy balance (Fig. 5B)

5.4.1. Hypothalamic nuclei and neuropeptides5.4.1.1. ARC at the crossroads of insulin and leptin signaling. Energyhomeostasis and body weight are now thought to be regulatedby neuronal circuits, which signal using specific neuropeptides,rather than by specific hypothalamic nuclei. However, the ARC, inparticular, is thought to play a pivotal role in the integration of reg-ulatory signals. As described earlier (Section 5.2.1.1), informationon the amount of adipose tissue, reflecting the long-term energystatus, is mainly signaled to the CNS by leptin and insulin [118].Although they both have receptors at multiple sites in the CNS,those concerned with energy balance reside for the most part inthe hypothalamus, particularly in ARC, ventromedial hypothala-mus (VMH) and DMH, as well as in the lateral hypothalamic area(LHA) and PVN [223,243,301].

In ARC, the effects of insulin and leptin are closely integrated,even intracellularly at the level of phosphoinositide 3-kinase, tomodulate two populations of neurons acting as the cardinal regula-tors of energy balance [118,243,397]: the orexigenic agouti-relatedpeptide/neuropeptide Y (AgRP/NPY) neurons reduce their activityand neurotransmitter expression by leptin and insulin stimulation,whereas the anorexigenic pro-opiomelanocortin/cocaine- andamphetamine-regulated transcript (POMC/CART)-expressing neu-rons function in the opposite way. However, the effects of insulinand leptin are not fully identical. While leptin depolarizes POMC/CART neurons to increase their firing rate and hyperpolarizesAgRP/NPY neurons to inhibit their electrical activity, insulin hyper-polarizes and silences both these two types of neurons by mecha-nisms not totally resolved [175,289,397].

Both AgRP/NPY and POMC/CART neurons express the inhibitoryneurotransmitter gamma-amino butyric acid (GABA), and POMC/CART neurons appear to be under constant suppression by theirneighboring AgRP/NPY-neurons through the release of GABA[132,276]. ARC also receives and projects various inhibitory andexcitatory contacts to other areas in the hypothalamus, and thesesynapses as well as the synapses amongst neurons in ARC arefound to be morphologically modulated by ghrelin and leptinadministration or fasting [6,289,397]. This synaptic or neuronalplasticity (based at least partly on the continuous activity of AgRPneurons) is proposed to be an important mechanism in hypotha-lamic energy balance regulation controlled by leptin/ghrelin bal-ance and explaining the long-term effects of ghrelin on foodintake [6,92].

In addition to leptin, insulin and ghrelin, plasma nutrient levelsinfluence the activity of orexigenic and anorexigenic neurons inARC. Glucose inhibits AgRP/NPY neurons and activates POMC/CARTneurons, while intracerebroventricularly infused oleic acid (a long-chain fatty acid) reduces NPY expression in the hypothalamus, anda protein-rich meal activates POMC/CART neurons [244,353,397].

5.4.1.2. Hypothalamic centers and networks outside the ARC. Beyondthe ARC, AgRP/NPY and POMC/CART neurons continue to have clo-

sely intertwined connections. Both of them innervate PVN, LHA,VMH and DMH as well as several other hypothalamic and extra-hypothalamic nuclei, notably the parabrachial nucleus (PBN)[6,51]; POMC/CART neurons extend their connections in the brain-stem also to the nucleus tractus solitarii (NTS) [430]. Of the inner-vated nuclei and areas, the VMH, PVN, LHA and DMH have beenassigned special roles in hypothalamic food intake regulation andexternal connections. Based on lesion studies, the VMH has tradi-tionally been referred to as the satiety center, and it is still todayassumed to have an independent role in food intake promotion,responding to at least glucose and leptin [300,397]. The anorexi-genic PVN and the orexigenic LHA are important relay sites ofinformation for autonomic and neuroendocrine effects and for foodintake with afferent connections to the brainstem. LHA has alsoconnections to cortical neurons and is an integrating nucleus ofthe energy balance and reward circuits [118,289,300]. The food in-take-promoting DMH is innervated by NPY/AgRP-positive fibersfrom ARC and responds directly to at least glucose and leptin. Ithas an important autonomous role in food intake and body weightregulation [32,218,532].

The strongly orexigenic NPY signals through Y1, Y2 and Y5receptors, and LHA seems to be the principal site of its action eventhough it also inhibits neuronal firing in PVN and VMH [6,69]. Theorexigenic effect through LHA is generally believed to be mediatedby melanin concentrating hormone (MCH), with a shorter-terminfluence by orexin/hypocretin. However, the mechanism by whichthe MCH secretion is stimulated is unknown [51,155], albeit recentfindings point to a marked effect of LHA GABAA receptors in feed-ing regulation [474]. The anorexigenic melanocortin system isbased on melanocyte-stimulating hormone (a-MSH), producedfrom POMC in POMC/CART neurons and secreted together withCART [6,92]. The principal second order nucleus in melanocortinsignaling is PVN, which contains high levels of melanocortin recep-tors 3 and 4 (MC3R, MC4R) stimulated by a-MSH [6]. It is also animportant convergence point of anorexigenic and orexigenic sig-naling, since AgRP, the second product of AgRP/NPY neurons, isan antagonist of melanonocortin receptors, and (as described ear-lier) CRF exerts its anorexic effect in the PVN.

5.4.1.3. Influence of hypothalamic lesions on the wasting syn-drome. The hypothalamus has been known for decades to be akey brain region in the regulation of feeding and energy balancewith the earliest publications dating back to the early 1950s [17].Lesions of hypothalamic nuclei are rather massive and indiscrimi-nate manipulations of brain regulatory sites as it has turned outthat the notion of distinct ‘hunger or satiety centers’ is far too sim-plistic, as outlined in the previous sections. Nevertheless, it is well-established that site-specific lesions of critical areas in this net-work (such as VMH or DMH) can markedly alter the defended bodyweight [249].

A lesion of the VMH induces metabolic obesity which resultsnot only from hyperphagia but also from disturbances in the auto-nomic nervous system [128,215]. The lesion usually involves ARCin addition to the actual ventromedial nucleus [249]. It was of greatinterest to see whether the metabolically counteracting impact ofVMH lesion could overcome the catabolic dyscrasia elicited byTCDD. It turned out that when VMH lesion preceded TCDD treat-ment, it surprisingly aggravated TCDD-induced wasting syndromein both TCDD-sensitive L-E and TCDD-resistant H/W rats [467]. Theexacerbation was greater if the rats had reached obesity beforeTCDD exposure (i.e. if exposed 2 weeks versus 4 days after theoperation). The doses of TCDD used in that experiment wereclearly sublethal to sham-operated rats and the response to VMHlesion without TCDD was hyperphagia and obesity. Interestingly,a single lesioned L-E rat responded to TCDD disparately by gainingweight at a tremendous pace and weighing 60% more at 6 weeks

Page 14: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 465

than the obese lesioned control rats. In the case when the rats werelesioned 6 weeks after TCDD exposure, at a time when body weighthad stabilized to a lower level, no aggravation was seen any longer(the lesioned TCDD-exposed rats grew slightly faster than thesham-operated TCDD-exposed partners). These findings suggestprotective involvement of regulatory networks either traversingthe VMH (including ARC) or operating in it at the initiation phaseof the wasting syndrome.

Subsequently, the possible interference of DMH and PVN lesionswith TCDD toxicity was also investigated in our laboratory on H/Wrats treated with a high but non-lethal dose of TCDD (1000 lg/kg).To maximize the effect of PVN lesion on body weight, the rats inthat experiment were on a highly palatable diet. Both lesions gen-erated the expected outcome in vehicle-treated rats (retarded orenhanced body weight gain for DMH and PVN lesions, respec-tively). There was no interaction between the effects of DMH lesionand TCDD, they were purely additive. In the case of PVN, in con-trast, the lesioned rats that initially were overweight diminishedtheir feed consumption more than the sham-operated counter-parts, with the result that 7 weeks after TCDD exposure the bodyweights of the two groups were indistinguishable [469,470]. Thus,neither of these two regulatory centers in the hypothalamus ap-pears to be directly involved in TCDD’s actions on feed intakeand body weight. The studies with TCDD further suggest that a le-sion of the DMH, but not that of the PVN, is able to reset the pon-derostat because the ultimate body weight level pursued by TCDD-treated rats was only affected by the former manipulation. Itshould be borne in mind that these experiments were carried outin TCDD-resistant H/W rats which do not exhibit the fatal formof the wasting syndrome.

5.4.1.4. Effects of TCDD on hypothalamic neuropeptides. At 6 daysafter a sublethal TCDD dose (15 lg/kg) to Sprague–Dawley rats,an increased expression of NPY, POMC and CART mRNA was de-tected in ARC, and in LHA also MCH expression was increased[117]. In keeping with these findings, a dose (50 lg/kg) close tothe LD50 value of the same strain of rat increased POMC mRNA lev-els in ARC at 7 and 14 days; at the latter time-point, a significantelevation was also found at 12.5 and 25 lg/kg doses [283]. In atime-course study spanning 6–96 h after TCDD exposure (50 lg/kg), a tendency to lowered levels in mRNA expression of orexigenicfactors was initially seen in hypothalamic blocks of TCDD-sensitiveL-E rats. By 96 h, this had converted to an increased expression ofNPY and AgRP, probably as a secondary change to TCDD-inducedhypophagia. Similar general patterns, although delayed, occurredin TCDD-resistant H/W rats [258]. Thus, there is limited evidenceof direct interference of TCDD with the regulatory neuropeptides.More information is needed especially on the early stages of dioxinintoxication.

5.4.2. Amine neurotransmittersThe classical monoamine neurotransmitters serotonin (5-

hydroxytryptamine, 5-HT), dopamine and noradrenaline as wellas the biogenic amine histamine act in conjunction with neuropep-tides and peripheral hormones to control energy balance. Seroto-nin has overall a suppressive effect on food intake and bodyweight e.g. in PVN, VMH and DMH acting mainly through 5-HT1B

and possibly 5-HT1C receptors [372]. The autoreceptor 5-HT1A hasthe opposite impact on feeding mediating hyperphagia [81]. Ithas recently even been proposed that leptin may exert its actionon food intake by inhibiting the production of serotonin in brain-stem neurons. This would then curtail appetite through reducedactivity of 5-HT1A in ARC leading to increased POMC expressionand decreased NPY and AgRP expression [520]. Serotonin is syn-thesized from tryptophan whose brain concentrations are rate-lim-iting for the synthesis [64].

Dopamine appears to be more influenced by food intake ratherthan influencing it [372]. However, mice which lack dopamine, dueto the absence of the tyrosine hydroxylase gene, have fatal hypo-phagia [519]. Noradrenaline may have dual effects on food intake:in PVN activation of a2 receptors stimulates feeding whereas a1

agonists suppress it [372]. Histamine is mainly involved in circa-dian feeding rhythms mediating or accentuating the effects of lep-tin via H1 receptors [193,524].

GABA is the main inhibitory neurotransmitter in the CNS. Re-cent findings underline the importance of GABAergic transmissionof the AgRP/NPY-neurons in maintaining feeding [517,518]. Loss ofGABAergic transmission of these cells to PBN leads to activation ofpostsynaptic neurons and starvation, and this has been proposed tobe related to the activation of circuits that normally promote nau-sea-induced anorexia. As described above, GABA is orexigenic byway of its inhibitory action on POMC/CART cells in the ARC whereboth GABAA and GABAB receptors have been demonstrated. Most ofthe AgRP/NPY neurons also contain the GABA synthesizing en-zyme, glutamic acid decarboxylase [276]. Glutamate, in turn, isthe dominant excitatory transmitter in hypothalamic neuroendo-crine regulation. When injected into the LHA, it elicits an intensefeeding response in satiated rats via activation of MCH and orexinneurons [276].

5.4.2.1. Effects of TCDD on amine neurotransmitters in the CNS. At 4and 10 days after a single dose of 50 lg/kg, TCDD was found to in-crease serotonin turnover uniformly across all macrodissectedbrain areas, including the hypothalamus, in L-E but not H/W rats[477]. The effect was specific to serotonin as there were no majorchanges in brain concentrations or turnover rates of noradrenalineand dopamine in either strain. This general pattern has been veri-fied in other rat strains (reviewed in [338]). The enhanced turnoveris due to increased plasma levels of free tryptophan which are re-flected in brain tryptophan concentrations [480]. Although it hasproven to be a reliable indicator of acutely toxic exposure to TCDDand other dioxins in rats [479], accelerated brain serotonin turn-over does not appear to be causally related to the wasting syn-drome, because depletion of brain serotonin by the selectiveneurotoxic analog, 5,7-dihydroxytryptamine, did not influence it[439]. In agreement with this, plasma and brain tryptophan werenot altered in TCDD-exposed guinea pigs in the face of markedbody weight loss while both were increased in TCDD-treated ham-sters with little effect on body weight [481].

A slight increase was recorded in histamine concentrations inthe whole hypothalamus at 28 h [464] and in the median eminenceat 25 h after TCDD treatment [471]. They might bear on the alteredcircadian feeding rhythms observed in TCDD-exposed rats (seeSection 5.7).

Fully supporting the reasoning above on the role of monoam-ines in TCDD toxicity, intervention with haloperidol or amperozide(dopamine antagonists), phenoxybenzamine (a-adrenoceptorblocking agent) or p-chlorophenylalanine (serotonin synthesisinhibitor) failed to modulate TCDD-induced wasting syndrome inL-E rats [339].

There is evidence from prenatally exposed animals that dioxinsmay affect GABAergic and glutamatergic neurotransmission in theCNS. For example, in rats the mRNA expression of glutamate recep-tor subunits NR2B (NMDA) and GluR1 (AMPA) was reduced whilethat of NR2A (NMDA) was enhanced in cortex and in the hippo-campus postnatally after prenatal TCDD exposure (0.1–0.8 lg/kg,GD15) [178,196]. TCDD administration on GD11.5 at a dose of5 lg/kg to the mouse dam was found to compromise the differen-tiation of GABAergic neurons in ventral telencephalon, whichmight offer an explanation for the experimentally observed, usu-ally subtle, learning alterations (either retardation or facilitation)in gestationally TCDD-exposed rats and monkeys [145,404,506].

Page 15: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

466 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

Furthermore, TCDD was postulated to induce the reported alteredmale sexual behavior and female-type LH-surge [197], ‘‘demascul-inization”, by interfering with the estradiol-dependent apoptosis ofGABA/glutamate neurons of the anteroventral periventricular nu-cleus in the preoptic area [166,327]. In vitro studies with neuronalcells have also revealed that TCDD may increase cellular calciumconcentration and activate protein kinase-C through NMDA recep-tors [213,254], and up-regulate NMDA receptor subunits (NR1,NR2A and NR2B) [74]. However, next to nothing is known aboutthe possible modulation of GABAergic or glutamatergic neuro-transmission by dioxins in the context of energy balance in adultanimals.

5.4.3. Nitric oxideIn recent years, the importance of nitric oxide in the regulation

of energy balance has emerged. The orexigenic peptides ghrelin,NPY and orexin-A all exert their effects on food intake, at least inpart, through nitric oxide which is synthesized by neuronal nitricoxide synthase (nNOS) [112,139,285]. Conversely, leptin inhibitsnNOS and thereby reduces nitric oxide levels [285]. Another iso-form, inducible NOS (iNOS), has been suggested to play an antago-nistic role to that of nNOS by mediating anorexia [286] and may beinvolved in cancer cachexia [494].

5.4.3.1. Involvement of nitric oxide in TCDD toxicity. Administrationof the irreversible selective inhibitor of nNOS, N-nitro-L-arginineto TCDD-treated animals either with slow-release osmotic mini-pumps or by repeated injections caused an interesting divergencein responses between mice and rats. In mice, the interventionattenuated the acute toxicity of TCDD, whereas in rats it exacer-bated it. However, in feed intake a departure from placebo-admin-istered, TCDD-treated rats was only seen during the first 5 days(feed intake was not measured in mice) [487]. In another study,a decrease in whole brain nNOS protein was first detected by Wes-tern blot in Long-Evans rats at one and 2 weeks post-exposure to50 lg/kg TCDD. Next, NAPDH-diaphorase staining was used toselectively analyze nNOS levels by immunohistochemistry inhypothalamic nuclei of identically treated rats of the same strain.Diminished staining was recorded in PVN, LHA and perifornical nu-cleus at 2 weeks after TCDD with a downward tendency at 1 week[72]. Thus, further studies are warranted on both iNOS and nNOS inTCDD-treated rats and mice.

5.4.4. Caudal brainstem in energy balance regulationThe caudal brainstem (‘‘hindbrain”), the NTS in particular, is the

main site to receive information from peripheral energy metabo-lism, chiefly by way of the vagal afferent fibers. The caudal brain-stem is involved in the control of meal size by signals arisingfrom the mouth and gastrointestinal tract based upon the chemicaland mechanical properties of food which are conveyed to the NTS,an important integrator in energy balance regulation [37,132,152].Caudal brainstem circuits are sufficient for meal size control, initi-ate glucoprivic feeding, contain systems for satiating signal (CCKand GLP-1) and leptin interaction, and have multiple afferent con-nections to the hypothalamus [152,383].

The caudal brainstem has not been addressed in the context ofthe impacts of TCDD on energy homeostasis. However, the promi-nent effects of TCDD on glucoprivic feeding (Section 5.2.2.3) andmeal size (Section 5.7) would definitely justify more attention tobe paid on this brain region in the future.

5.5. Satiation versus reward in the regulation of energy balance

The central system integrating information on the amount ofadipose tissue, blood nutrient levels and environment resides inhypothalamic and hindbrain nuclei, amongst a myriad of connec-

tions with other brain areas involved in motivation, reward andcognition. While this organization is generally accepted and thor-oughly described, the primacy of the central integration as opposedto a more decentralized regulation is debated [118,289,422].

In satiation-based food intake regulation, the energy balance setsthe general sensitivity of the hypothalamus – hindbrain system tosatiety signals affecting the meal size. An abundance of stored andcirculating nutrients enhances sensitivity to the meal terminatingsignals [289,300,514]. On the other hand, the incentive (reward)value of palatable food can override satiety signals and promoteexcess food ingestion. Aptly, this regulation of food intake is alsocalled non-homeostatic. The major components of reward are want-ing, liking and learning [35], with partly overlapping and only par-tially known CNS circuits and signals [118,223,300,422].

The reward circuitry is complex and involves interactions be-tween several signaling systems. The reward-based regulation ofeating is intimately linked to the midbrain mesolimbic dopaminer-gic (DA) system which contains a set of ventral tegmental area(VTA) neurons that innervate the striatum, amygdala and prefron-tal cortex [300,422]. Leptin and insulin act directly on VTA DA neu-rons reducing dopamine release and diminishing food seeking andintake [118], whereas food restriction and ghrelin induce an oppo-site effect [223,422]. Notably, a chronic absence of leptin impairsthe mesolimbic DA system [300]. LHA orexin neurons project toVTA and the stimulation of LHA promotes strongly food (and otherreward, e.g. drug) seeking, and this type of behavior is suppressedby leptin [118,300,422]. Subconscious ‘‘liking-reaction” to food canbe elicited by hindbrain neural circuits and seems to be related tothe cannabinoid and l-opioid systems with the involvement of thestriatal nucleus accumbens (NAc) and ventral pallidum [422]. NAchas also been shown to inhibit MCH neurons in LHA, which sendinhibitory efferents back to NAc, conceivably enabling continuousactivation of MCH-containing neurons [399]. Leptin and ghrelinhave both been shown to augment memory and learning, whilethey have opposite effects on food intake [422]. Opioids play animportant role in the reward circuitry, as a lack of either encepha-lin or b-endorphin in mice abolishes the reinforcing property offood, regardless of the palatability of the food tested [519].

5.5.1. Effects of TCDD on the reward circuitryOriginally, increased c-Fos expression (a widely-used index of

neuronal activation) was detected in several hypothalamic nuclei,central amygdala and BNST 3–4 days after TCDD (50 lg/kg) admin-istration to rats [71]. Later on (using an identical exposure setting),the same group reported an increase of methionine–encephalinimmunoreactivity in various forebrain nuclei, e.g. central amyg-dala, PVN, medial preoptic nucleus and BNST 2 weeks after TCDDtreatment [70]. This could be a compensatory reaction to the bodyweight loss and hypophagia provoked by TCDD. The opioid antag-onist naloxone decreased 24-h fast-induced feeding in control rats.In H/W rats treated with a high dose of TCDD (1000 lg/kg) 3months earlier, this suppression was blunted (the circadian feedingrhythm of the TCDD-treated rats was shifted (see Section 5.7), andthus they ate more than their corn oil-treated partners during the6-h observation period at daytime whether given saline or nalox-one) [337]. Hence, there is little evidence of a critical role for thereward circuitry in the TCDD-induced wasting syndrome, but theavailable data are also scant.

5.6. Enhanced neophobia in TCDD-treated animals

TCDD appears to affect feeding behavior in a specific manner.Certain alterations in feeding behavior emerge as early as a fewhours after TCDD exposure [473]. The selective nature of these im-pacts is revealed by the fact that other than behavioral effects onfeeding, few behavioral alterations are observed in TCDD-exposed

Page 16: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 467

adult animals [429]. Reported disturbances in the behavior do ex-ist, but they originate from prenatal and/or lactational exposure todioxins [16,45,420,506].

Exposure of adult rats to TCDD not only reduces total feed con-sumption but also results in a conspicuous and peculiar alterationin food preferences in certain contexts. Usually rodents like sweetand fatty tastes and therefore prefer cheese and chocolate to regu-lar chow, but if these new items are first introduced to the rats atTCDD exposure or soon afterwards, TCDD-treated rats display astriking avoidance of them. This change emerges rapidly, in a fewhours after exposure. Furthermore, the neophobia is not sex- orstrain-specific and bears no apparent correlation with sensitivityto the acute toxicity of TCDD. The deviant behavior, avoidance ofa tasty food item (chocolate), was observable at sublethal dosesand it persisted for at least 5 weeks [473].

Recently, we have further pursued these behavioral studies andour preliminary data show that the phenomenon is not confined torats but also occurs in TCDD-treated mice. In them, functional AHRseems to be needed for the neophobia (S. Lensu, R. Pohjanvirta, J.Tuomisto and J.T. Tuomisto, unpublished results). Moreover, theavoidance of novel food items is detectable at doses which are solow that they have no effect on body weight or total energy intakein rats. Therefore, this accentuated neophobia-like syndrome ap-pears to be an extremely sensitive toxic endpoint which emergesat a notably early phase of TCDD intoxication and warrants furtherstudies.

Gustatory or food neophobia means an innate fear of new foods,and it exists in humans and animals [97,119]. The abstaining ofingestion of large amounts of a new, potentially toxic food maybe genetically determined and beneficial for survival [184]. Gusta-tory neophobia is mediated by insular cortex [386] and basolateralamygdala (BLA) [376], and glutamate NMDA receptors participatein its mediation [119]. Conditioned taste aversion, an efficientlearned aversion of a novel taste associated with transient visceralillness, is mediated by the same CNS areas, principally BLA. As de-

Table 1Experimental evidence that TCDD may affect the central regulation of energy balance.

Factor/systema

Comments

CRF CRF, TRF and synaptic plasticity have been implicated as primary tashown to alter CRF mRNA expression in the CNS

Insulin TCDD reduces directly or indirectly plasma insulin and impairs islet cinsulin sensitivity. For insulin lethality, see ‘‘blood glucose”.

Leptin Small doses of TCDD did not induce changes in leptin levels but alt

Blood glucose Central and peripheral glucose depletion induces feeding. This respoby TCDD, possibly explaining sensitivity to insulin lethality

Satiatingsignals

After a relatively high dose (1000 lg/kg) of TCDD H/W rats exhibiteenhanced response to post-ingestive satiation signals. However, thewas not altered

VMH VMH lesion (inducing metabolic obesity) aggravated the wasting sybefore TCDD administration, but did not have a marked effect after

DMH DMH lesion (inducing hypophagia and suppressing 2DG-elicited feedweight in H/W rats

Neuropeptides In SD rats, a sublethal dose of TCDD increased NPY, CART, POMC andComparable doses increased POMC expression in ARC at 14 days, wPOMC expression in ARC both at 7 and 14 days. In a study on L-E rainduced an initial lowering tendency in the mRNAs of hypothalamiincreased expression of NPY and AgRP

Neophobia Avoidance of a novel tasty food item was observed in TCDD-treatedleast 5 weeks. Recent unpublished findings extend this phenomenonavoidance of novel food items seems to be an exceptionally sensitiv

Feedingrhythm

TCDD treatment has been shown to alter the normal circadian pattefindings point to strain- and dose-dependent dynamic chains of alte

a See also Fig. 5 A and B.

scribed earlier (Section 5.4.2.1), there is mounting evidence thatTCDD can interfere with the glutamate receptors by prenatal expo-sure. The augmented neophobia-like behavior of TCDD-treated ani-mals reinforces the need for information on whether this alsooccurs in the case of adult exposure.

5.7. Disrupted circadian feeding rhythm in TCDD-exposed rats

TCDD treatment has been shown to alter the normal circadianpattern of feeding in rats. About 2 weeks after exposure to a highbut sublethal dose of TCDD (1000 lg/kg), TCDD-resistant H/W ratsstarted to exhibit an increase in their proportional feed intake duringthe light hours of the day. This shift persisted for over 4 months[336,337]. Similar findings have been reported in TCDD-sensitiveL-E [339] and Sprague–Dawley rats [75]. To further investigate thisand other possible aberrations in consummatory behaviors, we haverecently set up an automated system enabling continuous monitor-ing of feed and water intake to analyze feeding and drinkingmicrostructures in both L-E and H/W rats. We have found strain-and dose-dependent dynamic chains of alterations to emerge soonafter exposure. The preliminary data have revealed that in L-E ratsat both supralethal (100 lg/kg) and sublethal (10 lg/kg) doses ofTCDD, the decrease in the total amount of feed eaten results from areduced size of meals with the reduction being larger at night thanat daytime. In contrast, in H/W rats the decrease was due to a low-ered number of meals throughout the day. We further observed thatof the two circadian peaks in the feeding, which in control rats takeplace shortly after lights off in the evening and just before lights on inthe morning, the morning peak diminished in TCDD-treated L-E rats(S. Lensu, J. Lindén, P. Tiittanen, J. Tuomisto and R. Pohjanvirta,unpublished data).

The reason(s) for these readjustments in feeding rhythms arestill obscure. Several biological phenomena including feedinghave circadian rhythms controlled by clock genes which belongto the bHLH/PAS protein superfamily [211]. Expression of the

References

rgets of energy balance regulation. TCDD was [6,60,92,106,246,283]

ell insulin secretion, but increases whole-body [125,148,149,238,270,308,310]

erations were reported at higher doses [310,465]

nse (elicited by 2DG or insulin) was abolished [149,337,340]

d feeding behavioral changes interpreted as anir response to peripherally administered CCK

[336,337,341]

ndrome in both L-E and H/W rats if performedTCDD-induced body weight reduction

[467]

ing) had an additive effect with TCDD on body [32,470]

MCH mRNA expression in ARC or LHA at 6 days.hile a dose close to the LD50 value increasedts focusing on days 1–4, a lethal dose of TCDD-c orexigenic factors. This was followed by an

[117,258,283]

rats at sublethal doses and it persisted for atto mice and attest to its AHR-dependency. Thee response to TCDD in rats

[473] Unpublished: seeSection 5.6

rn of feeding in rats. Recent unpublishedrations soon after TCDD exposure

[75,336,337] Unpublished: seeSection 5.7

Page 17: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

468 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

AHR protein fluctuates in several tissues of mice and rats in thecourse of the day [293,378]. Although the physiological role ofthe AHR in circadian rhythms is still debatable, activation of itby TCDD leads to disturbances in the circadian time-keeping sys-tem. Recent findings support a role for the clock genes Sim1,Per1 and Per2 in the modulation of AHR-mediated responses toTCDD in different organs in vivo [293,367,368] and in vitro[133,522]. However, in a study by Korkalainen et al. [224], nosignificant effects of TCDD were found on the hypothalamicmRNA levels of Sim1 or Per2 in L-E or H/W rats.

Table 1 summarizes the available evidence of the interference ofTCDD with the central regulation of energy balance.

6. Conclusions and future prospects

Despite extensive and numerous studies over the last three dec-ades, the ultimate reason for the extremely high acute toxicity ofcertain dioxins, foremost TCDD, has remained enigmatic. The acutetoxicity of TCDD is intimately coupled with the wasting syndrome,which in itself poses an interesting challenge to science. TCDDseems to target the regulatory mechanisms of body weight andfood intake in a highly specific manner as evidenced by the abilityand tendency of animals treated with sublethal doses of TCDD todefend their lowered body weight against external manipulations,by the irreversibility of TCDD-induced retardation in body weightgain and by the swift emergence and remarkable sensitivity ofthe neophobia-like response upon TCDD exposure. These featuresrender the wasting syndrome a potentially useful model, and TCDDan efficient tool, for physiological studies on the regulatory mech-anisms of energy balance.

The most resistant laboratory animals to the acute toxicity ofTCDD are hamsters and H/W rats. As seasonal animals, hamstersare genetically well-adapted to large variations in their bodyweight and may thereby tolerate the set-point-lowering effect ofTCDD. Interestingly, the behavioral response in H/W rats may alsobe unique: at very high doses of TCDD, they will often commence atotal fast which may last up to 11 consecutive days, and then rap-idly resume eating [331,467], whereas rats of TCDD-sensitivestrains keep on consuming small amounts of food until death.Since both hamsters and H/W rats possess AHRs that are remod-eled at the transactivation domain, the involvement of the AHRin the regulation of energy balance will be worth exploring.

The research of the wasting syndrome also has important bearingon the pathogenesis of diseases in which the control systems forbody weight are gravely affected such as anorexia nervosa and can-cer anorexia. The present view is that all the major toxic impacts ofdioxins are mediated by the AHR through its canonical signalingpathway. It seems, however, that the response pattern elicited byactivation of this pathway may depend on the ligand or the combina-tion of the AHR and its ligand (the same ligand might trigger a differ-ent response spectrum in different species due to variations in AHRstructure). In the future, this could pave the way to the developmentof compounds which might possess the potentially beneficial effectof dioxins on body weight (apparent lowering of the set-point level)but would be devoid of its toxicity. It is thus clear that there is still alot to explore in dioxins, the AHR and the central regulatory systemsof energy balance for years to come.

Acknowledgments

We thank Prof. Allan B. Okey for kindly allowing us to use one ofhis original figures of the AHR as the core for Fig. 3. This work wasfinancially supported by grants from the Academy of Finland(Grant Number 123345 [R.P.]) and from the Finnish Cultural Foun-dation, Regional fund of Northern Savo (S.L.).

References

[1] A. Abballe, T.J. Ballard, E. Dellatte, A. di Domenico, F. Ferri, A.R. Fulgenzi, G.Grisanti, N. Iacovella, A.M. Ingelido, R. Malisch, R. Miniero, M.G. Porpora, S.Risica, G. Ziemacki, E. De Felip, Persistent environmental contaminants inhuman milk: concentrations and time trends in Italy, Chemosphere 73 (Suppl.1) (2008) S220–S227.

[2] B.D. Abbott, L.S. Birnbaum, G.H. Perdew, Developmental expression of twomembers of a new class of transcription factors: I. Expression of arylhydrocarbon receptor in the C57BL/6N mouse embryo, Dev. Dyn. 204 (1995)133–143.

[3] B.D. Abbott, M.R. Probst, Developmental expression of two members of a newclass of transcription factors: II. Expression of aryl hydrocarbon receptornuclear translocator in the C57BL/6N mouse embryo, Dev. Dyn. 204 (1995)144–155.

[4] B.D. Abbott, J.E. Schmid, J.A. Pitt, A.R. Buckalew, C.R. Wood, G.A. Held, J.J.Diliberto, Adverse reproductive outcomes in the transgenic Ah receptor-deficient mouse, Toxicol. Appl. Pharmacol. 155 (1999) 62–70.

[5] P.A. Abbott, R.V. Bonnert, M.V. Caffrey, P.A. Cage, A.J. Cooke, D.K. Donald, M.Furber, S. Hill, J. Withnall, Fused mesoionic heterocycles: synthesis of[1,2,3]triazolo[1,5a]quinoline, [1,2,3]triazolo[1,5a]quinoxaline and [1,2,3]-triazolo[5,1-c]benzotriazine derivatives, Tetrahedron 58 (2002) 3185–3198.

[6] A. Abizaid, T.L. Horvath, Brain circuits regulating energy homeostasis, Regul.Peptides 149 (2008) 3–10.

[7] J. Adachi, Y. Mori, S. Matsui, H. Takigami, J. Fujino, H. Kitagawa, C.A. Miller 3rd,T. Kato, K. Saeki, T. Matsuda, Indirubin and indigo are potent aryl hydrocarbonreceptor ligands present in human urine, J. Biol. Chem. 276 (2001) 31475–31478.

[8] W.J. Adams, K.M. Blaine, A Water Solubility Determination of 2,3,7,8-Tcdd,Chemosphere 15 (1986) 1397–1400.

[9] M.H. Aitola, M.T. Pelto-Huikko, Expression of Arnt and Arnt2 mRNA indeveloping murine tissues, J. Histochem. Cytochem. 51 (2003) 41–54.

[10] E. Akahoshi, S. Yoshimura, M. Ishihara-Sugano, Over-expression of AhR (arylhydrocarbon receptor) induces neural differentiation of Neuro2a cells:neurotoxicology study, Environ. Health 5 (2006) 24.

[11] S. Alaluusua, P. Calderara, P.M. Gerthoux, P.L. Lukinmaa, O. Kovero, L.Needham, D.G. Patterson Jr., J. Tuomisto, P. Mocarelli, Developmental dentalaberrations after the dioxin accident in Seveso, Environ. Health Perspect. 112(2004) 1313–1318.

[12] S. Alaluusua, P.L. Lukinmaa, R. Pohjanvirta, M. Unkila, J. Tuomisto, Exposure to2,3,7,8-tetrachlorodibenzo-para-dioxin leads to defective dentin formationand pulpal perforation in rat incisor tooth, Toxicology 81 (1993) 1–13.

[13] S. Alaluusua, P.L. Lukinmaa, J. Torppa, J. Tuomisto, T. Vartiainen, Developingteeth as biomarker of dioxin exposure, Lancet 353 (1999) 206.

[14] S. Alaluusua, P.L. Lukinmaa, T. Vartiainen, M. Partanen, J. Torppa, J. Tuomisto,Polychlorinated dibenzo-p-dioxins and dibenzofurans via mother’s milk maycause developmental defects in the child’s teeth, Environ. Toxicol. Pharmacol.1 (1996) 193–197.

[15] D.L. Alexander, L. Zhang, M. Foroozesh, W.L. Alworth, C.R. Jefcoate,Metabolism-based polycyclic aromatic acetylene inhibition of CYP1B1 in10T1/2 cells potentiates aryl hydrocarbon receptor activity, Toxicol. Appl.Pharmacol. 161 (1999) 123–139.

[16] S. Amin, R.W. Moore, R.E. Peterson, S.L. Schantz, Gestational and lactationalexposure to TCDD or coplanar PCBs alters adult expression of saccharinpreference behavior in female rats, Neurotoxicol. Teratol. 22 (2000) 675–682.

[17] B.K. Anand, J.R. Brobeck, Localization of a feeding center in the hypothalamusof the rat, Proc. Soc. Exp. Biol. Med. 77 (1951) 323–324.

[18] E. Aries, D.R. Anderson, R. Fisher, T.A.T. Fray, D. Hemfrey, PCDD/F and dioxin-like PCB emissions from iron ore sintering plants in the UK, Chemosphere 65(2006) 1470–1480.

[19] S. Arpiainen, F. Raffalli-Mathieu, M.A. Lang, O. Pelkonen, J. Hakkola,Regulation of the Cyp2a5 gene involves an aryl hydrocarbon receptor-dependent pathway, Mol. Pharmacol. 67 (2005) 1325–1333.

[20] T. Baba, J. Mimura, N. Nakamura, N. Harada, M. Yamamoto, K. Morohashi,Y. Fujii-Kuriyama, Intrinsic function of the aryl hydrocarbon (dioxin)receptor as a key factor in female reproduction, Mol. Cell. Biol. 25 (2005)10040–10051.

[21] M. Backlund, M. Ingelman-Sundberg, Regulation of aryl hydrocarbonreceptor signal transduction by protein tyrosine kinases, Cell. Signal. 17(2005) 39–48.

[22] S.G. Bacsi, O. Hankinson, Functional characterization of DNA-binding domainsof the subunits of the heterodimeric aryl hydrocarbon receptor compleximputing novel and canonical basic helix–loop–helix protein–DNAinteractions, J. Biol. Chem. 271 (1996) 8843–8850.

[23] S.G. Bacsi, S. Reisz-Porszasz, O. Hankinson, Orientation of the heterodimericaryl hydrocarbon (dioxin) receptor complex on its asymmetric DNArecognition sequence, Mol. Pharmacol. 47 (1995) 432–438.

[24] W.A. Banks, The many lives of leptin, Peptides 25 (2004) 331–338.[25] W.A. Banks, The source of cerebral insulin, Eur. J. Pharmacol. 490 (2004) 5–12.[26] P.W. Beatty, M.A. Holscher, R.A. Neal, Toxicity of 2,3,7,8-tetrachlorodibenzo-

p-dioxin in larval and adult forms of Rana catesbeiana, Bull. Environ. Contam.Toxicol. 16 (1976) 578–581.

[27] S.R. Beedanagari, I. Bebenek, P. Bui, O. Hankinson, Resveratrol inhibits dioxin-induced expression of human CYP1A1 and CYP1B1 by inhibiting recruitment

Page 18: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 469

of the aryl hydrocarbon receptor complex and RNA polymerase II to theregulatory regions of the corresponding genes, Toxicol. Sci. 110 (2009) 61–67.

[28] T.V. Beischlag, J. Luis Morales, B.D. Hollingshead, G.H. Perdew, The arylhydrocarbon receptor complex and the control of gene expression, Crit. Rev.Eukar. Gene Expr. 18 (2008) 207–250.

[29] T.V. Beischlag, R.T. Taylor, D.W. Rose, D. Yoon, Y. Chen, W.H. Lee, M.G.Rosenfeld, O. Hankinson, Recruitment of thyroid hormone receptor/retinoblastoma-interacting protein 230 by the aryl hydrocarbon receptornuclear translocator is required for the transcriptional response to bothdioxin and hypoxia, J. Biol. Chem. 279 (2004) 54620–54628.

[30] T.V. Beischlag, S. Wang, D.W. Rose, J. Torchia, S. Reisz-Porszasz, K.Muhammad, W.E. Nelson, M.R. Probst, M.G. Rosenfeld, O. Hankinson,Recruitment of the NCoA/SRC-1/p160 family of transcriptional coactivatorsby the aryl hydrocarbon receptor/aryl hydrocarbon receptor nucleartranslocator complex, Mol. Cell. Biol. 22 (2002) 4319–4333.

[31] D.R. Bell, S. Clode, M.Q. Fan, A. Fernandes, P.M. Foster, T. Jiang, G. Loizou, A.MacNicoll, B.G. Miller, M. Rose, L. Tran, S. White, Interpretation of studies onthe developmental reproductive toxicology of 2.3.7,8-tetrachlorodibenzo-p-dioxin in male offspring, Food Chem. Toxicol. 48 (2010) 1439–1447.

[32] L.L. Bellinger, L.L. Bernardis, The dorsomedial hypothalamic nucleus and itsrole in ingestive behavior and body weight regulation: lessons learned fromlessoning studies, Physiol. Behav. 76 (2002) 431–442.

[33] K. Berger, M.S. Winzell, J. Mei, C. Erlanson-Albertsson, Enterostatin and itstarget mechanisms during regulation of fat intake, Physiol. Behav. 83 (2004)623–630.

[34] T. Bernshausen, B. Jux, C. Esser, J. Abel, E. Fritsche, Tissue distribution andfunction of the Aryl hydrocarbon receptor repressor (AhRR) in C57BL/6 andAryl hydrocarbon receptor deficient mice, Arch. Toxicol. 80 (2006) 206–211.

[35] K.C. Berridge, M.L. Kringelbach, Affective neuroscience of pleasure: reward inhumans and animals, Psychopharmacology (Berl) 199 (2008) 457–480.

[36] H.R. Berthoud, Multiple neural systems controlling food intake and bodyweight, Neurosci. Biobehav. Rev. 26 (2002) 393–428.

[37] H.R. Berthoud, Vagal, hormonal gut-brain communication from satiation tosatisfaction, Neurogastroenterol. Motil. 20 (Suppl. 1) (2008) 64–72.

[38] L.L. Bestervelt, Y. Cai, D.W. Piper, C.J. Nolan, J.A. Pitt, W.N. Piper, TCDD alterspituitary-adrenal function. I: adrenal responsiveness to exogenous ACTH,Neurotoxicol. Teratol. 15 (1993) 365–367.

[39] S.M. Billiard, M.E. Hahn, D.G. Franks, R.E. Peterson, N.C. Bols, P.V. Hodson,Binding of polycyclic aromatic hydrocarbons (PAHs) to teleost arylhydrocarbon receptors (AHRs), Comp. Biochem. Physiol. B. Biochem. Mol.Biol. 133 (2002) 55–68.

[40] L.S. Birnbaum, The mechanism of dioxin toxicity: relationship to riskassessment, Environ. Health Perspect. 102 (Suppl. 9) (1994) 157–167.

[41] L.S. Birnbaum, M.M. McDonald, P.C. Blair, A.M. Clark, M.W. Harris, Differentialtoxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in C57BL/6J micecongenic at the Ah Locus, Fund. Appl. Toxicol. 15 (1990) 186–200.

[42] L.S. Birnbaum, J. Tuomisto, Non-carcinogenic effects of TCDD in animals, FoodAddit. Contam. 17 (2000) 275–288.

[43] G. Biswas, S. Srinivasan, H.K. Anandatheerthavarada, N.G. Avadhani, Dioxin-mediated tumor progression through activation of mitochondria-to-nucleusstress signaling, Proc. Natl. Acad. Sci. USA 105 (2008) 186–191.

[44] M.A. Bittinger, L.P. Nguyen, C.A. Bradfield, Aspartate aminotransferasegenerates proagonists of the aryl hydrocarbon receptor, Mol. Pharmacol. 64(2003) 550–556.

[45] D.L. Bjerke, T.J. Brown, N.J. MacLusky, R.B. Hochberg, R.E. Peterson, Partialdemasculinization and feminization of sex behavior in male rats by in uteroand lactational exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin is notassociated with alterations in estrogen receptor binding or volumes ofsexually differentiated brain nuclei, Toxicol. Appl. Pharmacol. 127 (1994)258–267.

[46] K.W. Bock, C. Kohle, Ah receptor: dioxin-mediated toxic responses as hints toderegulated physiologic functions, Biochem. Pharmacol. 72 (2006) 393–404.

[47] J.E. Bohonowych, M.S. Denison, Persistent binding of ligands to the arylhydrocarbon receptor, Toxicol. Sci. 98 (2007) 99–109.

[48] P.C. Boutros, K.A. Bielefeld, R. Pohjanvirta, P.A. Harper, Dioxin-dependent anddioxin-independent gene batteries: comparison of liver and kidney in AHR-null mice, Toxicol. Sci. 112 (2009) 245–256.

[49] P.C. Boutros, R. Yan, I.D. Moffat, R. Pohjanvirta, A.B. Okey, Transcriptomicresponses to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in liver: comparisonof rat and mouse, BMC Genom. 9 (2008) 419.

[50] D.R. Boverhof, L.D. Burgoon, C. Tashiro, B. Sharratt, B. Chittim, J.R. Harkema,D.L. Mendrick, T.R. Zacharewski, Comparative toxicogenomic analysis of thehepatotoxic effects of TCDD in Sprague Dawley rats and C57BL/6 mice,Toxicol. Sci. 94 (2006) 398–416.

[51] C. Broberger, Brain regulation of food intake and appetite: molecules andnetworks, J. Int. Med. 258 (2005) 301–327.

[52] A. Brouwer, U.G. Ahlborg, F.X. van Leeuwen, M.M. Feeley, Report of the WHOworking group on the assessment of health risks for human infants fromexposure to PCDDs, PCDFs and PCBs, Chemosphere 37 (1998) 1627–1643.

[53] B. Brunstrom, J. Lund, Differences between chick and turkey embryos insensitivity to 3,30 ,4,40-tetrachloro-biphenyl and in concentration/affinity ofthe hepatic receptor for 2,3,7,8-tetrachlorodibenzo-p-dioxin, Comp. Biochem.Physiol. C. 91 (1988) 507–512.

[54] V. Bucinskaite, T. Tolessa, J. Pedersen, B. Rydqvist, L. Zerihun, J.J. Holst, P.M.Hellstrom, Receptor-mediated activation of gastric vagal afferents by

glucagon-like peptide-1 in the rat, Neurogastroenterol. Motil. 21 (2009)978–985.

[55] D.B. Buckley, C.D. Klaassen, Induction of mouse UDP-glucuronosyltransferasemRNA expression in liver and intestine by activators of aryl-hydrocarbonreceptor, constitutive androstane receptor, pregnane X receptor, peroxisomeproliferator-activated receptor alpha, and nuclear factor erythroid 2-relatedfactor 2, Drug Metab. Dispos. 37 (2009) 847–856.

[56] M. Bunge, U. Lechner, Anaerobic reductive dehalogenation of polychlorinateddioxins, Appl. Microbiol. Biotechnol. 84 (2009) 429–444.

[57] M.K. Bunger, E. Glover, S.M. Moran, J.A. Walisser, G.P. Lahvis, E.L. Hsu, C.A.Bradfield, Abnormal liver development and resistance to 2,3,7,8-tetrachlorodibenzo-p-dioxin toxicity in mice carrying a mutation in theDNA-binding domain of the aryl hydrocarbon receptor, Toxicol. Sci. 106(2008) 83–92.

[58] M.K. Bunger, S.M. Moran, E. Glover, T.L. Thomae, G.P. Lahvis, B.C. Lin, C.A.Bradfield, Resistance to 2,3,7,8-tetrachlorodibenzo-p-dioxin toxicity andabnormal liver development in mice carrying a mutation in the nuclearlocalization sequence of the aryl hydrocarbon receptor, J. Biol. Chem. 278(2003) 17767–17774.

[59] R.A. Butler, M.L. Kelley, W.H. Powell, M.E. Hahn, R.J. Van Beneden, An arylhydrocarbon receptor (AHR) homologue from the soft-shell clam, Myaarenaria: evidence that invertebrate AHR homologues lack 2,3,7,8-tetrachlorodibenzo-p-dioxin and beta-naphthoflavone binding, Gene 278(2001) 223–234.

[60] M. Cabanac, Regulation and the ponderostat, Int. J. Obes. Relat. Metab.Disorders 25 (Suppl. 5) (2001) S7–12.

[61] L.A. Campfield, F.J. Smith, Blood glucose dynamics and control of mealinitiation: a pattern detection and recognition theory, Physiol. Rev. 83 (2003)25–58.

[62] D.B. Carlson, G.H. Perdew, A dynamic role for the Ah receptor in cellsignaling? Insights from a diverse group of Ah receptor interacting proteins, J.Biochem. Mol. Toxicol. 16 (2002) 317–325.

[63] E.A. Carlson, C. McCulloch, A. Koganti, S.B. Goodwin, T.R. Sutter, J.B. Silkworth,Divergent transcriptomic responses to aryl hydrocarbon receptor agonistsbetween rat and human primary hepatocytes, Toxicol. Sci. 112 (2009) 257–272.

[64] A. Carlsson, M. Lindquist, Dependence of 5-HT and catecholamine synthesison precursor amino-acid levels in rat brain, Naunyn Shmiedelberg’s Arch.Pharmacol*********** 303 (1978) 157–164.

[65] W.K. Chan, G. Yao, Y.Z. Gu, C.A. Bradfield, Cross-talk between the arylhydrocarbon receptor and hypoxia inducible factor signaling pathways.Demonstration of competition and compensation, J. Biol. Chem. 274 (1999)12115–12123.

[66] S.F. Chang, Y.Y. Sun, L.Y. Yang, S.Y. Hu, S.Y. Tsai, W.S. Lee, Y.H. Lee, Bcl-2 genefamily expression in the brain of rat offspring after gestational and lactationaldioxin exposure, Ann. NY Acad. Sci. 1042 (2005) 471–480.

[67] D.E. Chapman, C.M. Schiller, Dose-related effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in C57BL/6J and DBA/2J mice, Toxicol.Appl. Pharmacol. 78 (1985) 147–157.

[68] O.B. Chaudhri, V. Salem, K.G. Murphy, S.R. Bloom, Gastrointestinal satietysignals, Annu. Rev. Physiol. 70 (2008) 239–255.

[69] M.J. Chee, W.F. Colmers, Y eat?, Nutrition 24 (2008) 869–877[70] S.B. Cheng, S. Kuchiiwa, A. Kawachi, H.Z. Gao, A. Gohshi, T. Kozako, T.

Kuchiiwa, S. Nakagawa, Up-regulation of methionine–enkephalin-likeimmunoreactivity by 2,3,7,8-tetrachlorodibenzo-p-dioxin treatment inthe forebrain of the Long-Evans rat, J. Chem. Neuroanat. 25 (2003) 73–82.

[71] S.B. Cheng, S. Kuchiiwa, I. Nagatomo, Y. Akasaki, M. Uchida, M. Tominaga, W.Hashiguchi, T. Kuchiiwa, S. Nakagawa, 2,3,7,8-Tetrachlorodibenzo-p-dioxintreatment induces c-Fos expression in the forebrain of the Long-Evans rat,Brain Res. 931 (2002) 176–180.

[72] S.B. Cheng, S. Kuchiiwa, X.Q. Ren, H.Z. Gao, T. Kuchiiwa, S. Nakagawa, Dioxinexposure down-regulates nitric oxide synthase and NADPH-diaphoraseactivities in the hypothalamus of Long-Evans rat, Neurosci. Lett. 345 (2003)5–8.

[73] H. Cheon, Y.S. Woo, J.Y. Lee, H.S. Kim, H.J. Kim, S. Cho, N.H. Won, J. Sohn,Signaling pathway for 2,3,7,8-tetrachlorodibenzo-p-dioxin-induced TNF-alpha production in differentiated THP-1 human macrophages, Exp. Mol.Med. 39 (2007) 524–534.

[74] S.J. Cho, J.S. Jung, I. Jin, Y.W. Jung, B.H. Ko, K.S. Nam, I.K. Park, I.S. Moon, Effectsof 2,3,7,8-tetrachlorodibenzo-p-dioxin on the expression of synaptic proteinsin dissociated rat cortical cells, Mol. Cells 14 (2002) 238–244.

[75] B.J. Christian, S.L. Inhorn, R.E. Peterson, Relationship of the wasting syndrometo lethality in rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin, Toxicol.Appl. Pharmacol. 82 (1986) 239–255.

[76] C.C. Coimbra, R.H. Migliorini, Insulin-sensitive glucoreceptors in rat preopticarea that regulate FFA mobilization, Am. J. Physiol. 251 (1986) E703–E706.

[77] A. Colles, G. Koppen, V. Hanot, V. Nelen, M.C. Dewolf, E. Noel, R. Malisch, A.Kotz, K. Kypke, P. Biot, C. Vinkx, G. Schoeters, Fourth WHO-coordinatedsurvey of human milk for persistent organic pollutants (POPs): Belgianresults, Chemosphere 73 (2008) 907–914.

[78] D. Consonni, A.C. Pesatori, C. Zocchetti, R. Sindaco, L.C. D’Oro, M. Rubagotti,P.A. Bertazzi, Mortality in a population exposed to dioxin after the Seveso,Italy, accident in 1976: 25 years of follow-up, Am. J. Epidemiol. 167 (2008)847–858.

Page 19: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

470 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

[79] Anonymous, Consultation on assessment of the health risk of dioxins; re-evaluation of the tolerable daily intake (TDI): executive summary, Food Addit.Contam. 17 (2000) 223–240.

[80] P. Coumailleau, L. Poellinger, J.A. Gustafsson, M.L. Whitelaw, Definition of aminimal domain of the dioxin receptor that is associated with Hsp90 andmaintains wild type ligand binding affinity and specificity, J. Biol. Chem. 270(1995) 25291–25300.

[81] G. Curzon, Serotonin and appetite, Ann. NY Acad. Sci. 600 (1990) 521–530(discussion 530-1).

[82] P. Dabir, T.E. Marinic, I. Krukovets, O.I. Stenina, Aryl hydrocarbon receptor isactivated by glucose and regulates the thrombospondin-1 gene promoter inendothelial cells, Circ. Res. 102 (2008) 1558–1565.

[83] Y. Date, M. Kojima, H. Hosoda, A. Sawaguchi, M.S. Mondal, T. Suganuma, S.Matsukura, K. Kangawa, M. Nakazato, Ghrelin, a novel growth hormone-releasing acylated peptide, is synthesized in a distinct endocrine cell type inthe gastrointestinal tracts of rats and humans, Endocrinology 141 (2000)4255–4261.

[84] S. Dauchy, F. Dutheil, R.J. Weaver, F. Chassoux, C. Daumas-Duport, P.O.Couraud, J.M. Scherrmann, I. De Waziers, X. Decleves, ABC transporterscytochromes P450 and their main transcription factors: expression at thehuman blood–brain barrier, J. Neurochem. 107 (2008) 1518–1528.

[85] M. Daujat, B. Peryt, P. Lesca, G. Fourtanier, J. Domergue, P. Maurel,Omeprazole, an inducer of human CYP1A1 and 1A2, is not a ligand for theAh receptor, Biochem. Biophys. Res. Commun. 188 (1992) 820–825.

[86] N.A. Davarinos, R.S. Pollenz, Aryl hydrocarbon receptor imported into thenucleus following ligand binding is rapidly degraded via the cytoplasmicproteasome following nuclear export, J. Biol. Chem. 274 (1999) 28708–28715.

[87] S.K. de Oliveira, M. Hoffmeister, S. Gambaryan, W. Muller-Esterl, J.A.Guimaraes, A.P. Smolenski, Phosphodiesterase 2A forms a complex with theco-chaperone XAP2 and regulates nuclear translocation of the arylhydrocarbon receptor, J. Biol. Chem. 282 (2007) 13656–13663.

[88] S.K. de Oliveira, A. Smolenski, Phosphodiesterases link the aryl hydrocarbonreceptor complex to cyclic nucleotide signaling, Biochem. Pharmacol. 77(2009) 723–733.

[89] M. Denis, S. Cuthill, A.C. Wikstrom, L. Poellinger, J.A. Gustafsson, Associationof the dioxin receptor with the Mr 90,000 heat shock protein: a structuralkinship with the glucocorticoid receptor, Biochem. Biophys. Res. Commun.155 (1988) 801–807.

[90] M.S. Denison, J.M. Fisher, J.P. Whitlock Jr., The DNA recognition site for thedioxin–Ah receptor complex. Nucleotide sequence and functional analysis, J.Biol. Chem. 263 (1988) 17221–17224.

[91] M.S. Denison, S.R. Nagy, Activation of the aryl hydrocarbon receptor bystructurally diverse exogenous and endogenous chemicals, Annu. Rev.Pharmacol. Toxicol. 43 (2003) 309–334.

[92] M.O. Dietrich, T.L. Horvath, Feeding signals and brain circuitry, Eur. J.Neurosci. 30 (2009) 1688–1696.

[93] K.M. Dolwick, J.V. Schmidt, L.A. Carver, H.I. Swanson, C.A. Bradfield, Cloningand expression of a human Ah receptor cDNA, Mol. Pharmacol. 44 (1993)911–917.

[94] B. Dong, F. Matsumura, Roles of cytosolic phospholipase A2 and Src kinase inthe early action of 2,3,7,8-tetrachlorodibenzo-p-dioxin through anongenomic pathway in MCF10A cells, Mol. Pharmacol. 74 (2008) 255–263.

[95] B. Dong, F. Matsumura, The conversion of rapid TCCD nongenomic signals topersistent inflammatory effects via select protein kinases in MCF10A cells,Mol. Endocrinol. 23 (2009) 549–558.

[96] B. Dong, N. Nishimura, C.F. Vogel, C. Tohyama, F. Matsumura, TCDD-inducedcyclooxygenase-2 expression is mediated by the nongenomic pathway inmouse MMDD1 macula densa cells and kidneys, Biochem. Pharmacol. 79(2010) 487–497.

[97] T.M. Dovey, P.A. Staples, E.L. Gibson, J.C. Halford, Food neophobia and ‘picky/fussy’ eating in children: a review, Appetite 50 (2008) 181–193.

[98] Y.P. Dragan, D. Schrenk, Animal studies addressing the carcinogenicity ofTCDD (or related compounds) with an emphasis on tumour promotion, FoodAddit. Contam. 17 (2000) 289–302.

[99] Y. Dumont, E. Moyse, A. Fournier, R. Quirion, Distribution of peripherallyinjected peptide YY ([125I] PYY (3–36)) and pancreatic polypeptide ([125I]hPP) in the CNS: enrichment in the area postrema, J. Mol. Neurosci. 33 (2007)294–304.

[100] D.M. Duncan, E.A. Burgess, I. Duncan, Control of distal antennal identity andtarsal development in Drosophila by spineless-aristapedia, a homolog of themammalian dioxin receptor, Genes Dev. 12 (1998) 1290–1303.

[101] T.J. Dunn, R. Lindahl, H.C. Pitot, Differential gene expression in response to2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). Noncoordinate regulation of aTCDD-induced aldehyde dehydrogenase and cytochrome P-450c in the rat, J.Biol. Chem. 263 (1988) 10878–10886.

[102] A.A. Dunn-Meynell, N.M. Sanders, D. Compton, T.C. Becker, J. Eiki, B.B. Zhang,B.E. Levin, Relationship among brain and blood glucose levels andspontaneous and glucoprivic feeding, J. Neurosci. 29 (2009) 7015–7022.

[103] F. Dutheil, S. Dauchy, M. Diry, V. Sazdovitch, O. Cloarec, L. Mellottee, I. Bieche,M. Ingelman-Sundberg, J.P. Flinois, I. de Waziers, P. Beaune, X. Decleves, C.Duyckaerts, M.A. Loriot, Xenobiotic-metabolizing enzymes and transportersin the normal human brain: regional and cellular mapping as a basis forputative roles in cerebral function, Drug Metab. Dispos. 37 (2009) 1528–1538.

[104] P.H. Dyke, C. Foan, M. Wenborn, P.J. Coleman, A review of dioxin releases toland and water in the UK, Sci. Total Environ. 207 (1997) 119–131.

[105] N. Dzeletovic, J. McGuire, M. Daujat, J. Tholander, M. Ema, Y. Fujii-Kuriyama, J.Bergman, P. Maurel, L. Poellinger, Regulation of dioxin receptor function byomeprazole, J. Biol. Chem. 272 (1997) 12705–12713.

[106] F.J. Ebling, P. Barrett, The regulation of seasonal changes in food intake andbody weight, J. Neuroendocrinol. 20 (2008) 827–833.

[107] R.H. Elbekai, A.O. El-Kadi, Transcriptional activation and posttranscriptionalmodification of Cyp1a1 by arsenite, cadmium, and chromium, Toxicol. Lett.172 (2007) 106–119.

[108] C.J. Elferink, N.L. Ge, A. Levine, Maximal aryl hydrocarbon receptor activitydepends on an interaction with the retinoblastoma protein, Mol. Pharmacol.59 (2001) 664–673.

[109] R.B. Emmons, D. Duncan, P.A. Estes, P. Kiefel, J.T. Mosher, M. Sonnenfeld, M.P.Ward, I. Duncan, S.T. Crews, The spineless-aristapedia and tango bHLH-PASproteins interact to control antennal and tarsal development in Drosophila,Development 126 (1999) 3937–3945.

[110] E. Enan, F. El-Sabeawy, M. Scott, J. Overstreet, B. Lasley, Alterations in thegrowth factor signal transduction pathways and modulators of the cell cyclein endocervical cells from macaques exposed to TCDD, Toxicol. Appl.Pharmacol. 151 (1998) 283–293.

[111] C. Esser, The immune phenotype of AhR null mouse mutants: not a simplemirror of xenobiotic receptor over-activation, Biochem. Pharmacol. 77 (2009)597–607.

[112] S.A. Farr, W.A. Banks, V.B. Kumar, J.E. Morley, Orexin-A-induced feeding isdependent on nitric oxide, Peptides 26 (2005) 759–765.

[113] M. Feeley, A. Brouwer, Health risks to infants from exposure to PCBs, PCDDsand PCDFs, Food Addit. Contam. 17 (2000) 325–333.

[114] P. Fernandez-Salguero, T. Pineau, D.M. Hilbert, T. McPhail, S.S. Lee, S. Kimura,D.W. Nebert, S. Rudikoff, J.M. Ward, F.J. Gonzalez, Immune systemimpairment and hepatic fibrosis in mice lacking the dioxin-binding Ahreceptor, Science 268 (1995) 722–726.

[115] P.M. Fernandez-Salguero, D.M. Hilbert, S. Rudikoff, J.M. Ward, F.J. Gonzalez,Aryl-hydrocarbon receptor-deficient mice are resistant to 2,3,7,8-tetrachlorodibenzo-p-dioxin-induced toxicity, Toxicol. Appl. Pharmacol. 140(1996) 173–179.

[116] P.M. Fernandez-Salguero, J.M. Ward, J.P. Sundberg, F.J. Gonzalez, Lesionsof aryl-hydrocarbon receptor-deficient mice, Vet. Pathol. 34 (1997) 605–614.

[117] S.O. Fetissov, P. Huang, Q. Zhang, J. Mimura, Y. Fujii-Kuriyama, A. Rannug, T.Hokfelt, S. Ceccatelli, Expression of hypothalamic neuropeptides after acuteTCDD treatment and distribution of Ah receptor repressor, Regul. Peptides119 (2004) 113–124.

[118] D.P. Figlewicz, S.C. Benoit, Insulin, leptin, and food reward: update 2008, Am.J. Physiol. Regul. Integr. Comp. Physiol. 296 (2009) R9–R19.

[119] Y. Figueroa-Guzmán, S. Reilly, NMDA receptors in the basolateral amygdalaand gustatory neophobia, Brain Res. 1210 (2008) 200–203.

[120] C.R. Filbrandt, Z. Wu, B. Zlokovic, L. Opanashuk, T.A. Gasiewicz, Presence andfunctional activity of the aryl hydrocarbon receptor in isolated murinecerebral vascular endothelial cells and astrocytes, Neurotoxicology 25 (2004)605–616.

[121] C.A. Flaveny, I.A. Murray, C.R. Chiaro, G.H. Perdew, Ligand selectivity and generegulation by the human aryl hydrocarbon receptor in transgenic mice, Mol.Pharmacol. 75 (2009) 1412–1420.

[122] C.A. Flaveny, I.A. Murray, G.H. Perdew, Differential gene regulation by thehuman and mouse aryl hydrocarbon receptor, Toxicol. Sci. 114 (2010) 217–225.

[123] N. Fletcher, D. Wahlstrom, R. Lundberg, C.B. Nilsson, K.C. Nilsson, K. Stockling,H. Hellmold, H. Hakansson, 2,3,7,8-Tetrachlorodibenzo-p-dioxin (TCDD)alters the mRNA expression of critical genes associated with cholesterolmetabolism, bile acid biosynthesis, and bile transport in rat liver: amicroarray study, Toxicol. Appl. Pharmacol. 207 (2005) 1–24.

[124] M.A. Franc, I.D. Moffat, P.C. Boutros, J.T. Tuomisto, J. Tuomisto, R.Pohjanvirta, A.B. Okey, Patterns of dioxin-altered mRNA expression inlivers of dioxin-sensitive versus dioxin-resistant rats, Arch. Toxicol. 82(2008) 809–830.

[125] K.W. Fried, G.L. Guo, N. Esterly, B. Kong, K.K. Rozman, 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) reverses hyperglycemia in a type IIdiabetes mellitus rat model by a mechanism unrelated to PPARgamma, DrugChem. Toxicol. (2010).

[126] A. Fujisawa-Sehara, K. Sogawa, M. Yamane, Y. Fujii-Kuriyama,Characterization of xenobiotic responsive elements upstream from thedrug-metabolizing cytochrome P-450c gene: a similarity to glucocorticoidregulatory elements, Nucleic Acids Res. 15 (1987) 4179–4191.

[127] B.N. Fukunaga, M.R. Probst, S. Reisz-Porszasz, O. Hankinson, Identification offunctional domains of the aryl hydrocarbon receptor, J. Biol. Chem. 270(1995) 29270–29278.

[128] M. Fukushima, K. Tokunaga, J. Lupien, J.W. Kemnitz, G.A. Bray, Dynamic andstatic phases of obesity following lesions in PVN and VMH, Am. J. Physiol.Regul. Integr. Comp. Physiol. 253 (1987) R523–R529.

[129] S.G. Furness, F. Whelan, The pleiotropy of dioxin toxicity – xenobioticmisappropriation of the aryl hydrocarbon receptor’s alternative physiologicalroles, Pharmacol. Ther. (2009).

[130] P. Furst, Dioxins, polychlorinated biphenyls and other organohalogencompounds in human milk. Levels, correlations, trends and exposurethrough breastfeeding, Mol. Nutr. Food Res. 50 (2006) 922–933.

[131] K.W. Gaido, S.C. Maness, L.S. Leonard, W.F. Greenlee, 2,3,7,8-Tetrachlorodibenzo-p-dioxin-dependent regulation of transforming growth

Page 20: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 471

factors-alpha and -beta 2 expression in a human keratinocyte cell lineinvolves both transcriptional and post-transcriptional control, J. Biol. Chem.267 (1992) 24591–24595.

[132] Q. Gao, T.L. Horvath, Neuronal control of energy homeostasis, FEBS Lett. 582(2008) 132–141.

[133] R.W. Garrett, T.A. Gasiewicz, The aryl hydrocarbon receptor agonist 2,3,7,8-tetrachlorodibenzo-p-dioxin alters the circadian rhythms, quiescence, andexpression of clock genes in murine hematopoietic stem and progenitor cells,Mol. Pharmacol. 69 (2006) 2076–2083.

[134] T.A. Gasiewicz, C.J. Elferink, E.C. Henry, Characterization of multiple forms ofthe Ah receptor: recognition of a dioxin-responsive enhancer involvesheteromer formation, Biochemistry 30 (1991) 2909–2916.

[135] T.A. Gasiewicz, M.A. Holscher, R.A. Neal, The effect of total parenteralnutrition on the toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin in the rat,Toxicol. Appl. Pharmacol. 54 (1980) 469–488.

[136] T.A. Gasiewicz, A.S. Kende, G. Rucci, B. Whitney, J.J. Willey, Analysis ofstructural requirements for Ah receptor antagonist activity: ellipticines,flavones, and related compounds, Biochem. Pharmacol. 52 (1996) 1787–1803.

[137] T.A. Gasiewicz, R.A. Neal, 2,3,7,8-Tetrachlorodibenzo-p-dioxin tissuedistribution, excretion, and effects on clinical chemical parameters inguinea pigs, Toxicol. Appl. Pharmacol. 51 (1979) 329–339.

[138] T.A. Gasiewicz, G. Rucci, E.C. Henry, R.B. Baggs, Changes in hamster hepaticcytochrome P-450, ethoxycoumarin O-deethylase, and reduced NAD(P):menadione oxidoreductase following treatment with 2,3,7,8-tetrachlorodibenzo-p-dioxin. Partial dissociation of temporal and dose-response relationships from elicited toxicity, Biochem. Pharmacol. 35(1986) 2737–2742.

[139] F.S. Gaskin, S.A. Farr, W.A. Banks, V.B. Kumar, J.E. Morley, Ghrelin-inducedfeeding is dependent on nitric oxide, Peptides 24 (2003) 913–918.

[140] N.L. Ge, C.J. Elferink, A direct interaction between the aryl hydrocarbonreceptor and retinoblastoma protein. Linking dioxin signaling to the cellcycle, J. Biol. Chem. 273 (1998) 22708–22713.

[141] B. Gereben, A.M. Zavacki, S. Ribich, B.W. Kim, S.A. Huang, W.S. Simonides, A.Zeold, A.C. Bianco, Cellular and molecular basis of deiodinase-regulatedthyroid hormone signaling, Endocr. Rev. 29 (2008) 898–938.

[142] K. Gerozissis, Brain insulin, energy and glucose homeostasis; genes,environment and metabolic pathologies, Eur. J. Pharmacol. 585 (2008) 38–49.

[143] A. Geusau, K. Abraham, K. Geissler, M.O. Sator, G. Stingl, E. Tschachler, Severe2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) intoxication: clinical andlaboratory effects, Environ. Health Perspect. 109 (2001) 865–869.

[144] A. Geusau, S. Schmaldienst, K. Derfler, O. Papke, K. Abraham, Severe 2,3,7,8-tetrachlorodibenzo- p-dioxin (TCDD) intoxication: kinetics and trials toenhance elimination in two patients, Arch. Toxicol. 76 (2002) 316–325.

[145] J.M. Gohlke, P.S. Stockton, S. Sieber, J. Foley, C.J. Portier, AhR-mediated geneexpression in the developing mouse telencephalon, Reprod. Toxicol. 28(2009) 321–328.

[146] J.A. Goldstein, P. Linko, Differential induction of two 2,3,7,8-tetrachlorodibenzo-p-dioxin-inducible forms of cytochrome P-450 inextrahepatic versus hepatic tissues, Mol. Pharmacol. 25 (1984) 185–191.

[147] F.J. Gonzalez, P. Fernandez-Salguero, The aryl hydrocarbon receptor:studies using the AHR-null mice, Drug Metab. Dispos. 26 (1998) 1194–1198.

[148] J.R. Gorski, G. Muzi, L.W. Weber, D.W. Pereira, R.J. Arceo, M.J. Iatropoulos, K.Rozman, Some endocrine and morphological aspects of the acute toxicity of2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD), Toxicol. Pathol. 16 (1988) 313–320.

[149] J.R. Gorski, K. Rozman, Dose-response and time course of hypothyroxinemiaand hypoinsulinemia and characterization of insulin hypersensitivity in2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)-treated rats, Toxicology 44(1987) 297–307.

[150] S. Gradelet, P. Astorg, T. Pineau, M.C. Canivenc, M.H. Siess, J. Leclerc, P. Lesca,Ah receptor-dependent CYP1A induction by two carotenoids, canthaxanthinand beta-apo-80-carotenal, with no affinity for the TCDD binding site,Biochem. Pharmacol. 54 (1997) 307–315.

[151] L.E. Gray, J.S. Ostby, W.R. Kelce, A dose-response analysis of the reproductiveeffects of a single gestational dose of 2,3,7,8-tetrachlorodibenzo-p-dioxin inmale Long Evans Hooded rat offspring, Toxicol. Appl. Pharmacol. 146 (1997)11–20.

[152] H.J. Grill, Leptin and the systems neuroscience of meal size control, Front.Neuroendocrinol. 31 (2010) 61–78.

[153] H.J. Grill, M.W. Schwartz, J.M. Kaplan, J.S. Foxhall, J. Breininger, D.G. Baskin,Evidence that the caudal brainstem is a target for the inhibitory effect ofleptin on food intake, Endocrinology 143 (2002) 239–246.

[154] J.L. Gross, R.H. Migliorini, Further evidence for a central regulation of freefatty acid mobilization in the rat, Am. J. Physiol. Endocrinol. Metab. 232(1977) E165–E171.

[155] A. Guyon, G. Conductier, C. Rovere, A. Enfissi, J.L. Nahon, Melanin-concentrating hormone producing neurons: activities and modulations,Peptides 30 (2009) 2031–2039.

[156] T. Haarmann-Stemmann, H. Bothe, J. Abel, Growth factors, cytokines andtheir receptors as downstream targets of arylhydrocarbon receptor (AhR)signaling pathways, Biochem. Pharmacol. 77 (2009) 508–520.

[157] M.E. Hahn, Dioxin toxicology and the aryl hydrocarbon receptor: insightsfrom fish and other non-traditional models, Mar. Biotechnol. 3 (2001) S224–S238.

[158] M.E. Hahn, Aryl hydrocarbon receptors: diversity and evolution, Chem. Biol.Interact. 141 (2002) 131–160.

[159] M.E. Hahn, L.L. Allan, D.H. Sherr, Regulation of constitutive and inducible AHRsignaling: complex interactions involving the AHR repressor, Biochem.Pharmacol. 77 (2009) 485–497.

[160] M.E. Hahn, S.I. Karchner, B.R. Evans, D.G. Franks, R.R. Merson, J.M. Lapseritis,Unexpected diversity of aryl hydrocarbon receptors in non-mammalianvertebrates: insights from comparative genomics, J. Exp. Zool. A. Comp. Exp.Biol. 305 (2006) 693–706.

[161] M.E. Hahn, A. Poland, E. Glover, J.J. Stegeman, Photoaffinity labeling of the Ahreceptor: phylogenetic survey of diverse vertebrate and invertebrate species,Arch. Biochem. Biophys. 310 (1994) 218–228.

[162] H. Hakk, J.J. Diliberto, L.S. Birnbaum, The effect of dose on 2,3,7,8-TCDD tissuedistribution, metabolism and elimination in CYP1A2 (–/–) knockout andC57BL/6N parental strains of mice, Toxicol. Appl. Pharmacol. 241 (2009) 119–126.

[163] J.L. Halaas, C. Boozer, J. Blair-West, N. Fidahusein, D.A. Denton, J.M. Friedman,Physiological response to long-term peripheral and central leptin infusion inlean and obese mice, Proc. Natl. Acad. Sci. USA 94 (1997) 8878–8883.

[164] W.H. Hanneman, M.E. Legare, R. Barhoumi, R.C. Burghardt, S. Safe, E. Tiffany-Castiglioni, Stimulation of calcium uptake in cultured rat hippocampalneurons by 2,3,7,8-tetrachlorodibenzo-p-dioxin, Toxicology 112 (1996) 19–28.

[165] E.B. Harstad, C.A. Guite, T.L. Thomae, C.A. Bradfield, Liver deformation in Ahr-null mice. evidence for aberrant hepatic perfusion in early development, Mol.Pharmacol. 69 (2006) 1534–1541.

[166] L.E. Hays, C.D. Carpenter, S.L. Petersen, Evidence that GABAergic neurons inthe preoptic area of the rat brain are targets of 2,3,7,8-tetrachlorodibenzo-p-dioxin during development, Environ. Health Perspect. 110 (Suppl. 3) (2002)369–376.

[167] J.A. Head, M.E. Hahn, S.W. Kennedy, Key amino acids in the aryl hydrocarbonreceptor predict dioxin sensitivity in avian species, Environ. Sci. Technol. 42(2008) 7535–7541.

[168] S. Heath-Pagliuso, W.J. Rogers, K. Tullis, S.D. Seidel, P.H. Cenijn, A. Brouwer,M.S. Denison, Activation of the Ah receptor by tryptophan and tryptophanmetabolites, Biochemistry 37 (1998) 11508–11515.

[169] E.C. Henry, J.C. Bemis, O. Henry, A.S. Kende, T.A. Gasiewicz, A potentialendogenous ligand for the aryl hydrocarbon receptor has potent agonistactivity in vitro and in vivo, Arch. Biochem. Biophys. 450 (2006) 67–77.

[170] E.C. Henry, T.A. Gasiewicz, Changes in thyroid hormones and thyroxineglucuronidation in hamsters compared with rats following treatment with2,3,7,8-tetrachlorodibenzo-p-dioxin, Toxicol. Appl. Pharmacol. 89 (1987)165–174.

[171] E.C. Henry, T.A. Gasiewicz, Molecular determinants of species-specific agonistand antagonist activity of a substituted flavone towards the aryl hydrocarbonreceptor, Arch. Biochem. Biophys. 472 (2008) 77–88.

[172] I. Hernandez-Ochoa, B.N. Karman, J.A. Flaws, The role of the aryl hydrocarbonreceptor in the female reproductive system, Biochem. Pharmacol. 77 (2009)547–559.

[173] G.R. Hervey, Regulation of energy balance, Nature 222 (1969) 629–631.[174] E.V. Hestermann, M. Brown, Agonist and chemopreventative ligands induce

differential transcriptional cofactor recruitment by aryl hydrocarbonreceptor, Mol. Cell. Biol. 23 (2003) 7920–7925.

[175] J.W. Hill, K.W. Williams, C. Ye, J. Luo, N. Balthasar, R. Coppari, M.A. Cowley,L.C. Cantley, B.B. Lowell, J.K. Elmquist, Acute effects of leptin require PI3Ksignaling in hypothalamic proopiomelanocortin neurons in mice, J. Clin.Invest. 118 (2008) 1796–1805.

[176] Y. Hirabayashi, T. Inoue, Aryl hydrocarbon receptor biology and xenobioticresponses in hematopoietic progenitor cells, Biochem. Pharmacol. 77 (2009)521–535.

[177] T. Hofer, R. Pohjanvirta, P. Spielmann, M. Viluksela, D.P. Buchmann, R.H.Wenger, M. Gassmann, Simultaneous exposure of rats to dioxin and carbonmonoxide reduces the xenobiotic but not the hypoxic response, Biol. Chem.385 (2004) 291–294.

[178] D.B. Hood, L. Woods, L. Brown, S. Johnson, F.F. Ebner, Gestational 2,3,7,8-tetrachlorodibenzo-p-dioxin exposure effects on sensory cortex function,Neurotoxicology 27 (2006) 1032–1042.

[179] G.E. Hook, J.K. Haseman, G.W. Lucier, Induction and suppression of hepaticand extrahepatic microsomal foreign-compound-metabolizing enzymesystems by 2,3,7,8-tetrachlorodibenzo-p-dioxin, Chem. Biol. Interact. 10(1975) 199–214.

[180] C.J. Huang Lu, R.B. Baggs, D. Redmond, E.C. Henry, A. Schecter, T.A. Gasiewicz,Toxicity and evidence for metabolic alterations in 2,3,7,8-tetrachlorodibenzo-p-dioxin-treated guinea pigs fed by total parenteral nutrition, Toxicol. Appl.Pharmacol. 84 (1986) 439–453.

[181] P. Huang, S. Ceccatelli, P. Hoegberg, T.J. Sten Shi, H. Hakansson, A. Rannug,TCDD-induced expression of Ah receptor responsive genes in the pituitaryand brain of cellular retinol-binding protein (CRBP-I) knockout mice, Toxicol.Appl. Pharmacol. 192 (2003) 262–274.

[182] P. Huang, A. Rannug, E. Ahlbom, H. Hakansson, S. Ceccatelli, Effect of 2,3,7,8-tetrachlorodibenzo-p-dioxin on the expression of cytochrome P450 1A1, thearyl hydrocarbon receptor, and the aryl hydrocarbon receptor nucleartranslocator in rat brain and pituitary, Toxicol. Appl. Pharmacol. 169 (2000)159–167.

[183] X. Huang, J.A. Powell-Coffman, Y. Jin, The AHR-1 aryl hydrocarbon receptorand its co-factor the AHA-1 aryl hydrocarbon receptor nuclear translocator

Page 21: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

472 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

specify GABAergic neuron cell fate in C. elegans, Development 131 (2004)819–828.

[184] R.E. Humphries, R.M. Sibly, A.P. Meehan, Cereal aversion in behaviourallyresistant house mice in Birmingham, UK, Appl. Anim. Behav. Sci. 66 (2000)323–333.

[185] C. Hundeiker, T. Pineau, G. Cassar, R.A. Betensky, E. Gleichmann, C. Esser,Thymocyte development in Ah-receptor-deficient mice is refractory to TCDD-inducible changes, Int. J. Immunopharmacol. 21 (1999) 841–859.

[186] Anonymous, IARC Working Group on the Evaluation of Carcinogenic Risks toHumans: Polychlorinated Dibenzo-Para-Dioxins and PolychlorinatedDibenzofurans. Lyon, France, 4–11 February 1997. IARC Monogr. Eval.Carcinog. Risks Hum. 69 (1997) 1–631.

[187] T. Ikuta, H. Eguchi, T. Tachibana, Y. Yoneda, K. Kawajiri, Nuclear localizationand export signals of the human aryl hydrocarbon receptor, J. Biol. Chem. 273(1998) 2895–2904.

[188] S. Jain, K.M. Dolwick, J.V. Schmidt, C.A. Bradfield, Potent transactivationdomains of the Ah receptor and the Ah receptor nuclear translocator map totheir carboxyl termini, J. Biol. Chem. 269 (1994) 31518–31524.

[189] S. Jain, E. Maltepe, M.M. Lu, C. Simon, C.A. Bradfield, Expression of ARNT,ARNT2, HIF1 alpha, HIF2 alpha and Ah receptor mRNAs in the developingmouse, Mech. Dev. 73 (1998) 117–123.

[190] M.H. Jin, C.H. Hong, H.Y. Lee, H.J. Kang, S.W. Han, Enhanced TGF-beta1 isinvolved in 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) induced oxidativestress in C57BL/6 mouse testis, Toxicol. Lett. 178 (2008) 202–209.

[191] A.W. Johnson, R. Canter, M. Gallagher, P.C. Holland, Assessing the role of thegrowth hormone secretagogue receptor in motivational learning and foodintake, Behav. Neurosci. 123 (2009) 1058–1065.

[192] L.C. Jones, J.P. Whitlock Jr., Dioxin-inducible transactivation in achromosomal setting. Analysis of the acidic domain of the Ah receptor, J.Biol. Chem. 276 (2001) 25037–25042.

[193] E.A. Jorgensen, U. Knigge, J. Warberg, A. Kjaer, Histamine and the regulationof body weight, Neuroendocrinology 86 (2007) 210–214.

[194] H. Kageyama, F. Takenoya, K. Shiba, S. Shioda, Neuronal circuits involvingghrelin in the hypothalamus-mediated regulation of feeding, Neuropeptides(2009).

[195] T. Kainu, J.A. Gustafsson, M. Pelto-Huikko, The dioxin receptor and its nucleartranslocator (Arnt) in the rat brain, Neuroreport 6 (1995) 2557–2560.

[196] M. Kakeyama, H. Sone, C. Tohyama, Changes in expression of NMDA receptorsubunit mRNA by perinatal exposure to dioxin, Neuroreport 12 (2001) 4009–4012.

[197] M. Kakeyama, C. Tohyama, Developmental neurotoxicity of dioxin and itsrelated compounds, Ind. Health 41 (2003) 215–230.

[198] S.I. Karchner, D.G. Franks, S.W. Kennedy, M.E. Hahn, The molecular basis fordifferential dioxin sensitivity in birds: role of the aryl hydrocarbon receptor,Proc. Natl. Acad. Sci. USA 103 (2006) 6252–6257.

[199] S.I. Karchner, M.J. Jenny, A.M. Tarrant, B.R. Evans, H.J. Kang, I. Bae, D.H. Sherr,M.E. Hahn, The active form of human aryl hydrocarbon receptor (AHR)repressor lacks exon 8, and its Pro 185 and Ala 185 variants repress both AHRand hypoxia-inducible factor, Mol. Cell. Biol. 29 (2009) 3465–3477.

[200] E. Karra, R.L. Batterham, The role of gut hormones in the regulation of bodyweight and energy homeostasis, Mol. Cell. Endocrinol. 316 (2010) 120–128.

[201] H. Kattainen, J. Tuukkanen, U. Simanainen, J.T. Tuomisto, O. Kovero, P.L.Lukinmaa, S. Alaluusua, J. Tuomisto, M. Viluksela, In utero/lactational 2,3,7,8-tetrachlorodibenzo-p-dioxin exposure impairs molar tooth development inrats, Toxicol. Appl. Pharmacol. 174 (2001) 216–224.

[202] K. Kawajiri, Y. Kobayashi, F. Ohtake, T. Ikuta, Y. Matsushima, J. Mimura, S.Pettersson, R.S. Pollenz, T. Sakaki, T. Hirokawa, T. Akiyama, M. Kurosumi, L.Poellinger, S. Kato, Y. Fujii-Kuriyama, Aryl hydrocarbon receptor suppressesintestinal carcinogenesis in ApcMin/+ mice with natural ligands, Proc. Natl.Acad. Sci. USA 106 (2009) 13481–13486.

[203] A. Kazlauskas, L. Poellinger, I. Pongratz, Evidence that the co-chaperone p23regulates ligand responsiveness of the dioxin (Aryl hydrocarbon) receptor, J.Biol. Chem. 274 (1999) 13519–13524.

[204] A. Kazlauskas, L. Poellinger, I. Pongratz, The immunophilin-like protein XAP2regulates ubiquitination and subcellular localization of the dioxin receptor, J.Biol. Chem. 275 (2000) 41317–41324.

[205] A. Kazlauskas, S. Sundstrom, L. Poellinger, I. Pongratz, The hsp90 chaperonecomplex regulates intracellular localization of the dioxin receptor, Mol. Cell.Biol. 21 (2001) 2594–2607.

[206] S. Ke, A.B. Rabson, J.F. Germino, M.A. Gallo, Y. Tian, Mechanism of suppressionof cytochrome P-450 1A1 expression by tumor necrosis factor-alpha andlipopolysaccharide, J. Biol. Chem. 276 (2001) 39638–39644.

[207] L.B. Kedderis, M.E. Andersen, L.S. Birnbaum, Effect of dose, time, andpretreatment on the biliary excretion and tissue distribution of 2,3,7,8-tetrachlorodibenzo-p-dioxin in the rat, Fund. Appl. Toxicol. 21 (1993) 405–411.

[208] R.E. Keesey, M.D. Hirvonen, Body weight set-points: determination andadjustment, J. Nutr. 127 (1997) 1875S–1883S.

[209] C.K. Kelling, B.J. Christian, S.L. Inhorn, R.E. Peterson, Hypophagia-inducedweight loss in mice, rats, and guinea pigs treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin, Fund. Appl. Toxicol. 5 (1985) 700–712.

[210] G.C. Kennedy, The role of depot fat in the hypothalamic control of food intakein the rat, Proc. Roy. Soc. Lond. B Biol. Sci. 140 (1953) 578–592.

[211] R.J. Kewley, M.L. Whitelaw, A. Chapman-Smith, The mammalian basic helix–loop–helix/PAS family of transcriptional regulators, Int. J. Biochem. Cell Biol.36 (2004) 189–204.

[212] S. Kim, E. Dere, L.D. Burgoon, C.C. Chang, T.R. Zacharewski, Comparativeanalysis of AhR-mediated TCDD-elicited gene expression in human liveradult stem cells, Toxicol. Sci. 112 (2009) 229–244.

[213] S.Y. Kim, J.H. Yang, Neurotoxic effects of 2,3,7,8-tetrachlorodibenzo-p-dioxinin cerebellar granule cells, Exp. Mol. Med. 37 (2005) 58–64.

[214] A. Kimura, T. Naka, T. Nakahama, I. Chinen, K. Masuda, K. Nohara, Y. Fujii-Kuriyama, T. Kishimoto, Aryl hydrocarbon receptor in combination withStat1 regulates LPS-induced inflammatory responses, J. Exp. Med. 206 (2009)2027–2035.

[215] B.M. King, The rise, fall, and resurrection of the ventromedial hypothalamusin the regulation of feeding behavior and body weight, Physiol. Behav. 87(2006) 221–244.

[216] H. Kiviranta, J.T. Tuomisto, J. Tuomisto, E. Tukiainen, T. Vartiainen,Polychlorinated dibenzo-p-dioxins, dibenzofurans, and biphenyls in thegeneral population in Finland, Chemosphere 60 (2005) 854–869.

[217] A. Kobayashi, K. Numayama-Tsuruta, K. Sogawa, Y. Fujii-Kuriyama, CBP/p300functions as a possible transcriptional coactivator of Ah receptor nucleartranslocator (Arnt), J. Biochem. 122 (1997) 703–710.

[218] P. Kobelt, A.-S. Wisser, A. Stengel, M. Goebel, T. Inhoff, S. Noetzel, R.W. Veh, N.Bannert, I. van der Voort, B. Wiedenmann, B.F. Klapp, Y. Taché, H. Mönnikes,Peripheral injection of ghrelin induces Fos expression in the dorsomedialhypothalamic nucleus in rats, Brain Res. 1204 (2008) 77–86.

[219] R.J. Kociba, D.G. Keyes, J.E. Beyer, R.M. Carreon, C.E. Wade, D.A. Dittenber, R.P.Kalnins, L.E. Frauson, C.N. Park, S.D. Barnard, R.A. Hummel, C.G. Humiston,Results of a two-year chronic toxicity and oncogenicity study of 2,3,7,8-tetrachlorodibenzo-p-dioxin in rats, Toxicol. Appl. Pharmacol. 46 (1978)279–303.

[220] J. Koistinen, J. Koivusaari, I. Nuuja, P.J. Vuorinen, J. Paasivirta, J.P. Giesy,2,3,7,8-Tetrachlorodibenzo-p-dioxin equivalents in extracts of Baltic white-tailed sea eagles, Environ. Toxicol. Chem. 16 (1997) 1533–1544.

[221] M. Kojima, H. Hosoda, Y. Date, M. Nakazato, H. Matsuo, K. Kangawa, Ghrelin isa growth-hormone-releasing acylated peptide from stomach, Nature 402(1999) 656–660.

[222] S. Kojima, N. Ueno, A. Asakawa, K. Sagiyama, T. Naruo, S. Mizuno, A. Inui, Arole for pancreatic polypeptide in feeding and body weight regulation,Peptides 28 (2007) 459–463.

[223] A.C. Könner, T. Klöckener, J.C. Brüning, Control of energy homeostasis byinsulin and leptin: targeting the arcuate nucleus and beyond, Physiol. Behav.97 (2009) 632–638.

[224] M. Korkalainen, J. Linden, J. Tuomisto, R. Pohjanvirta, Effect of TCDD on mRNAexpression of genes encoding bHLH/PAS proteins in rat hypothalamus,Toxicology 208 (2005) 1–11.

[225] M. Korkalainen, J. Tuomisto, R. Pohjanvirta, Restructured transactivationdomain in hamster AH receptor, Biochem. Biophys. Res. Commun. 273 (2000)272–281.

[226] M. Korkalainen, J. Tuomisto, R. Pohjanvirta, Identification of novel splicevariants of ARNT and ARNT2 in the rat, Biochem. Biophys. Res. Commun. 303(2003) 1095–1100.

[227] M.K. Korkalainen, A.R. Torronen, S.O. Karenlampi, Comparison of expressionof aldehyde dehydrogenase 3 and CYP1A1 in dominant and recessive arylhydrocarbon hydroxylase-deficient mutant mouse hepatoma cells, Chem.Biol. Interact. 94 (1995) 121–134.

[228] D.D. Krahn, B.A. Gosnell, A.S. Levine, J.E. Morley, Behavioral effects ofcorticotropin-releasing factor: localization and characterization of centraleffects, Brain Res. 443 (1988) 63–69.

[229] S. Kress, W.F. Greenlee, Cell-specific regulation of human CYP1A1 andCYP1B1 genes, Cancer Res. 57 (1997) 1264–1269.

[230] S. Kress, J. Reichert, M. Schwarz, Functional analysis of the humancytochrome P4501A1 (CYP1A1) gene enhancer, Eur. J. Biochem. 258 (1998)803–812.

[231] S. Krusekopf, I. Roots, A.G. Hildebrandt, U. Kleeberg, Time-dependenttranscriptional induction of CYP1A1, CYP1A2 and CYP1B1 mRNAs by H+/K+-ATPase inhibitors and other xenobiotics, Xenobiotica 33 (2003) 107–118.

[232] K. Kudo, T. Takeuchi, Y. Murakami, M. Ebina, H. Kikuchi, Characterization ofthe region of the aryl hydrocarbon receptor required for ligand dependencyof transactivation using chimeric receptor between Drosophila and Musmusculus, Biochim. Biophys. Acta 1789 (2009) 477–486.

[233] P.S. Kulkarni, J.G. Crespo, C.A.M. Afonso, Dioxins sources and currentremediation technologies – a review, Environ. Int. 34 (2008) 139–153.

[234] M.B. Kumar, G.H. Perdew, Nuclear receptor coactivator SRC-1 interacts withthe Q-rich subdomain of the AhR and modulates its transactivation potential,Gene Expr. 8 (1999) 273–286.

[235] M.B. Kumar, P. Ramadoss, R.K. Reen, J.P. Vanden Heuvel, G.H. Perdew, The Q-rich subdomain of the human Ah receptor transactivation domain is requiredfor dioxin-mediated transcriptional activity, J. Biol. Chem. 276 (2001) 42302–42310.

[236] M.B. Kumar, R.W. Tarpey, G.H. Perdew, Differential recruitment of coactivatorRIP140 by Ah and estrogen receptors. Absence of a role for LXXLL motifs, J.Biol. Chem. 274 (1999) 22155–22164.

[237] N. Kuramoto, K. Baba, K. Gion, C. Sugiyama, H. Taniura, Y. Yoneda, Xenobioticresponse element binding enriched in both nuclear and microsomal fractionsof rat cerebellum, J. Neurochem. 85 (2003) 264–273.

[238] H. Kurita, W. Yoshioka, N. Nishimura, N. Kubota, T. Kadowaki, C. Tohyama,Aryl hydrocarbon receptor-mediated effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin on glucose-stimulated insulin secretion in mice, J. Appl. Toxicol. 29(2009) 689–694.

Page 22: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 473

[239] M.J. Kwon, K.S. Jeong, E.J. Choi, B.H. Lee, 2,3,7,8-Tetrachlorodibenzo-p-dioxin(TCDD)-induced activation of mitogen-activated protein kinase signalingpathway in Jurkat T cells, Pharmacol. Toxicol. 93 (2003) 186–190.

[240] G.P. Lahvis, C.A. Bradfield, Ahr null alleles: distinctive or different?, BiochemPharmacol. 56 (1998) 781–787.

[241] G.P. Lahvis, S.L. Lindell, R.S. Thomas, R.S. McCuskey, C. Murphy, E. Glover, M.Bentz, J. Southard, C.A. Bradfield, Portosystemic shunting and persistent fetalvascular structures in aryl hydrocarbon receptor-deficient mice, Proc. Natl.Acad. Sci. USA 97 (2000) 10442–10447.

[242] G.P. Lahvis, R.W. Pyzalski, E. Glover, H.C. Pitot, M.K. McElwee, C.A. Bradfield,The aryl hydrocarbon receptor is required for developmental closure of theductus venosus in the neonatal mouse, Mol. Pharmacol. 67 (2005) 714–720.

[243] C.K. Lam, M. Chari, T.K. Lam, CNS regulation of glucose homeostasis,Physiology (Bethesda) 24 (2009) 159–170.

[244] T.K. Lam, G.J. Schwartz, L. Rossetti, Hypothalamic sensing of fatty acids, Nat.Neurosci. 8 (2005) 579–584.

[245] J.A. Lavine, A.J. Rowatt, T. Klimova, A.J. Whitington, E. Dengler, C. Beck, W.H.Powell, Aryl hydrocarbon receptors in the frog Xenopus laevis: two AhR1paralogs exhibit low affinity for 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD),Toxicol. Sci. 88 (2005) 60–72.

[246] R.M. Lechan, C. Fekete, The TRH neuron: a hypothalamic integrator of energymetabolism, Prog. Brain Res. 153 (2006) 209–235.

[247] P.M. Lemieux, C.C. Lutes, D.A. Santoianni, Emissions of organic air toxics fromopen burning: a comprehensive review, Prog. Energy Combust. Sci. 30 (2004)1–32.

[248] P. Lesca, B. Peryt, G. Larrieu, M. Alvinerie, P. Galtier, M. Daujat, P. Maurel, L.Hoogenboom, Evidence for the ligand-independent activation of the AHreceptor, Biochem. Biophys. Res. Commun. 209 (1995) 474–482.

[249] B.E. Levin, Developmental gene � environment interactions affecting systemsregulating energy homeostasis and obesity, Front. Neuroendocrinol. (2010),doi:10.1016/j.yfrne.2010.02.005.

[250] W. Li, S. Donat, O. Dohr, K. Unfried, J. Abel, Ah receptor in different tissues ofC57BL/6J and DBA/2J mice. use of competitive polymerase chain reaction tomeasure Ah-receptor mRNA expression, Arch. Biochem. Biophys. 315 (1994)279–284.

[251] A.K. Liem, P. Furst, C. Rappe, Exposure of populations to dioxins and relatedcompounds, Food Addit. Contam. 17 (2000) 241–259.

[252] S. Lignell, M. Aune, P.O. Darnerud, S. Cnattingius, A. Glynn, Persistentorganochlorine and organobromine compounds in mother’s milk fromSweden 1996–2006: compound-specific temporal trends, Environ. Res. 109(2009) 760–767.

[253] C.H. Lin, C.C. Chen, C.M. Chou, C.Y. Wang, C.C. Hung, J.Y. Chen, H.W. Chang,Y.C. Chen, G.C. Yeh, Y.H. Lee, Knockdown of the aryl hydrocarbon receptorattenuates excitotoxicity and enhances NMDA-induced BDNF expression incortical neurons, J. Neurochem. 111 (2009) 777–789.

[254] C.H. Lin, S.H. Juan, C.Y. Wang, Y.Y. Sun, C.M. Chou, S.F. Chang, S.Y. Hu, W.S.Lee, Y.H. Lee, Neuronal activity enhances aryl hydrocarbon receptor-mediated gene expression and dioxin neurotoxicity in cortical neurons, J.Neurochem. 104 (2008) 1415–1429.

[255] T.M. Lin, K. Ko, R.W. Moore, D.L. Buchanan, P.S. Cooke, R.E. Peterson, Role ofthe aryl hydrocarbon receptor in the development of control and 2,3,7,8-tetrachlorodibenzo-p-dioxin-exposed male mice, J. Toxicol. Environ. Health A64 (2001) 327–342.

[256] T.M. Lin, K. Ko, R.W. Moore, U. Simanainen, T.D. Oberley, R.E. Peterson, Effectsof aryl hydrocarbon receptor null mutation and in utero and lactational2,3,7,8-tetrachlorodibenzo-p-dioxin exposure on prostate and seminalvesicle development in C57BL/6 mice, Toxicol. Sci. 68 (2002) 479–487.

[257] M.C. Lindebro, L. Poellinger, M.L. Whitelaw, Protein-protein interaction viaPAS domains: role of the PAS domain in positive and negative regulation ofthe bHLH/PAS dioxin receptor-Arnt transcription factor complex, EMBO J. 14(1995) 3528–3539.

[258] J. Linden, M. Korkalainen, S. Lensu, J. Tuomisto, R. Pohjanvirta, Effects of2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) and leptin on hypothalamicmRNA expression of factors participating in food intake regulation in aTCDD-sensitive and a TCDD-resistant rat strain, J. Biochem. Mol. Toxicol. 19(2005) 139–148.

[259] G. Ling, J. Gu, M.B. Genter, X. Zhuo, X. Ding, Regulation of cytochrome P450gene expression in the olfactory mucosa, Chem. Biol. Interact. 147 (2004)247–258.

[260] P.C. Liu, F. Matsumura, Differential effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin on the ‘‘adipose- type” and ‘‘brain-type” glucose transporters in mice,Mol. Pharmacol. 47 (1995) 65–73.

[261] A.K. Lund, L.N. Agbor, N. Zhang, A. Baker, H. Zhao, G.D. Fink, N.L. Kanagy, M.K.Walker, Loss of the aryl hydrocarbon receptor induces hypoxemia,endothelin-1, and systemic hypertension at modest altitude, Hypertension51 (2008) 803–809.

[262] J. Ma, Y. Horii, J. Cheng, W. Wang, Q. Wu, T. Ohura, K. Kannan, Chlorinated andparent polycyclic aromatic hydrocarbons in environmental samples from anelectronic waste recycling facility and a chemical industrial complex inChina, Environ. Sci. Technol. 43 (2009) 643–649.

[263] Q. Ma, K.T. Baldwin, 2,3,7,8-tetrachlorodibenzo-p-dioxin-induced degradationof aryl hydrocarbon receptor (AhR) by the ubiquitin-proteasome pathway. Roleof the transcription activation and DNA binding of AhR, J. Biol. Chem. 275(2000) 8432–8438.

[264] Q. Ma, K. Kinneer, Y. Bi, J.Y. Chan, Y.W. Kan, Induction of murine NAD(P)H:quinone oxidoreductase by 2,3,7,8-tetrachlorodibenzo-p-dioxin requires the

CNC (cap ‘n’ collar) basic leucine zipper transcription factor Nrf2 (nuclearfactor erythroid 2-related factor 2): cross-interaction between AhR (arylhydrocarbon receptor) and Nrf2 signal transduction, Biochem. J. 377 (2004)205–213.

[265] Q. Ma, J.P. Whitlock Jr., A novel cytoplasmic protein that interacts with the Ahreceptor, contains tetratricopeptide repeat motifs, and augments thetranscriptional response to 2,3,7,8-tetrachlorodibenzo-p-dioxin, J. Biol.Chem. 272 (1997) 8878–8884.

[266] T.A. Mably, D.L. Bjerke, R.W. Moore, A. Gendron-Fitzpatrick, R.E. Peterson, Inutero and lactational exposure of male rats to 2,3,7,8-tetrachlorodibenzo-p-dioxin. 3. Effects on spermatogenesis and reproductive capability, Toxicol.Appl. Pharmacol. 114 (1992) 118–126.

[267] P.K. Mandal, Dioxin: a review of its environmental effects and its arylhydrocarbon receptor biology, J. Comp. Physiol. B. 175 (2005) 221–230.

[268] L. Marple, B. Berridge, L. Throop, Measurement of the water octanol partition-coefficient of 2,3,7,8-tetrachlorodibenzo-para-dioxin, Environ. Sci. Technol.20 (1986) 397–399.

[269] L. Marple, R. Brunck, L. Throop, Water Solubility of 2,3,7,8-Tetrachlorodibenzo-Para-Dioxin, Environ. Sci. Technol. 20 (1986) 180–182.

[270] M. Matsuda, R.A. DeFronzo, Insulin sensitivity indices obtained from oralglucose tolerance testing: comparison with the euglycemic insulin clamp,Diabetes Care 22 (1999) 1462–1470.

[271] F. Matsumura, The significance of the nongenomic pathway in mediatinginflammatory signaling of the dioxin-activated Ah receptor to cause toxiceffects, Biochem. Pharmacol. 77 (2009) 608–626.

[272] J. Matthews, J.A. Gustafsson, Estrogen receptor and aryl hydrocarbon receptorsignaling pathways, Nucl. Recept. Signal. 4 (2006) e016.

[273] J. Matthews, B. Wihlen, J. Thomsen, J.A. Gustafsson, Aryl hydrocarbonreceptor-mediated transcription: ligand-dependent recruitment of estrogenreceptor alpha to 2,3,7,8-tetrachlorodibenzo-p-dioxin-responsive promoters,Mol. Cell. Biol. 25 (2005) 5317–5328.

[274] G. McKay, Dioxin characterisation, formation and minimisation duringmunicipal solid waste (MSW) incineration: review, Chem. Eng. J. 86 (2002)343–368.

[275] B.J. McMillan, C.A. Bradfield, The aryl hydrocarbon receptor is activated bymodified low-density lipoprotein, Proc. Natl. Acad. Sci. USA 104 (2007) 1412–1417.

[276] B. Meister, Neurotransmitters in key neurons of the hypothalamus thatregulate feeding behavior and body weight, Physiol. Behav. 92 (2007) 263–271.

[277] J. Mimura, M. Ema, K. Sogawa, Y. Fujii-Kuriyama, Identification of a novelmechanism of regulation of Ah (dioxin) receptor function, Genes Dev. 13(1999) 20–25.

[278] J. Mimura, K. Yamashita, K. Nakamura, M. Morita, T.N. Takagi, K. Nakao, M.Ema, K. Sogawa, M. Yasuda, M. Katsuki, Y. Fujii-Kuriyama, Loss of teratogenicresponse to 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in mice lacking theAh (dioxin) receptor, Genes Cells 2 (1997) 645–654.

[279] P. Mocarelli, P.M. Gerthoux, E. Ferrari, D.G. Patterson Jr., S.M. Kieszak, P.Brambilla, N. Vincoli, S. Signorini, P. Tramacere, V. Carreri, E.J. Sampson, W.E.Turner, L.L. Needham, Paternal concentrations of dioxin and sex ratio ofoffspring, Lancet 355 (2000) 1858–1863.

[280] P. Mocarelli, L.L. Needham, A. Marocchi, D.G. Patterson Jr., P. Brambilla, P.M.Gerthoux, L. Meazza, V. Carreri, Serum concentrations of 2,3,7,8-tetrachlorodibenzo-p-dioxin and test results from selected residents ofSeveso, Italy, J. Toxicol. Environ. Health 32 (1991) 357–366.

[281] I.D. Moffat, P.C. Boutros, T. Celius, J. Linden, R. Pohjanvirta, A.B. Okey,MicroRNAs in adult rodent liver are refractory to dioxin treatment, Toxicol.Sci. 99 (2007) 470–487.

[282] P. Monteiro, D. Gilot, E. Le Ferrec, C. Rauch, D. Lagadic-Gossmann, O. Fardel,Dioxin-mediated up-regulation of aryl hydrocarbon receptor target genes isdependent on the calcium/calmodulin/CaMKIalpha pathway, Mol.Pharmacol. 73 (2008) 769–777.

[283] B.H. Moon, C.G. Hong, S.Y. Kim, H.J. Kim, S.K. Shin, S. Kang, K.J. Lee, Y.K. Kim,M.S. Lee, K.H. Shin, A single administration of 2,3,7,8-tetrachlorodibenzo-p-dioxin that produces reduced food and water intake induces long-lastingexpression of corticotropin-releasing factor, arginine vasopressin, andproopiomelanocortin in rat brain, Toxicol. Appl. Pharmacol. 233 (2008)314–322.

[284] P.J. Morgan, A.W. Ross, J.G. Mercer, P. Barrett, What can we learn fromseasonal animals about the regulation of energy balance?, Prog Brain Res. 153(2006) 325–337.

[285] J.E. Morley, M.M. Alshaher, S.A. Farr, J.F. Flood, V.B. Kumar, Leptin andneuropeptide Y (NPY) modulate nitric oxide synthase: further evidence for arole of nitric oxide in feeding, Peptides 20 (1999) 595–600.

[286] J.E. Morley, S.A. Farr, Cachexia and neuropeptide Y, Nutrition 24 (2008) 815–819.

[287] D.L. Morris, L. Rui, Recent advances in understanding leptin signaling andleptin resistance, Am. J. Physiol. Endocrinol. Metab. 297 (2009) E1247–E1259.

[288] D.C. Morse, A.P. Stein, P.E. Thomas, H.E. Lowndes, Distribution and inductionof cytochrome P450 1A1 and 1A2 in rat brain, Toxicol. Appl. Pharmacol. 152(1998) 232–239.

[289] G.J. Morton, D.E. Cummings, D.G. Baskin, G.S. Barsh, M.W. Schwartz, Centralnervous system control of food intake and body weight, Nature 443 (2006)289–295.

[290] N.A. Mufti, N.A. Bleckwenn, J.G. Babish, M.L. Shuler, Possible involvement ofthe Ah receptor in the induction of cytochrome P-450IA1 under conditions of

Page 23: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

474 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

hydrodynamic shear in microcarrier-attached hepatoma cell lines, Biochem.Biophys. Res. Commun. 208 (1995) 144–152.

[291] N.A. Mufti, M.L. Shuler, Induction of cytochrome P-450IA1 activity inresponse to sublethal stresses in microcarrier-attached Hep G2 cells,Biotechnol. Prog. 11 (1995) 659–663.

[292] N.A. Mufti, M.L. Shuler, Possible role of arachidonic acid in stress-inducedcytochrome P450IA1 activity, Biotechnol. Prog. 12 (1996) 847–854.

[293] M. Mukai, T.M. Lin, R.E. Peterson, P.S. Cooke, S.A. Tischkau, Behavioralrhythmicity of mice lacking AhR and attenuation of light-induced phase shiftby 2,3,7,8-tetrachlorodibenzo-p-dioxin, J. Biol. Rhythms 23 (2008) 200–210.

[294] M. Mukai, S.A. Tischkau, Effects of tryptophan photoproducts in the circadiantiming system: searching for a physiological role for aryl hydrocarbonreceptor, Toxicol. Sci. 95 (2007) 172–181.

[295] I.A. Murray, G. Krishnegowda, B.C. DiNatale, C. Flaveny, C. Chiaro, J.M. Lin, A.K.Sharma, S. Amin, G.H. Perdew, Development of a selective modulator of arylhydrocarbon (Ah) receptor activity that exhibits anti-inflammatoryproperties, Chem. Res. Toxicol. 23 (2010) 955–966.

[296] I.A. Murray, J.L. Morales, C.A. Flaveny, B.C. Dinatale, C. Chiaro, K. Gowdahalli,S. Amin, G.H. Perdew, Evidence for ligand-mediated selective modulation ofaryl hydrocarbon receptor activity, Mol. Pharmacol. 77 (2010) 247–254.

[297] I.A. Murray, J.L. Morales, C.A. Flaveny, B.C. Dinatale, C.R. Chiaro, K. Gowdahalli,S.G. Amin, G.H. Perdew, Evidence for ligand-mediated selective modulation ofaryl hydrocarbon receptor activity, Mol. Pharmacol. 77 (2010) 247–254.

[298] G. Muzi, J.R. Gorski, K. Rozman, Composition of diet modifies toxicity of2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in cold-adapted rats, Arch.Toxicol. 61 (1987) 34–39.

[299] G. Muzi, J.R. Gorski, K. Rozman, Mode of metabolism is altered in 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)-treated rats, Toxicol. Lett. 47 (1989) 77–86.

[300] M.G. Myers Jr., H. Munzberg, G.M. Leinninger, R.L. Leshan, The geometry ofleptin action in the brain: more complicated than a simple ARC, Cell. Metab. 9(2009) 117–123.

[301] M.G. Myers, M.A. Cowley, H. Munzberg, Mechanisms of leptin action andleptin resistance, Annu. Rev. Physiol. 70 (2008) 537–556.

[302] S.C. Nair, E.J. Toran, R.A. Rimerman, S. Hjermstad, T.E. Smithgall, D.F. Smith, Apathway of multi-chaperone interactions common to diverse regulatoryproteins: estrogen receptor, Fes tyrosine kinase, heat shock transcriptionfactor Hsf1, and the aryl hydrocarbon receptor, Cell Stress Chaperones 1(1996) 237–250.

[303] L.P. Nguyen, E.L. Hsu, G. Chowdhury, M. Dostalek, F.P. Guengerich, C.A.Bradfield, D-amino acid oxidase generates agonists of the aryl hydrocarbonreceptor from D-Tryptophan, Chem. Res. Toxicol. 22 (2009) 1897–1904.

[304] T.A. Nguyen, D. Hoivik, J.E. Lee, S. Safe, Interactions of nuclear receptorcoactivator/corepressor proteins with the aryl hydrocarbon receptorcomplex, Arch. Biochem. Biophys. 367 (1999) 250–257.

[305] N. Nishimura, F. Matsumura, C.F. Vogel, H. Nishimura, J. Yonemoto, W.Yoshioka, C. Tohyama, Critical role of cyclooxygenase-2 activation inpathogenesis of hydronephrosis caused by lactational exposure of mice todioxin, Toxicol. Appl. Pharmacol. 231 (2008) 374–383.

[306] N. Nishimura, J. Yonemoto, Y. Miyabara, Y. Fujii-Kuriyama, C. Tohyama,Altered thyroxin and retinoid metabolic response to 2,3,7,8-tetrachlorodibenzo-p-dioxin in aryl hydrocarbon receptor-null mice, Arch.Toxicol. 79 (2005) 260–267.

[307] S. Nishiumi, N. Yamamoto, R. Kodoi, I. Fukuda, K. Yoshida, H. Ashida,Antagonistic and agonistic effects of indigoids on the transformation of anaryl hydrocarbon receptor, Arch. Biochem. Biophys. 470 (2008) 187–199.

[308] M. Nishizumi, Y. Higaki, Studies on mechanism of wasting syndrome inTCDD-intoxicated rats, Fukuoka Igaku Zasshi 88 (1997) 200–204.

[309] K. Noren, D. Meironyte, Certain organochlorine and organobrominecontaminants in Swedish human milk in perspective of past 20–30 years,Chemosphere 40 (2000) 1111–1123.

[310] M. Novelli, S. Piaggi, V. De Tata, 2,3,7,8-Tetrachlorodibenzo-p-dioxin-inducedimpairment of glucose-stimulated insulin secretion in isolated rat pancreaticislets, Toxicol. Lett. 156 (2005) 307–314.

[311] M. Nukaya, S. Moran, C.A. Bradfield, The role of the dioxin-responsiveelement cluster between the Cyp1a1 and Cyp1a2 loci in aryl hydrocarbonreceptor biology, Proc. Natl. Acad. Sci. USA 106 (2009) 4923–4928.

[312] B. Oesch-Bartlomowicz, A. Huelster, O. Wiss, P. Antoniou-Lipfert, C. Dietrich,M. Arand, C. Weiss, E. Bockamp, F. Oesch, Aryl hydrocarbon receptoractivation by cAMP vs. dioxin: divergent signaling pathways, Proc. Natl.Acad. Sci. USA 102 (2005) 9218–9223.

[313] F. Ohtake, A. Baba, I. Takada, M. Okada, K. Iwasaki, H. Miki, S. Takahashi, A.Kouzmenko, K. Nohara, T. Chiba, Y. Fujii-Kuriyama, S. Kato, Dioxin receptor isa ligand-dependent E3 ubiquitin ligase, Nature 446 (2007) 562–566.

[314] A.B. Okey, M.A. Franc, I.D. Moffat, N. Tijet, P.C. Boutros, M. Korkalainen, J.Tuomisto, R. Pohjanvirta, Toxicological implications of polymorphisms inreceptors for xenobiotic chemicals: the case of the aryl hydrocarbon receptor,Toxicol. Appl. Pharmacol. 207 (2005) 43–51.

[315] S.T. Okino, D. Pookot, S. Basak, R. Dahiya, Toxic and chemopreventive ligandspreferentially activate distinct aryl hydrocarbon receptor pathways:implications for cancer prevention, Cancer. Prev. Res. 2 (2009) 251–256.

[316] S.T. Okino, L.C. Quattrochi, D. Pookot, M. Iwahashi, R. Dahiya, A dioxin-responsive enhancer 30 of the human CYP1A2 gene, Mol. Pharmacol. 72(2007) 1457–1465.

[317] J.R. Olson, M.A. Holscher, R.A. Neal, Toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin in the golden Syrian hamster, Toxicol. Appl. Pharmacol. 55 (1980) 67–78.

[318] M. Oshima, J. Mimura, H. Sekine, H. Okawa, Y. Fujii-Kuriyama, SUMOmodification regulates the transcriptional repressor function of arylhydrocarbon receptor repressor, J. Biol. Chem. 284 (2009) 11017–11026.

[319] G. Paajarvi, M. Viluksela, R. Pohjanvirta, U. Stenius, J. Hogberg, TCDD activatesMdm2 and attenuates the p53 response to DNA damaging agents,Carcinogenesis 26 (2005) 201–208.

[320] S.J. Park, W.K. Yoon, H.J. Kim, H.Y. Son, S.W. Cho, K.S. Jeong, T.H. Kim, S.H. Kim,S.R. Kim, S.Y. Ryu, 2,3,7,8-Tetrachlorodibenzo-p-dioxin activates ERK and p38mitogen-activated protein kinases in RAW 264.7 cells, Anticancer Res. 25(2005) 2831–2836.

[321] R.D. Patel, I.A. Murray, C.A. Flaveny, A. Kusnadi, G.H. Perdew, Ah receptorrepresses acute-phase response gene expression without binding to itscognate response element, Lab. Invest. 89 (2009) 695–707.

[322] D.G. Patterson Jr., W.E. Turner, S.P. Caudill, L.L. Needham, Total TEQ referencerange (PCDDs, PCDFs, cPCBs, mono-PCBs) for the US population 2001–2002,Chemosphere 73 (2008) S261–S277.

[323] G.H. Perdew, C.A. Bradfield, Mapping the 90 kDa heat shock protein bindingregion of the Ah receptor, Biochem. Mol. Biol. Int. 39 (1996) 589–593.

[324] A.C. Pesatori, D. Consonni, M. Rubagotti, P. Grillo, P.A. Bertazzi, Cancerincidence in the population exposed to dioxin after the ‘‘Seveso accident”:twenty years of follow-up, Environ. Health 8 (2009) 39.

[325] F. Peter Guengerich, M.V. Martin, W.A. McCormick, L.P. Nguyen, E. Glover,C.A. Bradfield, Aryl hydrocarbon receptor response to indigoids in vitro andin vivo, Arch. Biochem. Biophys. 423 (2004) 309–316.

[326] S.L. Petersen, M.A. Curran, S.A. Marconi, C.D. Carpenter, L.S. Lubbers, M.D.McAbee, Distribution of mRNAs encoding the arylhydrocarbon receptor,arylhydrocarbon receptor nuclear translocator, and arylhydrocarbon receptornuclear translocator-2 in the rat brain and brainstem, J. Comp. Neurol. 427(2000) 428–439.

[327] S.L. Petersen, S. Krishnan, E.D. Hudgens, The aryl hydrocarbon receptorpathway and sexual differentiation of neuroendocrine functions,Endocrinology 147 (2006) S33–S42.

[328] J.R. Petrulis, A. Kusnadi, P. Ramadoss, B. Hollingshead, G.H. Perdew, Thehsp90 Co-chaperone XAP2 alters importin beta recognition of the bipartitenuclear localization signal of the Ah receptor and represses transcriptionalactivity, J. Biol. Chem. 278 (2003) 2677–2685.

[329] J.R. Petrulis, G.H. Perdew, The role of chaperone proteins in the arylhydrocarbon receptor core complex, Chem. Biol. Interact. 141 (2002) 25–40.

[330] D. Phelan, G.M. Winter, W.J. Rogers, J.C. Lam, M.S. Denison, Activation of theAh receptor signal transduction pathway by bilirubin and biliverdin, Arch.Biochem. Biophys. 357 (1998) 155–163.

[331] R. Pohjanvirta, Studies on the mechanism of acute toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in rats, Publications of the NationalPublic Health Institute (NPHI) A1 (Thesis), 1991.

[332] R. Pohjanvirta, Transgenic mouse lines expressing rat AH receptor variants –a new animal model for research on AH receptor function and dioxin toxicitymechanisms, Toxicol. Appl. Pharmacol. 236 (2009) 166–182.

[333] R. Pohjanvirta, R. Juvonen, S. Karenlampi, H. Raunio, J. Tuomisto, Hepatic Ah-receptor levels and the effect of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)on hepatic microsomal monooxygenase activities in a TCDD-susceptible and-resistant rat strain, Toxicol. Appl. Pharmacol. 92 (1988) 131–140.

[334] R. Pohjanvirta, M. Korkalainen, J. McGuire, U. Simanainen, R. Juvonen, J.T.Tuomisto, M. Unkila, M. Viluksela, J. Bergman, L. Poellinger, J. Tuomisto,Comparison of acute toxicities of indolo[3, 2-b]carbazole (ICZ) and 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in TCDD-sensitive rats, Food Chem.Toxicol. 40 (2002) 1023–1032.

[335] R. Pohjanvirta, T. Kulju, A.F. Morselt, R. Tuominen, R. Juvonen, K. Rozman, P.Mannisto, Y. Collan, E.L. Sainio, J. Tuomisto, Target tissue morphology andserum biochemistry following 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)exposure in a TCDD-susceptible and a TCDD-resistant rat strain, Fund. Appl.Toxicol. 12 (1989) 698–712.

[336] R. Pohjanvirta, J. Tuomisto, 2,3,7,8-Tetrachlorodibenzo-p-dioxin enhancesresponsiveness to post-ingestive satiety signals, Toxicology 63 (1990) 285–299.

[337] R. Pohjanvirta, J. Tuomisto, Remarkable residual alterations in responses tofeeding regulatory challenges in Han/Wistar rats after recovery from theacute toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD), Food Chem.Toxicol. 28 (1990) 677–686.

[338] R. Pohjanvirta, J. Tuomisto, Short-term toxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin in laboratory animals: effects, mechanisms, and animal models,Pharmacol. Rev. 46 (1994) 483–549.

[339] R. Pohjanvirta, J. Tuomisto, K. Vikkula, Screening of pharmacological agentsgiven peripherally with respect to TCDD-induced wasting syndrome in Long-Evans rats, Pharmacol. Toxicol. 63 (1988) 240–247.

[340] R. Pohjanvirta, M. Unkila, J. Tuomisto, The loss of glucoprivic feeding is anearly-stage alteration in TCDD-treated Han/Wistar rats, Pharmacol. Toxicol.67 (1990) 441–443.

[341] R. Pohjanvirta, M. Unkila, J. Tuomisto, Characterization of the enhancedresponsiveness to postingestive satiety signals in 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)-treated Han/Wistar rats, Pharmacol. Toxicol. 69 (1991) 433–441.

Page 24: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 475

[342] R. Pohjanvirta, M. Unkila, J. Tuomisto, TCDD-induced hypophagia is notexplained by nausea, Pharmacol. Biochem. Behav. 47 (1994) 273–282.

[343] R. Pohjanvirta, M. Unkila, J.T. Tuomisto, J. Tuomisto, Modulation of TCDD-induced wasting syndrome by diabetes, Organohalogen Compd. 21 (1994)315–318.

[344] R. Pohjanvirta, T. Vartiainen, A. Uusi-Rauva, J. Monkkonen, J. Tuomisto, Tissuedistribution, metabolism, and excretion of 14C-TCDD in a TCDD-susceptibleand a TCDD-resistant rat strain, Pharmacol. Toxicol. 66 (1990) 93–100.

[345] R. Pohjanvirta, J.M. Wong, W. Li, P.A. Harper, J. Tuomisto, A.B. Okey, Pointmutation in intron sequence causes altered carboxyl-terminal structure inthe aryl hydrocarbon receptor of the most 2,3,7,8-tetrachlorodibenzo-p-dioxin-resistant rat strain, Mol. Pharmacol. 54 (1998) 86–93.

[346] A. Poland, E. Glover, Comparison of 2,3,7,8-tetrachlorodibenzo-p-dioxin, apotent inducer of aryl hydrocarbon hydroxylase, with 3-methylcholanthrene,Mol. Pharmacol. 10 (1974) 349–359.

[347] A. Poland, J.C. Knutson, 2,3,7,8-Tetrachlorodibenzo-P-dioxin and relatedhalogenated aromatic hydrocarbons: examination of the mechanism oftoxicity, Annu. Rev. Pharmacol. Toxicol. 22 (1982) 517–554.

[348] A. Poland, D. Palen, E. Glover, Analysis of the four alleles of the murine arylhydrocarbon receptor, Mol. Pharmacol. 46 (1994) 915–921.

[349] A. Poland, P. Teitelbaum, E. Glover, [125I] 2-iodo-3,7,8-trichlorodibenzo-p-dioxin-binding species in mouse liver induced by agonists for the Ahreceptor: characterization and identification, Mol. Pharmacol. 36 (1989)113–120.

[350] A. Poland, P. Teitelbaum, E. Glover, A. Kende, Stimulation of in vivo hepaticuptake and in vitro hepatic binding of [125I] 2-lodo-3,7,8-trichlorodibenzo-p-dioxin by the administration of agonist for the Ah receptor, Mol.Pharmacol. 36 (1989) 121–127.

[351] I. Pongratz, C. Antonsson, M.L. Whitelaw, L. Poellinger, Role of the PASdomain in regulation of dimerization and DNA binding specificity of thedioxin receptor, Mol. Cell. Biol. 18 (1998) 4079–4088.

[352] K.A. Posey, D.J. Clegg, R.L. Printz, J. Byun, G.J. Morton, A. Vivekanandan-Giri, S.Pennathur, D.G. Baskin, J.W. Heinecke, S.C. Woods, M.W. Schwartz, K.D.Niswender, Hypothalamic proinflammatory lipid accumulation,inflammation, and insulin resistance in rats fed a high-fat diet, Am. J.Physiol. Endocrinol. Metab. 296 (2009) E1003–E1012.

[353] M. Potier, N. Darcel, D. Tome, Protein, amino acids and the control of foodintake, Curr. Opin. Clin. Nutr. Metab. Care 12 (2009) 54–58.

[354] C.L. Potter, L.A. Menahan, R.E. Peterson, Relationship of alterations in energymetabolism to hypophagia in rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin, Fund. Appl. Toxicol. 6 (1986) 89–97.

[355] C.L. Potter, R.W. Moore, S.L. Inhorn, T.C. Hagen, R.E. Peterson, Thyroid statusand thermogenesis in rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin,Toxicol. Appl. Pharmacol. 84 (1986) 45–55.

[356] J.A. Powell-Coffman, C.A. Bradfield, W.B. Wood, Caenorhabditis elegansorthologs of the aryl hydrocarbon receptor and its heterodimerizationpartner the aryl hydrocarbon receptor nuclear translocator, Proc. Natl.Acad. Sci. USA 95 (1998) 2844–2849.

[357] A. Pravettoni, A. Colciago, P. Negri-Cesi, S. Villa, F. Celotti, Ontogeneticdevelopment, sexual differentiation, and effects of Aroclor 1254 exposure onexpression of the arylhydrocarbon receptor and of the arylhydrocarbonreceptor nuclear translocator in the rat hypothalamus, Reprod. Toxicol. 20(2005) 521–530.

[358] A.M. Prentice, B.J. Hennig, A.J. Fulford, Evolutionary origins of the obesityepidemic: natural selection of thrifty genes or genetic drift followingpredation release?, Int J. Obesity 32 (2008) 1607–1610.

[359] A. Puga, C. Ma, J.L. Marlowe, The aryl hydrocarbon receptor cross-talks withmultiple signal transduction pathways, Biochem. Pharmacol. 77 (2009) 713–722.

[360] A. Puga, D.W. Nebert, F. Carrier, Dioxin induces expression of c-fos and c-junproto-oncogenes and a large increase in transcription factor AP-1, DNA CellBiol. 11 (1992) 269–281.

[361] A. Puga, M.A. Sartor, M.Y. Huang, J.K. Kerzee, Y.D. Wei, C.R. Tomlinson, C.S.Baxter, M. Medvedovic, Gene expression profiles of mouse aorta and culturedvascular smooth muscle cells differ widely, yet show common responses todioxin exposure, Cardiovasc. Toxicol. 4 (2004) 385–404.

[362] A. Puga, C.R. Tomlinson, Y. Xia, Ah receptor signals cross-talk with multipledevelopmental pathways, Biochem. Pharmacol. 69 (2005) 199–207.

[363] S.M. Puhvel, M. Sakamoto, D.C. Ertl, R.M. Reisner, Hairless mice as models forchloracne: a study of cutaneous changes induced by topical application ofestablished chloracnegens, Toxicol. Appl. Pharmacol. 64 (1982) 492–503.

[364] P. Pusztai, B. Sarman, E. Ruzicska, J. Toke, K. Racz, A. Somogyi, Z. Tulassay,Ghrelin: a new peptide regulating the neurohormonal system, energyhomeostasis and glucose metabolism, Diabetes Metab. Res. Rev. 24 (2008)343–352.

[365] H. Qin, J.A. Powell-Coffman, The Caenorhabditis elegans aryl hydrocarbonreceptor, AHR-1, regulates neuronal development, Dev. Biol. 270 (2004) 64–75.

[366] H. Qin, Z. Zhai, J.A. Powell-Coffman, The Caenorhabditis elegans AHR-1transcription complex controls expression of soluble guanylate cyclasegenes in the URX neurons and regulates aggregation behavior, Dev. Biol.298 (2006) 606–615.

[367] X. Qu, R.P. Metz, W.W. Porter, V.M. Cassone, D.J. Earnest, Disruption of clockgene expression alters responses of the aryl hydrocarbon receptor signalingpathway in the mouse mammary gland, Mol. Pharmacol. 72 (2007) 1349–1358.

[368] X. Qu, R.P. Metz, W.W. Porter, V.M. Cassone, D.J. Earnest, Disruption of periodgene expression alters the inductive effects of dioxin on the AhR signalingpathway in the mouse liver, Toxicol. Appl. Pharmacol. 234 (2009) 370–377.

[369] U. Quass, M. Fermann, G. Broker, The European dioxin air emission inventoryproject – final results, Chemosphere 54 (2004) 1319–1327.

[370] A. Raasmaja, M. Viluksela, K.K. Rozman, Decreased liver type I 50-deiodinaseand increased brown adipose tissue type II 50-deiodinase activity in 2,3,7,8-tetrachlorobibenzo-p-dioxin (TCDD)-treated Long-Evans rats, Toxicology 114(1996) 199–205.

[371] P. Ramadoss, G.H. Perdew, The transactivation domain of the Ah receptor is akey determinant of cellular localization and ligand-independentnucleocytoplasmic shuttling properties, Biochemistry 44 (2005) 11148–11159.

[372] E.J. Ramos, M.M. Meguid, A.C. Campos, J.C. Coelho, Neuropeptide Y, alpha-melanocyte-stimulating hormone, and monoamines in food intakeregulation, Nutrition 21 (2005) 269–279.

[373] U. Rannug, M. Sjogren, A. Rannug, M. Gillner, R. Toftgard, J.A. Gustafsson, H.Rosenkranz, G. Klopman, Use of artificial intelligence in structure-affinitycorrelations of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) receptor ligands,Carcinogenesis 12 (1991) 2007–2015.

[374] C. Rappe, Dioxin, patterns and source identification, Fresen. J. Anal. Chem.348 (1994) 63–75.

[375] S.S. Ray, H.I. Swanson, Dioxin-induced immortalization of normal humankeratinocytes and silencing of p53 and p16INK4a, J. Biol. Chem. 279 (2004)27187–27193.

[376] S. Reilly, M.A. Bornovalova, Conditioned taste aversion and amygdala lesionsin the rat: a critical review, Neurosci. Biobehav. Rev. 29 (2005) 1067–1088.

[377] H. Reyes, S. Reisz-Porszasz, O. Hankinson, Identification of the Ah receptornuclear translocator protein (Arnt) as a component of the DNA binding formof the Ah receptor, Science 256 (1992) 1193–1195.

[378] V.M. Richardson, M.J. Santostefano, L.S. Birnbaum, Daily cycle of bHLH-PASproteins, Ah receptor and Arnt, in multiple tissues of female Sprague-Dawleyrats, Biochem. Biophys. Res. Commun. 252 (1998) 225–231.

[379] C.A. Richter, D.E. Tillitt, M. Hannink, Regulation of subcellular localization ofthe aryl hydrocarbon receptor (AhR), Arch. Biochem. Biophys. 389 (2001)207–217.

[380] D.S. Riddick, Y. Huang, P.A. Harper, A.B. Okey, 2,3,7,8-Tetrachlorodibenzo-p-dioxin versus 3-methylcholanthrene: comparative studies of Ah receptorbinding, transformation, and induction of CYP1A1, J. Biol. Chem. 269 (1994)12118–12128.

[381] D.S. Riddick, C. Lee, A. Bhathena, Y.E. Timsit, P.Y. Cheng, E.T. Morgan, R.A.Prough, S.L. Ripp, K.K. Miller, A. Jahan, J.Y. Chiang, Transcriptionalsuppression of cytochrome P450 genes by endogenous and exogenouschemicals, Drug Metab. Dispos. 32 (2004) 367–375.

[382] K. Riecke, D. Grimm, M. Shakibaei, P. Kossmehl, G. Schulze-Tanzil, M. Paul, R.Stahlmann, Low doses of 2,3,7,8-tetrachlorodibenzo- p-dioxin increasetransforming growth factor beta and cause myocardial fibrosis inmarmosets (Callithrix jacchus), Arch. Toxicol. 76 (2002) 360–366.

[383] S. Ritter, Glucoprivation and the glucoprivic control of food intake, in: R.C.Ritter, S. Ritter, C.D. Barnes (Eds.), Feeding Behaviour: Neural and HumoralControls, Academic Press, Orlando, FL (USA), 1986, pp. 271–313.

[384] B.J. Roberts, M.L. Whitelaw, Degradation of the basic helix–loop–helix/Per-ARNT-sim homology domain dioxin receptor via the ubiquitin/proteasomepathway, J. Biol. Chem. 274 (1999) 36351–36356.

[385] A.C. Roman, J.M. Carvajal-Gonzalez, E.M. Rico-Leo, P.M. Fernandez-Salguero, Dioxin receptor deficiency impairs angiogenesis by amechanism involving VEGF-A depletion in the endothelium andtransforming growth factor-beta overexpression in the stroma, J. Biol.Chem. 284 (2009) 25135–25148.

[386] C. Roman, S. Reilly, Effects of insular cortex lesions on conditioned tasteaversion and latent inhibition in the rat, Eur. J. Neurosci. 26 (2007) 2627–2632.

[387] J.D. Roth, H. Maier, S. Chen, B.L. Roland, Implications of amylin receptoragonism: integrated neurohormonal mechanisms and therapeuticapplications, Arch. Neurol. 66 (2009) 306–310.

[388] J.C. Rowlands, I.J. McEwan, J.A. Gustafsson, Trans-activation by the humanaryl hydrocarbon receptor and aryl hydrocarbon receptor nucleartranslocator proteins: direct interactions with basal transcription factors,Mol. Pharmacol. 50 (1996) 538–548.

[389] J. Ruegg, E. Swedenborg, D. Wahlstrom, A. Escande, P. Balaguer, K. Pettersson,I. Pongratz, The transcription factor aryl hydrocarbon receptor nucleartranslocator functions as an estrogen receptor beta-selective coactivator,and its recruitment to alternative pathways mediates antiestrogenic effectsof dioxin, Mol. Endocrinol. 22 (2008) 304–316.

[390] P. Ruokojarvi, M. Ettala, P. Rahkonen, J. Tarhanen, J. Ruuskanen,Polychlorinated dibenzo-P-dioxins and dibenzo-P-furans (Pcdds and Pcdfs)in municipal waste landfill fires, Chemosphere 30 (1995) 1697–1708.

[391] S.R. Rushing, M.S. Denison, The silencing mediator of retinoic acid andthyroid hormone receptors can interact with the aryl hydrocarbon (Ah)receptor but fails to repress Ah receptor-dependent gene expression, Arch.Biochem. Biophys. 403 (2002) 189–201.

[392] S. Safe, Polychlorinated biphenyls (PCBs), dibenzo-p-dioxins (PCDDs),dibenzofurans (PCDFs), and related compounds: environmental andmechanistic considerations which support the development of toxicequivalency factors (TEFs), Crit. Rev. Toxicol. 21 (1990) 51–88.

Page 25: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

476 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

[393] S. Safe, M. Wormke, I. Samudio, Mechanisms of inhibitory aryl hydrocarbonreceptor–estrogen receptor crosstalk in human breast cancer cells, J.Mammary Gland Biol. Neoplasia 5 (2000) 295–306.

[394] S.H. Safe, Comparative toxicology and mechanism of action of polychlorinateddibenzo-p-dioxins and dibenzofurans, Annu. Rev. Pharmacol. Toxicol. 26(1986) 371–399.

[395] A. Sahu, Minireview: a hypothalamic role in energy balance with specialemphasis on leptin, Endocrinology 145 (2004) 2613–2620.

[396] A. Sahu, Effects of chronic central leptin infusion on proopiomelanocortin andneurotensin gene expression in the rat hypothalamus, Neurosci. Lett. 440(2008) 125–129.

[397] C. Sanchez-Lasheras, A.C. Könner, J.C. Brüning, Integrative neurobiology ofenergy homeostasis–neurocircuits, signals and mediators, Front.Neuroendocrinol. 31 (2010) 4–15.

[398] J.T. Sanderson, G.D. Bellward, Hepatic microsomal ethoxyresorufin O-deethylase-inducing potency in ovo and cytosolic Ah receptor bindingaffinity of 2,3,7,8-tetrachlorodibenzo-p-dioxin: comparison of four avianspecies, Toxicol. Appl. Pharmacol. 132 (1995) 131–145.

[399] H. Sano, M. Yokoi, Striatal medium spiny neurons terminate in a distinctregion in the lateral hypothalamic area and do not directly innervate orexin/hypocretin- or melanin-concentrating hormone-containing neurons, J.Neurosci. 27 (2007) 6948–6955.

[400] E. Sasamori, S. Shimoyama, S. Fukushige, H. Kikuchi, Involvement of CREM inCYP1A1 induction through ligand-independent activation pathway of arylhydrocarbon receptor in HepG2 cells, Arch. Biochem. Biophys. 478 (2008)26–35.

[401] S. Sato, H. Shirakawa, S. Tomita, Y. Ohsaki, K. Haketa, O. Tooi, N. Santo, M.Tohkin, Y. Furukawa, F.J. Gonzalez, M. Komai, Low-dose dioxins alter geneexpression related to cholesterol biosynthesis, lipogenesis, and glucosemetabolism through the aryl hydrocarbon receptor-mediated pathway inmouse liver, Toxicol. Appl. Pharmacol. 229 (2008) 10–19.

[402] J.F. Savouret, M. Antenos, M. Quesne, J. Xu, E. Milgrom, R.F. Casper, 7-Ketocholesterol is an endogenous modulator for the arylhydrocarbonreceptor, J. Biol. Chem. 276 (2001) 3054–3059.

[403] C.M. Schaldach, J. Riby, L.F. Bjeldanes, Lipoxin A4: a new class of ligand for theAh receptor, Biochemistry 38 (1999) 7594–7600.

[404] S.L. Schantz, R.E. Bowman, Learning in monkeys exposed perinatally to2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD), Neurotoxicol. Teratol. 11 (1989)13–19.

[405] A. Schecter, O. Papke, K.C. Tung, J. Joseph, T.R. Harris, J. Dahlgren,Polybrominated diphenyl ether flame retardants in the US population:current levels, temporal trends, and comparison with dioxins, dibenzofurans,and polychlorinated biphenyls, J. Occup. Environ. Med. 47 (2005) 199–211.

[406] B. Schilter, C.J. Omiecinski, Regional distribution and expression modulationof cytochrome P-450 and epoxide hydrolase mRNAs in the rat brain, Mol.Pharmacol. 44 (1993) 990–996.

[407] J.V. Schmidt, G.H. Su, J.K. Reddy, M.C. Simon, C.A. Bradfield, Characterizationof a murine Ahr null allele: involvement of the Ah receptor in hepatic growthand development, Proc. Natl. Acad. Sci. USA 93 (1996) 6731–6736.

[408] M. Schuhmacher, H. Kiviranta, P. Ruokojarvi, M. Nadal, J.L. Domingo,Concentrations of PCDD/Fs, PCBs and PBDEs in breast milk of women fromCatalonia, Spain: a follow-up study, Environ. Int. 35 (2009) 607–613.

[409] A.G. Schuur, F.M. Boekhorst, A. Brouwer, T.J. Visser, Extrathyroidal effects of2,3,7,8-tetrachlorodibenzo-p-dioxin on thyroid hormone turnover in maleSprague-Dawley rats, Endocrinology 138 (1997) 3727–3734.

[410] M.W. Schwartz, D.G. Baskin, K.J. Kaiyala, S.C. Woods, Model for the regulationof energy balance and adiposity by the central nervous system, Am. J. Clin.Nutr. 69 (1999) 584–596.

[411] M.W. Schwartz, S.C. Woods, D. Porte Jr., R.J. Seeley, D.G. Baskin, Centralnervous system control of food intake, Nature 404 (2000) 661–671.

[412] M.W. Schwartz, S.C. Woods, R.J. Seeley, G.S. Barsh, D.G. Baskin, R.L. Leibel, Isthe energy homeostasis system inherently biased toward weight gain?,Diabetes 52 (2003) 232–238

[413] E.M. Sciullo, B. Dong, C.F. Vogel, F. Matsumura, Characterization of thepattern of the nongenomic signaling pathway through which TCDD-inducesearly inflammatory responses in U937 human macrophages, Chemosphere74 (2009) 1531–1537.

[414] E.M. Sciullo, C.F. Vogel, W. Li, F. Matsumura, Initial and extendedinflammatory messages of the nongenomic signaling pathway of the TCDD-activated Ah receptor in U937 macrophages, Arch. Biochem. Biophys. 480(2008) 143–155.

[415] A. Sclafani, Effects of gastrointestinal surgery on ingestive behavior inanimals, Gastroenterol. Clin. North Am. 16 (1987) 461–477.

[416] M.D. Seefeld, S.W. Corbett, R.E. Keesey, R.E. Peterson, Characterization of thewasting syndrome in rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin,Toxicol. Appl. Pharmacol. 73 (1984) 311–322.

[417] M.D. Seefeld, R.E. Keesey, R.E. Peterson, Body weight regulation in rats treatedwith 2,3,7,8-tetrachlorodibenzo-p-dioxin, Toxicol. Appl. Pharmacol. 76(1984) 526–536.

[418] M.D. Seefeld, R.E. Peterson, Digestible energy and efficiency of feed utilizationin rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin, Toxicol. Appl.Pharmacol. 74 (1984) 214–222.

[419] H. Sekine, J. Mimura, M. Oshima, H. Okawa, J. Kanno, K. Igarashi, F.J. Gonzalez,T. Ikuta, K. Kawajiri, Y. Fujii-Kuriyama, Hypersensitivity of aryl hydrocarbonreceptor-deficient mice to lipopolysaccharide-induced septic shock, Mol. Cell.Biol. 29 (2009) 6391–6400.

[420] B.W. Seo, B.E. Powers, J.J. Widholm, S.L. Schantz, Radial arm mazeperformance in rats following gestational and lactational exposure to2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD), Neurotoxicol. Teratol. 22(2000) 511–519.

[421] M. Shibazaki, T. Takeuchi, S. Ahmed, H. Kikuchi, Suppression by p38 MAPkinase inhibitors (pyridinyl imidazole compounds) of Ah receptor target geneactivation by 2,3,7,8-tetrachlorodibenzo-p-dioxin and the possiblemechanism, J. Biol. Chem. 279 (2004) 3869–3876.

[422] A.C. Shin, H. Zheng, H.R. Berthoud, An expanded view of energy homeostasis:neural integration of metabolic, cognitive, and emotional drives to eat,Physiol. Behav. 97 (2009) 572–580.

[423] S. Shin, N. Wakabayashi, V. Misra, S. Biswal, G.H. Lee, E.S. Agoston, M.Yamamoto, T.W. Kensler, NRF2 modulates aryl hydrocarbon receptorsignaling: influence on adipogenesis, Mol. Cell. Biol. 27 (2007) 7188–7197.

[424] U. Simanainen, T. Haavisto, J.T. Tuomisto, J. Paranko, J. Toppari, J. Tuomisto,R.E. Peterson, M. Viluksela, Pattern of male reproductive system effects afterin utero and lactational 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)exposure in three differentially TCDD-sensitive rat lines, Toxicol. Sci. 80(2004) 101–108.

[425] U. Simanainen, J.T. Tuomisto, R. Pohjanvirta, P. Syrjala, J. Tuomisto, M.Viluksela, Postnatal development of resistance to short-term high-dose toxiceffects of 2,3,7,8-tetrachlorodibenzo-p-dioxin in TCDD-resistant and -semiresistant rats, Toxicol. Appl. Pharmacol. 196 (2004) 11–19.

[426] U. Simanainen, J.T. Tuomisto, J. Tuomisto, M. Viluksela, Structure-activityrelationships and dose responses of polychlorinated dibenzo-p-dioxins forshort-term effects in 2,3,7,8-tetrachlorodibenzo-p-dioxin-resistant and -sensitive rat strains, Toxicol. Appl. Pharmacol. 181 (2002) 38–47.

[427] C.J. Sinal, J.R. Bend, Aryl hydrocarbon receptor-dependent induction ofcyp1a1 by bilirubin in mouse hepatoma hepa 1c1c7 cells, Mol. Pharmacol.52 (1997) 590–599.

[428] S. Sinkkonen, J. Paasivirta, Degradation half-life times of PCDDs, PCDFs andPCBs for environmental fate modeling, Chemosphere 40 (2000) 943–949.

[429] U. Sirkka, R. Pohjanvirta, S.A. Nieminen, J. Tuomisto, P. Ylitalo, Acuteneurobehavioural effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) inHan/Wistar rats, Pharmacol. Toxicol. 71 (1992) 284–288.

[430] K.P. Skibicka, H.J. Grill, Hypothalamic and hindbrain melanocortin receptorscontribute to the feeding, thermogenic, and cardiovascular action ofmelanocortins, Endocrinology 150 (2009) 5351–5361.

[431] A.G. Smith, B. Clothier, P. Carthew, N.L. Childs, P.R. Sinclair, D.W. Nebert, T.P.Dalton, Protection of the Cyp1a2(–/–) null mouse against uroporphyria andhepatic injury following exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin,Toxicol. Appl. Pharmacol. 173 (2001) 89–98.

[432] K. Sogawa, K. Iwabuchi, H. Abe, Y. Fujii-Kuriyama, Transcriptional activationdomains of the Ah receptor and Ah receptor nuclear translocator, J. CancerRes. Clin. Oncol. 121 (1995) 612–620.

[433] K. Sogawa, K. Numayama-Tsuruta, T. Takahashi, N. Matsushita, C. Miura, J.Nikawa, O. Gotoh, Y. Kikuchi, Y. Fujii-Kuriyama, A novel inductionmechanism of the rat CYP1A2 gene mediated by Ah receptor-Arntheterodimer, Biochem. Biophys. Res. Commun. 318 (2004) 746–755.

[434] K.M. Sojka, C.B. Kern, R.S. Pollenz, Expression and subcellular localization ofthe aryl hydrocarbon receptor nuclear translocator (ARNT) protein in mouseand chicken over developmental time, Anat. Rec. 260 (2000) 327–334.

[435] A. Solomon, B.A. De Fanti, J.A. Martinez, Peripheral ghrelin interacts withorexin neurons in glucostatic signalling, Regul. Peptides 144 (2007) 17–24.

[436] O. Sorg, M. Zennegg, P. Schmid, R. Fedosyuk, R. Valikhnovskyi, O. Gaide, V.Kniazevych, J.H. Saurat, 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)poisoning in Victor Yushchenko: identification and measurement of TCDDmetabolites, Lancet 374 (2009) 1179–1185.

[437] J.R. Speakman, Thrifty genes for obesity, an attractive but flawed idea, analternative perspective: the ‘drifty gene’ hypothesis, Int. J. Obes. (Lond) 32(2008) 1611–1617.

[438] D.C. Spink, C.L. Hayes, N.R. Young, M. Christou, T.R. Sutter, C.R. Jefcoate, J.F.Gierthy, The effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin on estrogenmetabolism in MCF-7 breast cancer cells: evidence for induction of a novel17 beta-estradiol 4-hydroxylase, J. Steroid Biochem. Mol. Biol. 51 (1994)251–258.

[439] B.U. Stahl, R.H. Alper, K. Rozman, Depletion of brain serotonin does not alter2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)-induced starvation syndrome inthe rat, Toxicol. Lett. 59 (1991) 65–72.

[440] B.R. Stanmore, The formation of dioxins in combustion systems, Combust.Flame 136 (2004) 398–427.

[441] E.A. Stevens, J.D. Mezrich, C.A. Bradfield, The aryl hydrocarbon receptor: aperspective on potential roles in the immune system, Immunology 127(2009) 299–311.

[442] J.H. Strubbe, S.C. Woods, The timing of meals, Psychol. Rev. 111 (2004) 128–141.

[443] C.H. Sutter, H. Yin, Y. Li, J.S. Mammen, S. Bodreddigari, G. Stevens, J.A. Cole,T.R. Sutter, EGF receptor signaling blocks aryl hydrocarbon receptor-mediated transcription and cell differentiation in human epidermalkeratinocytes, Proc. Natl. Acad. Sci. USA 106 (2009) 4266–4271.

[444] H.I. Swanson, DNA binding and protein interactions of the AHR/ARNTheterodimer that facilitate gene activation, Chem. Biol. Interact. 141 (2002)63–76.

[445] H.I. Swanson, W.K. Chan, C.A. Bradfield, DNA binding specificities and pairingrules of the Ah receptor, ARNT, and SIM proteins, J. Biol. Chem. 270 (1995)26292–26302.

Page 26: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478 477

[446] E. Swedenborg, I. Pongratz, AhR and ARNT modulate ER signaling, Toxicology(2009).

[447] M.H. Sweeney, P. Mocarelli, Human health effects after exposure to 2,3,7,8-TCDD, Food Addit. Contam. 17 (2000) 303–316.

[448] K. Takimoto, R. Lindahl, T.J. Dunn, H.C. Pitot, Structure of the 50 flankingregion of class 3 aldehyde dehydrogenase in the rat, Arch. Biochem. Biophys.312 (1994) 539–546.

[449] Z. Tan, X. Chang, A. Puga, Y. Xia, Activation of mitogen-activated proteinkinases (MAPKs) by aromatic hydrocarbons: role in the regulation of arylhydrocarbon receptor (AHR) function, Biochem. Pharmacol. 64 (2002) 771–780.

[450] R.T. Taylor, F. Wang, E.L. Hsu, O. Hankinson, Roles of coactivator proteins indioxin induction of CYP1A1 and CYP1B1 in human breast cancer cells,Toxicol. Sci. 107 (2009) 1–8.

[451] J.P. Thaler, S.J. Choi, M.W. Schwartz, B.E. Wisse, Hypothalamic inflammationand energy homeostasis: resolving the paradox, Front. Neuroendocrinol. 31(2010) 79–84.

[452] T.H. Thatcher, S.B. Maggirwar, C.J. Baglole, H.F. Lakatos, T.A. Gasiewicz, P.R.Phipps, P.J. Sime, Aryl hydrocarbon receptor-deficient mice developheightened inflammatory responses to cigarette smoke and endotoxinassociated with rapid loss of the nuclear factor-kappaB component RelB,Am. J. Pathol. 170 (2007) 855–864.

[453] B. Thorens, Glucose sensing and the pathogenesis of obesity and type 2diabetes, Int. J. Obes. 32 (2008) S62–S71.

[454] T.S. Thurmond, A.E. Silverstone, R.B. Baggs, F.W. Quimby, J.E. Staples, T.A.Gasiewicz, A chimeric aryl hydrocarbon receptor knockout mouse modelindicates that aryl hydrocarbon receptor activation in hematopoietic cellscontributes to the hepatic lesions induced by 2,3,7,8-tetrachlorodibenzo-p-dioxin, Toxicol. Appl. Pharmacol. 158 (1999) 33–40.

[455] Y. Tian, S. Ke, M.S. Denison, A.B. Rabson, M.A. Gallo, Ah receptor and NF-kappaB interactions, a potential mechanism for dioxin toxicity, J. Biol. Chem.274 (1999) 510–515.

[456] Y. Tian, S. Ke, T. Thomas, R.J. Meeker, M.A. Gallo, Regulation of estrogenreceptor mRNA by 2,3,7,8-tetrachlorodibenzo-p-dioxin as measured bycompetitive RT-PCR, J. Biochem. Mol. Toxicol. 12 (1998) 71–77.

[457] Y. Tian, S. Ke, T. Thomas, R.J. Meeker, M.A. Gallo, Transcriptional suppressionof estrogen receptor gene expression by 2,3,7,8-tetrachlorodibenzo-p-dioxin(TCDD), J. Steroid Biochem. Mol. Biol. 67 (1998) 17–24.

[458] Y. Tian, A.B. Rabson, M.A. Gallo, Ah receptor and NF-kappaB interactions:mechanisms and physiological implications, Chem. Biol. Interact. 141 (2002)97–115.

[459] N. Tijet, P.C. Boutros, I.D. Moffat, A.B. Okey, J. Tuomisto, R. Pohjanvirta, Arylhydrocarbon receptor regulates distinct dioxin-dependent and dioxin-independent gene batteries, Mol. Pharmacol. 69 (2006) 140–153.

[460] T. Todaka, H. Hirakawa, J. Kajiwara, T. Hori, K. Tobiishi, D. Onozuka, S. Kato, S.Sasaki, S. Nakajima, Y. Saijo, F. Sata, R. Kishi, T. Iida, M. Furue, Concentrationsof polychlorinated dibenzo-p-dioxins, polychlorinated dibenzofurans, anddioxin-like polychlorinated biphenyls in blood and breast milk collected from60 mothers in Sapporo City, Japan, Chemosphere 72 (2008) 1152–1158.

[461] M. Tohkin, M. Fukuhara, G. Elizondo, S. Tomita, F.J. Gonzalez, Arylhydrocarbon receptor is required for p300-mediated induction of DNAsynthesis by adenovirus E1A, Mol. Pharmacol. 58 (2000) 845–851.

[462] J. Tuomisto, Does mechanistic understanding help in risk assessment – theexample of dioxins, Toxicol. Appl. Pharmacol. 207 (2005) 2–10.

[463] J. Tuomisto, W. Andrzejewski, M. Unkila, R. Pohjanvirta, J. Linden, T.Vartiainen, L. Tuomisto, Modulation of TCDD-induced wasting syndrome byportocaval anastomosis and vagotomy in Long-Evans and Han/Wistar rats,Eur. J. Pharmacol. 292 (1995) 277–285.

[464] J. Tuomisto, R. Pohjanvirta, E. MacDonald, L. Tuomisto, Changes in rat brainmonoamines, monoamine metabolites and histamine after a singleadministration of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD), Pharmacol.Toxicol. 67 (1990) 260–265.

[465] J.T. Tuomisto, M. Laaksonen, M. Viluksela, R. Pohjanvirta, J. Tuomisto, Minorchanges in leptin levels after 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)exposure, Pharmacol. Toxicol. 80 (3) (1997) 113–114.

[466] J.T. Tuomisto, J. Pekkanen, H. Kiviranta, E. Tukiainen, T. Vartiainen, J.Tuomisto, Soft-tissue sarcoma and dioxin: a case-control study, Int. J.Cancer 108 (2004) 893–900.

[467] J.T. Tuomisto, R. Pohjanvirta, M. Unkila, J. Tuomisto, 2,3,7,8-Tetrachlorodibenzo-p-dioxin-induced anorexia and wasting syndrome inrats: aggravation after ventromedial hypothalamic lesion, Eur. J. Pharmacol.293 (1995) 309–317.

[468] J.T. Tuomisto, R. Pohjanvirta, M. Unkila, J. Tuomisto, TCDD-induced anorexiaand wasting syndrome in rats: effects of diet-induced obesity and nutrition,Pharmacol. Biochem. Behav. 62 (1999) 735–742.

[469] J.T. Tuomisto, R. Pohjanvirta, M. Unkila, M. Viluksela, J. Tuomisto, TCDDblocks the weight increasing effect of paraventricular lesion, OrganohalogenCompd 29 (1996) 371–374.

[470] J.T. Tuomisto, R. Pohjanvirta, M. Viluksela, J. Tuomisto, 2,3,7,8-Tetrachlorodibenzo-p-dioxin (TCDD) and lesion of dorsomedial hypothalamicnucleus have additive effects on body weight loss, Organohalogen Compd 37(1998) 81–83.

[471] J.T. Tuomisto, M. Unkila, R. Pohjanvirta, M. Koulu, L. Tuomisto, Effect of asingle dose of TCDD on the level of histamine in discrete nuclei in rat brain,Agents Actions 33 (1991) 154–156.

[472] J.T. Tuomisto, M. Viluksela, R. Pohjanvirta, J. Tuomisto, The AH receptor and anovel gene determine acute toxic responses to TCDD: segregation of theresistant alleles to different rat lines, Toxicol. Appl. Pharmacol. 155 (1999)71–81.

[473] J.T. Tuomisto, M. Viluksela, R. Pohjanvirta, J. Tuomisto, Changes in food intakeand food selection in rats after 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)exposure, Pharmacol. Biochem. Behav. 65 (2000) 381–387.

[474] C.I. Turenius, M.M. Htut, D.A. Prodon, P.L. Ebersole, P.T. Ngo, R.N. Lara, J.L.Wilczynski, B.G. Stanley, GABA(A) receptors in the lateral hypothalamus asmediators of satiety and body weight regulation, Brain Res. 1262 (2009) 16–24.

[475] A.W. Turunen, P.K. Verkasalo, H. Kiviranta, E. Pukkala, A. Jula, S. Mannisto, R.Rasanen, J. Marniemi, T. Vartiainen, Mortality in a cohort with high fishconsumption, Int. J. Epidemiol. 37 (2008) 1008–1017.

[476] M. Unkila, R. Pohjanvirta, P. Honkakoski, R. Torronen, J. Tuomisto, 2,3,7,8-Tetrachlorodibenzo-p-dioxin (TCDD) induced ethoxyresorufin-O-deethylase(EROD) and aldehyde dehydrogenase (ALDH3) activities in the brain andliver. A comparison between the most TCDD-susceptible and the most TCDD-resistant rat strain, Biochem. Pharmacol. 46 (1993) 651–659.

[477] M. Unkila, R. Pohjanvirta, E. MacDonald, J. Tuomisto, Differential effect ofTCDD on brain serotonin metabolism in a TCDD-susceptible and a TCDD-resistant rat strain, Chemosphere 27 (1993) 401–406.

[478] M. Unkila, R. Pohjanvirta, J. Tuomisto, Biochemical effects of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) and related compounds on the centralnervous system, Int. J. Biochem. Cell Biol. 27 (1995) 443–455.

[479] M. Unkila, R. Pohjanvirta, J. Tuomisto, Body weight loss and changes intryptophan homeostasis by chlorinated dibenzo-p-dioxin congeners in themost TCDD-susceptible and the most TCDD-resistant rat strain, Arch. Toxicol.72 (1998) 769–776.

[480] M. Unkila, R. Pohjanvirta, J. Tuomisto, Dioxin-induced perturbations intryptophan homeostasis in laboratory animals, Adv. Exp. Med. Biol. 467(1999) 433–442.

[481] M. Unkila, M. Ruotsalainen, R. Pohjanvirta, M. Viluksela, E. MacDonald, J.T.Tuomisto, K. Rozman, J. Tuomisto, Effect of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) on tryptophan and glucose homeostasis in the most TCDD-susceptible and the most TCDD-resistant species, guinea pigs and hamsters,Arch. Toxicol. 69 (1995) 677–683.

[482] S. Uno, T.P. Dalton, P.R. Sinclair, N. Gorman, B. Wang, A.G. Smith, M.L. Miller,H.G. Shertzer, D.W. Nebert, Cyp1a1(–/–) male mice. protection against high-dose TCDD-induced lethality and wasting syndrome, and resistance tointrahepatocyte lipid accumulation and uroporphyria, Toxicol. Appl.Pharmacol. 196 (2004) 410–421.

[483] A.P. van Birgelen, M. van den Berg, Toxicokinetics, Food Addit. Contam. 17(2000) 267–273.

[484] M. Van den Berg, L.S. Birnbaum, M. Denison, M. De Vito, W. Farland, M.Feeley, H. Fiedler, H. Hakansson, A. Hanberg, L. Haws, M. Rose, S. Safe, D.Schrenk, C. Tohyama, A. Tritscher, J. Tuomisto, M. Tysklind, N. Walker, R.E.Peterson, The 2005 World Health Organization reevaluation of human andmammalian toxic equivalency factors for dioxins and dioxin-like compounds,Toxicol. Sci. 93 (2006) 223–241.

[485] F.X.R. Van Leeuwen, R. Malisch, Results of the third round of the WHO-coordinated exposure study on the levels of PCBs, PCDDs and PCDFs inhuman milk, Organohalogen Compd 56 (2002) 311–316.

[486] M. Viluksela, Y. Bager, J.T. Tuomisto, G. Scheu, M. Unkila, R. Pohjanvirta, S.Flodstrom, V.M. Kosma, J. Maki-Paakkanen, T. Vartiainen, C. Klimm, K.W.Schramm, L. Warngard, J. Tuomisto, Liver tumor-promoting activity of2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in TCDD-sensitive and TCDD-resistant rat strains, Cancer Res. 60 (2000) 6911–6920.

[487] M. Viluksela, R. Pohjanvirta, J.T. Tuomisto, M. Unkila, J. Tuomisto, Nitric oxideantagonist N-nitro-L-arginine decreases the lethality of TCDD in mice, butincreases it in rats, Organohalogen Compd 29 (1996) 272–276.

[488] C.F. Vogel, F. Matsumura, A new cross-talk between the aryl hydrocarbonreceptor and RelB, a member of the NF-kappaB family, Biochem. Pharmacol.77 (2009) 734–745.

[489] C.F. Vogel, E. Sciullo, S. Park, C. Liedtke, C. Trautwein, F. Matsumura, Dioxinincreases C/EBPbeta transcription by activating cAMP/protein kinase A, J.Biol. Chem. 279 (2004) 8886–8894.

[490] R. Voorman, S.D. Aust, TCDD (2,3,7,8-tetrachlorodibenzo-p-dioxin) is a tightbinding inhibitor of cytochrome P-450d, J. Biochem. Toxicol. 4 (1989) 105–109.

[491] B.A. Vorderstrasse, L.B. Steppan, A.E. Silverstone, N.I. Kerkvliet, Arylhydrocarbon receptor-deficient mice generate normal immune responses tomodel antigens and are resistant to TCDD-induced immune suppression,Toxicol. Appl. Pharmacol. 171 (2001) 157–164.

[492] J.A. Walisser, M.K. Bunger, E. Glover, E.B. Harstad, C.A. Bradfield, Patentductus venosus and dioxin resistance in mice harboring a hypomorphic Arntallele, J. Biol. Chem. 279 (2004) 16326–16331.

[493] S. Wang, O. Hankinson, Functional involvement of the Brahma/SWI2-relatedgene 1 protein in cytochrome P4501A1 transcription mediated by the arylhydrocarbon receptor complex, J. Biol. Chem. 277 (2002) 11821–11827.

[494] W. Wang, E. Svanberg, D. Delbro, K. Lundholm, NOS isoenzyme content inbrain nuclei as related to food intake in experimental cancer cachexia, BrainRes. Mol. Brain Res. 134 (2005) 205–214.

[495] A.J. Watson, O. Hankinson, Dioxin- and Ah receptor-dependent proteinbinding to xenobiotic responsive elements and G-rich DNA studied by in vivofootprinting, J. Biol. Chem. 267 (1992) 6874–6878.

Page 27: Dioxins, the aryl hydrocarbon receptor and the central regulation of energy balance

478 J. Lindén et al. / Frontiers in Neuroendocrinology 31 (2010) 452–478

[496] K. Watt, T.J. Jess, S.M. Kelly, N.C. Price, I.J. McEwan, Induced alpha-helixstructure in the aryl hydrocarbon receptor transactivation domain modulatesprotein–protein interactions, Biochemistry 44 (2005) 734–743.

[497] L.W. Weber, S.W. Ernst, B.U. Stahl, K. Rozman, Tissue distribution andtoxicokinetics of 2,3,7,8-tetrachlorodibenzo-p-dioxin in rats afterintravenous injection, Fund. Appl. Toxicol. 21 (1993) 523–534.

[498] L.W. Weber, M. Lebofsky, B.U. Stahl, J.R. Gorski, G. Muzi, K. Rozman, Reducedactivities of key enzymes of gluconeogenesis as possible cause of acutetoxicity of 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) in rats, Toxicology 66(1991) 133–144.

[499] C. Weiss, D. Faust, H. Durk, S.K. Kolluri, A. Pelzer, S. Schneider, C. Dietrich, F.Oesch, M. Gottlicher, TCDD induces c-jun expression via a novel Ah (dioxin)receptor-mediated p38-MAPK-dependent pathway, Oncogene 24 (2005)4975–4983.

[500] S. Wen, F. Yang, J.G. Li, Y. Gong, X.L. Zhang, Y. Hui, Y.N. Wu, Y.F. Zhao, Y. Xu,Polychlorinated dibenzo-p-dioxin and dibenzofurans (PCDD/Fs),polybrominated diphenyl ethers (PBDEs), and polychlorinated biphenyls(PCBs) monitored by tree bark in an E-waste recycling area, Chemosphere 74(2009) 981–987.

[501] S.S. White, L.S. Birnbaum, An overview of the effects of dioxins and dioxin-like compounds on vertebrates, as documented in human and ecologicalepidemiology, J. Environ. Sci. Health. C. Environ. Carcinog. Ecotoxicol. Rev. 27(2009) 197–211.

[502] M.L. Whitelaw, J.A. Gustafsson, L. Poellinger, Identification of transactivationand repression functions of the dioxin receptor and its basic helix–loop–helix/PAS partner factor Arnt: inducible versus constitutive modes ofregulation, Mol. Cell. Biol. 14 (1994) 8343–8355.

[503] J.P. Whitlock Jr., C.H. Chichester, R.M. Bedgood, S.T. Okino, H.P. Ko, Q. Ma, L.Dong, H. Li, R. Clarke-Katzenberg, Induction of drug-metabolizing enzymesby dioxin, Drug Metab. Rev. 29 (1997) 1107–1127.

[504] A.N. Wick, D.R. Drury, H.I. Nakada, J.B. Wolfe, Localization of the primarymetabolic block produced by 2-deoxyglucose, J. Biol. Chem. 224 (1957) 963–969.

[505] M. Widerak, C. Ghoneim, M.F. Dumontier, M. Quesne, M.T. Corvol, J.F.Savouret, The aryl hydrocarbon receptor activates the retinoic acidreceptoralpha through SMRT antagonism, Biochimie 88 (2006) 387–397.

[506] J.J. Widholm, B.W. Seo, B.J. Strupp, R.F. Seegal, S.L. Schantz, Effects of perinatalexposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin on spatial and visualreversal learning in rats, Neurotoxicol. Teratol. 25 (2003) 459–471.

[507] B. Wihlen, S. Ahmed, J. Inzunza, J. Matthews, Estrogen receptor subtype- andpromoter-specific modulation of aryl hydrocarbon receptor-dependenttranscription, Mol. Cancer. Res. 7 (2009) 977–986.

[508] M.A. Williamson, T.A. Gasiewicz, L.A. Opanashuk, Aryl hydrocarbon receptorexpression and activity in cerebellar granule neuroblasts: implications fordevelopment and dioxin neurotoxicity, Toxicol. Sci. 83 (2005) 340–348.

[509] E. Wincent, N. Amini, S. Luecke, H. Glatt, J. Bergman, C. Crescenzi, A. Rannug,U. Rannug, The suggested physiologic aryl hydrocarbon receptor activatorand cytochrome P4501 substrate 6-formylindolo[3,2-b]carbazole is presentin humans, J. Biol. Chem. 284 (2009) 2690–2696.

[510] D. Wirtshafter, J.D. Davis, Set points, settling points, and the control of bodyweight, Physiol. Behav. 19 (1977) 75–78.

[511] B.E. Wisse, F. Kim, M.W. Schwartz, An integrative view of obesity, Science 318(2007) 928–929.

[512] R.M. Wittich, Degradation of dioxin-like compounds by microorganisms,Appl. Microbiol. Biotechnol. 49 (1998) 489–499.

[513] S.C. Woods, The control of food intake: behavioral versus molecularperspectives, Cell. Metab. 9 (2009) 489–498.

[514] S.C. Woods, D.A. D’Alessio, Central control of body weight and appetite, J. Clin.Endocrinol. Metab. 93 (2008) S37–50.

[515] S.C. Woods, R.J. Seeley, D. Cota, Regulation of food intake throughhypothalamic signaling networks involving mTOR, Annu. Rev. Nutr. 28(2008) 295–311.

[516] A.M. Wren, S.R. Bloom, Gut hormones and appetite control, Gastroenterology132 (2007) 2116–2130.

[517] Q. Wu, M.P. Boyle, R.D. Palmiter, Loss of GABAergic signaling by AgRPneurons to the parabrachial nucleus leads to starvation, Cell 137 (2009)1225–1234.

[518] Q. Wu, M.P. Howell, R.D. Palmiter, Ablation of neurons expressing agouti-related protein activates fos and gliosis in postsynaptic target regions, J.Neurosci. 28 (2008) 9218–9226.

[519] K. Wynne, S. Stanley, B. McGowan, S. Bloom, Appetite control, J. Endocrinol.184 (2005) 291–318.

[520] V.K. Yadav, F. Oury, N. Suda, Z.W. Liu, X.B. Gao, C. Confavreux, K.C.Klemenhagen, K.F. Tanaka, J.A. Gingrich, X.E. Guo, L.H. Tecott, J.J. Mann, R.Hen, T.L. Horvath, G. Karsenty, A serotonin-dependent mechanism explainsthe leptin regulation of bone mass, appetite, and energy expenditure, Cell 138(2009) 976–989.

[521] T. Yamada, H. Katagiri, Avenues of communication between the brain andtissues/organs involved in energy homeostasis, Endocr. J. 54 (2007) 497–505.

[522] C. Yang, F. Boucher, A. Tremblay, J.L. Michaud, Regulatory interactionbetween arylhydrocarbon receptor and SIM1, two basic helix–loop–helixPAS proteins involved in the control of food intake, J. Biol. Chem. 279 (2004)9306–9312.

[523] R.L. Yeager, S.A. Reisman, L.M. Aleksunes, C.D. Klaassen, Introducing theTCDD-inducible AhR-Nrf2 gene battery, Toxicol. Sci. 111 (2009) 238–246.

[524] H. Yoshimatsu, Hypothalamic neuronal histamine regulates body weightthrough the modulation of diurnal feeding rhythm, Nutrition 24 (2008) 827–831.

[525] M.F. Yueh, J.A. Bonzo, R.H. Tukey, The role of Ah receptor in induction ofhuman UDP-glucuronosyltransferase 1A1, Methods Enzymol. 400 (2005) 75–91.

[526] M.F. Yueh, Y.H. Huang, A. Hiller, S. Chen, N. Nguyen, R.H. Tukey, Involvementof the xenobiotic response element (XRE) in Ah receptor-mediated inductionof human UDP-glucuronosyltransferase 1A1, J. Biol. Chem. 278 (2003)15001–15006.

[527] H. Zaher, P.M. Fernandez-Salguero, J. Letterio, M.S. Sheikh, A.J. Fornace Jr., A.B.Roberts, F.J. Gonzalez, The involvement of aryl hydrocarbon receptor in theactivation of transforming growth factor-beta and apoptosis, Mol. Pharmacol.54 (1998) 313–321.

[528] S. Zhang, C. Qin, S.H. Safe, Flavonoids as aryl hydrocarbon receptor agonists/antagonists: effects of structure and cell context, Environ. Health Perspect.111 (2003) 1877–1882.

[529] S. Zhang, C. Rowlands, S. Safe, Ligand-dependent interactions of the Ahreceptor with coactivators in a mammalian two-hybrid assay, Toxicol. Appl.Pharmacol. 227 (2008) 196–206.

[530] J.G. Zhou, E.C. Henry, C.M. Palermo, S.D. Dertinger, T.A. Gasiewicz, Species-specific transcriptional activity of synthetic flavonoids in guinea pig andmouse cells as a result of differential activation of the aryl hydrocarbonreceptor to interact with dioxin-responsive elements, Mol. Pharmacol. 63(2003) 915–924.

[531] B.T. Zhu, M.A. Gallo, C.W. Burger Jr., R.J. Meeker, M.X. Cai, S. Xu, A.H. Conney,Effect of 2,3,7,8-tetrachlorodibenzo-p-dioxin administration and high-fatdiet on the body weight and hepatic estrogen metabolism in female C3H/HeNmice, Toxicol. Appl. Pharmacol. 226 (2008) 107–118.

[532] J.N. Zhu, C.L. Guo, H.Z. Li, J.J. Wang, Dorsomedial hypothalamic nucleusneurons integrate important peripheral feeding-related signals in rats, J.Neurosci. Res. 85 (2007) 3193–3204.