Review doi: 10.1111/joim.12255

Brown adipose tissue and its therapeutic potential M. E. Lidell, M. J. Betz† & S. Enerb€ack From the Department of Medical and Clinical Genetics, Institute of Biomedicine, The Sahlgrenska Academy, University of Gothenburg, Gothenburg, Sweden †Present address: Department of Internal Medicine, University Hospital Basel, Basel, Switzerland

Abstract. Lidell ME, Betz MJ, Enerb€ ack S (Institute of Biomedicine, The Sahlgrenska Academy, University of Gothenburg, Gothenburg, Sweden). Brown adipose tissue and its therapeutic potential (Review). J Intern Med 2014; 276: 364–377. Obesity and related diseases are a major cause of human morbidity and mortality and constitute a substantial economic burden for society. Effective treatment regimens are scarce, and new therapeutic targets are needed. Brown adipose tissue, an energy-expending tissue that produces heat, represents a potential therapeutic target. Its presence is associated with low body mass index, low total adipose tissue content and a lower risk of type 2 diabetes mellitus. Knowledge about the development and function of thermogenic adipocytes in brown adipose tissue has increased substantially in the last decade. Important transcriptional regulators have been identified, and hormones able to modulate the thermogenic capacity of the tissue

Introduction In 1551, Swiss naturalist Konrad Gessner first described brown adipose tissue (BAT) in the interscapular region of marmots [Muris (Marmota) alpinus]. However, it was not until the early 20th century that the existence of BAT in humans was recognized [1, 2]. At the time, the tissue was referred to as the interscapular gland in humans or the hibernating gland in smaller mammals. These early studies described the anatomical location, gross morphology and histomorphology of the highly vascularized and innervated tissue, which is largely composed of rounded or polygonal multilocular fat-filled cells that have granular cytoplasm and centrally located nuclei. BAT was initially believed to function during hibernation, but its presence in nonhibernating animals and humans suggested other uses. Although Polimanti speculated about a thermoregulatory function in 1912, it was not until 1958 that Johansson concluded from a literature review that ‘brown fat, at least in some animals, appears to be important in the regulation of body temperature’ [3, 4]. During the 1960s,

364

ª 2014 The Association for the Publication of the Journal of Internal Medicine

have been recognized. Intriguingly, it is now clear that humans, like rodents, possess two types of thermogenic adipocytes: the classical brown adipocytes found in the interscapular brown adipose organ and the so-called beige adipocytes primarily found in subcutaneous white adipose tissue after adrenergic stimulation. The presence of two distinct types of energy-expending adipocytes in humans is conceptually important because these cells might be stimulated and recruited by different signals, raising the possibility that they might be separate potential targets for therapeutic intervention. In this review, we will discuss important features of the energy-expending brown adipose tissue and highlight those that may serve as potential targets for pharmacological intervention aimed at expanding the tissue and/or enhancing its function to counteract obesity. Keywords: beige adipocytes, brown adipocytes, brown adipose tissue, energy expenditure, obesity.

several studies evaluated the tissue as a thermogenic organ [5]. Silverman et al. [6] showed that during cold exposure of human newborns, the skin temperature of human newborns fell the least over the nape of the neck, the location coinciding with BAT. Dawkins and Scopes found that the increased oxygen consumption that occurred in infants subjected to cold was associated with elevated plasma glycerol levels but no accompanying rise in plasma free fatty acids (FFAs) [7]. As BAT does not express glycerol kinase, these results indicated that lipolysis occurred locally in BAT, which used the released FFAs for heat production. Additionally, histological analysis of BAT from deceased neonates revealed that the tissue was filled with fat if the infants had been nursed at thermoneutrality, but that it was depleted from fat if they had been conventionally swaddled and nursed at room temperature (reviewed in [5]). In the 1960s and early 1970s, several studies were conducted to determine the distribution of BAT in humans. The studies focused primarily on the presence of BAT in infants in whom the tissue was abundant and widely distributed [8–10]. Heaton also presented

M. E. Lidell et al.

data, suggesting that BAT gradually disappeared with increasing age, which supported the prevailing opinion that thermogenically active BAT was common in human infants but regressed with age, leaving little or no active BAT in adults [9]. However, the view that adult humans virtually lack active BAT was challenged by unexpected findings in nuclear medicine about 20 years later. Specifically, during positron emission tomography (PET) with the tracer [18F]-fluorodeoxyglucose (FDG) for cancer staging or surveillance, a confounding symmetrical tracer uptake was often found in the neck and shoulder area of patients [11]. These areas were unrelated to the tumours, and computed tomography (CT) showed that they had features of adipose tissue. Uptake in these regions was less prevalent when patients were acclimatized to warm indoor temperatures prior to the scan, which led to the hypothesis that adult humans retain significant amounts of metabolically active BAT. Several studies were subsequently initiated to test the hypothesis, and in April 2009, three independent studies, all confirming the presence of metabolically active BAT in adult humans, were published [12–14]. Since then, many studies related to BAT in humans have been published, and it is now well accepted that most adults have metabolically active BAT [15–17]. Rodent studies suggest that activation and/or expansion of energy-expending BAT is associated with a healthy metabolic phenotype [18–21], and increasing evidence suggests a similar connection in humans. For example, several studies have shown an inverse association between the presence of BAT and obesity and type 2 diabetes mellitus [13, 15, 22, 23]. Hence, from having been an obscure gland connected with hibernation at the beginning of last century, BAT is now considered a potential target for therapy to treat obesity and obesity-related diseases. Brown adipose tissue – a thermogenic organ When rodents or humans experience cold, thermogenesis is initially mediated by shivering. During prolonged cold exposure, shivering gradually subsides, whilst nonshivering thermogenesis increases [24, 25]. Animal studies have shown that this type of thermogenesis depends on BAT [26]. Upon cold exposure, efferent signals from the hypothalamus that activate BAT thermogenesis are relayed to the tissue by sympathetic innervation (Fig. 1). Norepinephrine acts as the primary

Review: Brown adipose tissue

transmitter and activates b-adrenergic signalling pathways within the brown adipocytes [27]. The subsequent induction of lipolysis generates FFAs, the main substrate for heat production in BAT. The FFAs in turn activate uncoupling protein 1 (UCP1), the core protein of the ‘thermogenic engine’ [27, 28], which is only found in the inner mitochondrial membrane of brown adipocytes. Activated UCP1 uncouples oxidative phosphorylation from ATP regeneration by facilitating the reflux of protons across the inner mitochondrial membrane down the proton gradient, thus bypassing ATP synthase [29]. It is this ‘short circuit’ of the proton gradient that generates heat. BAT can therefore be seen as capable of transforming energy stored as triglycerides into heat. The rich vascularization of the tissue is essential for supplying the tissue with oxygen and transporting the generated heat to the rest of the body. The importance of a rich blood supply for cold-induced thermogenesis is illustrated by BAT perfusion more than doubling when human subjects are subjected to acute cold [30]. Apart from the acute effect of increased adrenergic activity, prolonged activity due to persistent cold exposure has additional consequences for BAT, including increased amounts of UCP1, increased mitochondrial biogenesis and both hyperplasia and hypertrophy of the tissue [31, 32]. Two important players in these events are peroxisome proliferator-activated receptor c (PPARc) coactivator 1a (PGC-1a) and type 2 iodothyronine deiodinase (DIO2), which are both up-regulated and activated by adrenergic signalling [33, 34]. PGC-1a, a transcriptional coactivator of the nuclear receptor PPARc, is an integral regulator of genes that are involved in mitochondrial biogenesis and oxidative metabolism [35]. In addition, PGC-1a activates transcription of the Ucp1 gene by coactivating nuclear receptors, such as PPARc and the thyroid hormone receptor (THR), that are assembled on Ucp1-regulating DNA elements [31]. The enzyme DIO2 activates THR by generating its most active ligand triiodothyronine (T3) from thyroxine (T4) locally in the brown adipocytes [34]. Hence, the thyroid and sympathetic systems act synergistically to increase the thermogenic potential of BAT by increasing its UCP1 content. The essential role of functional BAT for thermoregulation in small mammals such as rodents is undisputed and illustrated by several mouse models. Mice with reduced BAT mass due to transgenic expression of the cell toxic diphtheria ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

365

M. E. Lidell et al.

Review: Brown adipose tissue

Fig. 1 Overview of cold-induced activation of thermogenesis in brown adipose tissue (BAT). In response to cold exposure, afferent signals from cold receptors are integrated in the hypothalamus, which in turn activates BAT via the sympathetic nervous system (SNS). Norepinephrine (NE) stimulates b3-adrenergic receptors (b3) on the surface of thermogenic adipocytes. The activation of the adrenergic signalling pathway induces lipolysis and expression of thermogenic genes in the adipocytes. Free fatty acids released during lipolysis activate UCP1 and serve as the main fuel for thermogenesis. By short-circuiting the proton gradient built up over the inner mitochondrial membrane by the electron transport chain, the activated UCP1 uncouples the electromotive force from ATP synthesis by ATP synthase and heat is generated.

toxin A-chain in brown adipocytes are cold sensitive, as are Ucp1-ablated mice [36, 37]. Although multiple lines of evidence suggest that BAT is important for thermoregulation in human infants, less is known about its role in adults. However, it is clear that a large proportion of adults retain BAT that can be activated by cold [12, 17, 30]. This fact together with BAT apparently being more active during the cold months of the year strongly suggests that the tissue is also involved in thermoregulation in adults [13, 22, 38, 39]. Two types of brown adipocytes Animal studies have shown that apart from the brown adipocytes found in classical BAT depots, brown-like adipocytes occur within white adipose tissue (WAT) depots in response to chronic cold exposure or administration of b3-adrenergic agonists or thiazolidinediones (TZDs), a group of PPARc agonists [40–43]. These cells, referred to as inducible brown adipocytes, brite (brown in white) or beige adipocytes, present with a phenotype similar to that of brown adipocytes in classical BAT depots, including multilocular lipid droplets, a high mitochondrial content and expression of Ucp1 and 366

ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

Pgc1a genes. It has also been confirmed that these cells can induce energy-expending thermogenesis [44, 45]. Importantly, despite their similarities, classical brown and beige adipocytes differ in their basal expression of Ucp1. Whilst classical brown adipocytes express Ucp1 at high levels in the basal state, beige adipocytes exhibit much lower basal expression. However, stimulation with b-adrenergic agonists or activators of PPARc can induce Ucp1 expression in beige adipocytes to levels comparable to those in classical brown adipocytes [45]. Due to white and brown adipocytes sharing many features, the two cell types were until recently thought to stem from a common progenitor cell. However, using a lineage tracing approach, Seale et al. [46] showed that brown adipocytes and skeletal muscle cells, but not white adipocytes, descend from progenitor cells expressing Myf5, the gene encoding myogenic regulatory factor MYF-5. This close developmental connection had previously been suggested by Atit et al. [47] who demonstrated that dorsal dermis, epaxial skeletal muscle and interscapular BAT (iBAT) derive from

M. E. Lidell et al.

Review: Brown adipose tissue

En1-expressing cells of the central dermomyotome. Most importantly, it became apparent that brown and beige adipocytes were developmentally different cell types because, like white adipocytes, the beige adipocytes within the WAT of mice treated with a selective b3-adrenergic agonist did not derive from Myf5-expressing progenitors [46]. Discrete and unique gene signatures of brown and beige adipocytes provide further evidence that the two entities are distinct cell types [45, 48].

The presence of two distinct types of thermogenically competent and energy-expending adipocytes in humans is conceptually important because they might be stimulated and recruited by different signals; they therefore represent separate potential targets for therapeutic intervention. Additionally, the finding in mice that genetic variability affects the development of beige adipocytes, but not classical brown adipocytes support differential regulation of the two cell types [54].

Recent studies have contributed important information about the development of beige adipocytes. Utilizing genetic lineage tracing techniques, Lee et al. [49] identified bipotent precursor cells expressing platelet-derived growth factor receptor (PDGFR)a, Sca-1 and CD34 in close proximity to blood vessels in WAT. When stimulated by a b3adrenergic agonist, these cells developed into beige adipocytes. If the animals were fed a high-fat diet, however, the cells developed into white adipocytes.

Important players in the development and function of BAT

The supraclavicular BAT depot of adult humans was recently found to express a gene signature resembling that of beige adipocytes, raising the question of whether humans possess classical brown adipocytes at all [45, 48, 50]. The old indication that human infants might have an iBAT depot, a depot known to contain classical brown adipocytes in rodents, led us to test the hypothesis that at least human infants, like other small mammals, have an anatomically distinguishable iBAT depot consisting of classical brown adipocytes[8, 51]. Postmortem magnetic resonance imaging of human infants indeed revealed tissue in the interscapular region with an intermediate-fat fraction, as expected for BAT. The histomorphology of the sampled tissue closely resembled that of iBAT depots of rodents, presenting with densely packed multilocular and UCP1-positive cells delineated from subcutaneous WAT by a layer of connective tissue. Gene expression analysis using previously described marker genes for classical brown and beige adipocytes revealed a gene signature more similar to that expected for classical brown adipocytes. Hence, it appears that at least human infants, like rodents, possess bona fide iBAT consisting of classical brown adipocytes, which suggests that humans actually have both types of thermogenic adipocytes. Additionally, two independent studies characterizing BAT from the neck region of adults support our findings and indicate that classical brown adipocytes are not restricted to infants but also exist in adults [52, 53].

PR domain zinc finger protein 16 Prdm16, the gene encoding PR domain zinc finger protein 16 (PRDM16), is one of a few murine transcriptional components preferentially expressed in brown versus white adipocytes [55]. Overexpression of PRDM16 in fibroblasts or white adipocyte precursors induces a full brown adipocyte gene programme and stimulates both mitochondrial biogenesis and uncoupled cellular respiration. In contrast, a reduction of cellular PRDM16 levels in brown preadipocytes causes a corresponding decrease in the expression of brown adipocyte-associated genes. In line with this, transgenic overexpression of Prdm16 in WAT of mice increases the expression of brown adipocyteassociated genes and boosts the number of brownlike adipocytes (presumably beige adipocytes) in epididymal WAT (a classical WAT depot) after b3adrenergic stimulation. In both in vitro and in vivo experiments, the induction of the brown adipocyte phenotype is accompanied by reduced expression of white adipocyte-associated markers. This dual function of PRDM16 as an inducer of brown adipocyte-associated genes and a repressor of white adipocyte-associated genes is mediated through interaction with PGC1a on promoters of brown adipocyte-associated genes and with corepressing C-terminal-binding proteins 1 and 2 (CtBP-1 and CtBP-2) on promoters of white adipocyte-associated genes, respectively [56]. Although preferentially expressed in BAT as compared with WAT, Prdm16 was recently shown to also be selectively expressed in subcutaneous WAT depots in contrast to intra-abdominal WAT depots [57]. In addition, inguinal subcutaneous WAT of mice overexpressing Prdm16 in adipose tissue presented with a BAT-like phenotype even without b3-adrenergic agonist treatment. However, this alteration was not seen either in the epidydimal WAT depot of the transgenic mice or in the inguinal ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

367

M. E. Lidell et al.

depot of wild-type littermates. The importance of Prdm16 for the b3-adrenergic-mediated appearance of presumed beige adipocytes in WAT was further highlighted by Prdm16 heterozygous mice developing many fewer clusters of brown-like adipocytes in their inguinal WAT compared to wild-type littermates when subjected to a b3adrenergic agonist. PRDM16 is also required for the browning of WAT in response to TZDs, and it has been suggested that the effects of TZDs on browning are due in large measure to the stabilization and accumulation of PRDM16 [58].

Review: Brown adipose tissue

Despite its importance for adaptive thermogenesis, PGC1a does not seem to be a master regulator of BAT formation per se as both iBAT and subcutaneous WAT depots normally containing clusters of beige adipocytes, present with a seemingly normal morphology in PGC1a-deficient mice [62]. The differentiation of brown preadipocytes lacking PGC1a also seems to be normal as they express brown adipocyte marker genes and develop into cells with a characteristic brown adipocyte morphology [61]. Hence, PGC1a seems to be an essential regulator of adaptive thermogenesis but not an indispensable factor for determination of brown or beige adipocytes.

Apart from its suggested role in beige adipocyte formation, PRDM16 acts as the molecular switch for directing Myf5-positive progenitor cells into the brown adipocyte lineage, whilst repressing their development into skeletal muscle cells [46]. Interactions between PRDM16 and both PPARc and the transcription factor C/EBPb are important in controlling this cell fate switch [46, 59].

PRDM16 and PGC1a – players at the crossroad of BAT formation and function

From the preceding discussion, it is clear that PRDM16 plays a vital role in the formation of both classical brown and beige adipocytes. Hence, this protein, often referred to as a master regulator of brown adipocytes, constitutes a very interesting target for pharmacologic intervention aimed at increasing the presence of thermogenically competent adipocytes.

Euchromatic histone-lysine N-methyltransferase (EHMT1), a protein selectively expressed in brown adipocytes, was recently shown to be an integral component of the machinery that induces brown adipocyte cell fate [63]. Deletion of Ehmt1 in brown adipocytes leads to the loss of BAT characteristics accompanied by the induction of a skeletal muscle gene programme. In addition, mice with adipose tissue-specific deletion of Ehmt1 are cold sensitive due to a marked reduction in BAT-mediated adaptive thermogenesis and fail to develop beige adipocytes in WAT depots in response to b3-adrenergic stimulation. In contrast, overexpression of EHMT1 in brown adipocytes increases expression of thermogenic genes and raises the oxygen consumption rate. EHMT1 controls brown adipocyte cell fate and BAT thermogenesis by interacting with and stabilizing PRDM16.

Pgc1a PGC1a was first identified as a protein that interacted with and coactivated PPARc and THR in brown adipocytes [33]. Its expression was shown to be dramatically increased in both BAT of coldexposed mice and in adipocytes subjected to a b-adrenergic agonist. Forced expression of the protein in white adipocytes not only increased the expression of Ucp1 and genes encoding proteins of the respiratory chain but also the mitochondrial DNA content of the cells, suggesting increased mitochondriogenesis. Hence, PGC1a appears to be a factor that could induce a full thermogenic programme in response to adrenergic signalling triggered by a cold environment. In support of this view, PGC1a-deficient mice have a severely blunted capacity for cold-induced thermogenesis [60]. In addition, PGC1a-deficient brown adipocytes fail to induce expression of thermogenic genes in response to adrenergic signalling [61]. 368

ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

Several transcriptional regulators affect BAT formation and function, and the effects of many of them are, at least in part, exerted through modulation of the cellular content or function of either PRDM16 or PGC1a.

The NAD-dependent protein deacetylase sirtuin-1 (SIRT1) is another protein with the capacity to modulate adaptive thermogenesis. Increased activity of SIRT1 can mimic the action of TZDs and induce browning of subcutaneous WAT depots by facilitating the interaction between PPARc and PRDM16 by deacetylating PPARc [64]. Several transcriptional regulators have been demonstrated to affect the development and function of thermogenic adipocytes, at least in part, by influ-

M. E. Lidell et al.

encing the activity or cellular levels of PGC1a. One such protein is nuclear receptor-interacting protein 1 (NRIP1, often referred to as RIP140), which binds directly to PGC1a and inhibits its activity [65]. In agreement with this, NRIP1-deficient mice attain clusters of beige adipocytes in their WAT depots, and NRIP1-null cells express high levels of UCP1 and have elevated energy expenditure [66, 67]. Steroid receptor coactivators (SRCs) belonging to the p160 family have also been shown to affect BAT function by modulating the activity of PGC1a. Whilst the expression levels of thermogenic genes are reduced and adaptive thermogenesis is impaired in Src1-deficient mice, Src2-deficient mice display increased expression levels of thermogenic genes and have increased capacity for adaptive thermogenesis [68]. Picard et al. [68] showed that SRC-1 stabilizes the interaction between PGC1a and PPARc, whereas SRC-2 inhibits this interaction by competing with SRC-1 by forming a less active complex with PGC1a. SRC-3 has also been shown to inhibit the activity of PGC1a [69]. SRC-3 appears to mediate its effects by increasing the expression of Kat2a (also called Gcn5), the gene encoding histone acetyltransferase KAT2A. This enzyme can acetylate PGC1a and thereby inhibit its activity [70]. Hence, Src3-deficient mice have increased expression of thermogenic genes and increased energy expenditure [69]. Identification of proteins that influence the activity or cellular levels of PRDM16 and PGC1a is of great importance because such factors represent potential targets for pharmacological intervention aimed at expanding BAT and/or enhancing its function. The importance of BAT as a metabolic regulator Given BAT’s capability to dissipate chemical energy as heat, it is not hard to envision BAT playing a major role in the regulation of metabolism. In 1979, Rothwell and Stock discovered that overfeeding rats with a palatable diet induced increased thermogenesis in BAT [71]. This discovery led to the concept of diet-induced thermogenesis, meaning that activation of BAT in response to caloric excess could reduce metabolic efficiency and avoid or diminish obesity. Consistent with this view, early studies using surgical denervation or transgenic ablation of BAT by overexpression of diphtheria toxin A-chain specifically in brown adipocytes demonstrated increased body weight

Review: Brown adipose tissue

and insulin resistance in treated mice [37, 72]. As previously mentioned, ablation of the Ucp1 gene in mice leads to reduced cold-induced thermogenesis. Surprisingly, such mice did not become obese when housed at room temperature [36]. However, they did become obese when housed at thermoneutral temperature (29 °C), which suggests an important metabolic role of BAT [73]. Although the idea that BAT, in addition to being a source of heat, plays a role in an innate defence against obesity is controversial, it is clear that an increased number or activity of thermogenic cells in the form of brown or beige adipocytes can counteract obesity and insulin resistance. This is illustrated in several mouse models in which the number of such cells has been artificially increased by genetic manipulation of BAT-regulating genes. As an example, increasing the number of beige adipocytes in WAT by overexpressing UCP1 in the adipose tissue of a mouse strain genetically prone to obesity normalized the phenotype of the animals [74]. In line with this, adipose tissue-specific overexpression of FOXC2, a transcription factor thought to sensitize adrenergic signalling, increased the amount of beige adipocytes in WAT depots and protected mice against diet-induced obesity, hypertriglyceridaemia and insulin resistance [19, 75]. Importantly, the capability of rodents to recruit beige adipocytes within WAT depots upon cold exposure or adrenergic stimulation is subject to genetic variation and varies greatly between different mouse strains; the capacity is lowest in strains prone to obesity and insulin resistance. However, the iBAT depot does not show this genetic variation in rodents [41, 76, 77]. Table 1 gives an overview of the metabolic implications of BAT in rodents. Upon cold exposure, BAT is activated within minutes and its energy demands are met by rapid lipolysis of the intracellular lipid stores [78]. To replenish the intracellular triglycerides, BAT takes up FFAs released from the lipids of triglyceriderich lipoproteins (TRL) by lipoprotein lipase in the endothelium of its dense vasculature. BAT activation through short-term cold exposure was recently shown to increase TRL metabolism by controlling vascular lipoprotein homeostasis in mice and ameliorating hyperlipidaemia [79]. Recently, metformin was demonstrated to reduce plasma cholesterol and lipid levels in an animal model of human lipoprotein metabolism by increasing BAT activity, leading to increased triglyceride and VLDL uptake through BAT [80]. ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

369

M. E. Lidell et al.

Review: Brown adipose tissue

Table 1 Importance of brown adipose tissue for metabolism in rodents References Ablation of BAT causes obesity and insulin resistance.

[37, 72]

Thyroid hormones increase BAT differentiation and activity and are locally

[87–89]

regulated by type 2 iodothyronine deiodinase. Thermogenic adipocytes are present as classical brown adipocytes in iBAT and

[40–45]

as beige adipocytes in WAT after adrenergic stimulation. The ability to recruit beige adipocytes in WAT is associated with decreased obesity.

[41, 76, 77]

Expansion of BAT in transgenic mouse models leads to resistance to obesity and hypertriglyceridaemia.

[19, 75]

BAT, brown adipose tissue; WAT, white adipose tissue; iBAT, interscapular BAT.

resistance upon physiological or pharmacological stimulation of BAT [81–83]. Importantly, several observational studies in humans indicate an association of increased BAT activity and insulin sensitivity as well as reduced obesity [13, 15, 17, 22, 23]. Table 2 summarizes relevant findings on BAT in humans. Taken together, several lines of evidence demonstrate major effects of BAT on metabolism both in animals and in humans. Moreover, BAT could be a potential target for therapeutic intervention to treat metabolic disease in humans.

Fig. 2 Cold-induced glucose uptake in supraclavicular brown adipose tissue. Computed tomography (CT) and positron emission tomography (PET) images of a subject exposed to cold (PET cold) and room temperature (PET warm), respectively. Two hours prior to the ‘PET cold scans’, the subject was exposed to 17–19 °C temperature whilst wearing light clothing. During imaging one of the subject’s feet was occasionally placed in ice water. PET was performed with the tracer [18F]-fluorodeoxyglucose. The location of the supraclavicular BAT depots are indicated with arrows. (From Virtanen KA, Lidell ME, Orava J, Heglind M, Westergren R, Niemi T, Taittonen M, Laine J, Savisto NJ, Enerb€ a ck S, Nuutila P. Functional brown adipose tissue in healthy adults. New England Journal of Medicine 360:1520. Copyright © 2009 Massachusetts Medical Society. Reprinted with permission.)

Table 2 Relevance of brown adipose tissue in humans References BAT is present in a majority of

[12–17]

adult humans. BAT activity is increased in

[91]

hyperthyroidism. BAT activity can be increased by

[119]

repeated mild cold exposure. BAT activity is inversely associated

[13, 15, 17, 22, 23]

with obesity and type 2 diabetes. Severe obesity is associated with

[110, 111]

less BAT.

Similar to the effects in exercising muscle, BAT activation leads to dramatically increased glucose uptake by cells in humans (Fig. 2) [30]. Although evidence for a potential therapeutic effect in humans is still missing, data from rodents provide robust indications for amelioration of insulin 370

ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

BAT in adult humans is

[45, 48, 50–53]

predominantly of the ‘beige’ type. Outdoor temperature is inversely associated with BAT activity. BAT, brown adipose tissue.

[13, 22, 38, 39]

M. E. Lidell et al.

Endocrine regulation of BAT Several hormones influence both the activity and expansion of BAT. Physiologically, increased activity of the sympathetic nervous system transmitted by norepinephrine acting primarily on b3-adrenoreceptors activates BAT [27]. Accordingly, the role of catecholamines in adrenergic stimulation of BAT activation is well established. Hypersecretion of catecholamines from pheochromocytoma in humans has been noted to lead to the transformation of WAT depots to a BAT phenotype, and an inverse association of abdominal obesity and plasma catecholamine levels in patients with pheochromocytoma has been described [84–86]. Thyroid hormones play a pivotal role in the proper function of BAT by modulating adrenergic signalling, and it has long been known that hyperthyroidism increases resting metabolic rate. The hormones are predominantly secreted as T4, which is converted into the biologically more potent T3 in peripheral tissues by deiodinases. The expression of DIO2 is a characteristic feature of brown adipocytes. Upon noradrenergic stimulation, DIO2 expression in BAT increases dramatically, leading to a high local availability of T3 [87]. Stimulation of THRa1 leads to increased sensitivity of the tissue towards catecholamines, whilst THRb is crucial for expression of UCP1 in BAT [88, 89]. Additionally, expression of PGC1a, the master regulator of BAT function, is increased by T3 via a thyroid hormone response element in the Pgc1a promoter [90]. Although these mechanisms are known from animal models, the effect of hyperthyroidism on human BAT was only recently demonstrated by FDG-PET/CT and indirect calorimetry [91]. The important interplay between thyroid hormones and catecholamines is further modulated by bile acids that increase the expression of DIO2 by activating the G-protein-coupled bile acid receptor TGR5 [92]. Specific activation of TGR5 also leads to increased insulin sensitivity in mice [93]. In recent years, effects of several peptides on BAT function and development have been described, mostly in animal models. Bone morphogenetic proteins (BMPs) are a family of growth factors that were first described in the context of bone and cartilage development. However, they also play a major role in the morphogenesis of other tissues, and BMPs have recently been implicated in the development and growth of BAT. BMP7 activates brown adipogenesis by inducing early regulators such as PRDM16 and PGC1a. Adenovirus-mediated

Review: Brown adipose tissue

overexpression of this cytokine leads to a significant increase in BAT mass and blunts weight gain [94]. Subcutaneous infusion of BMP7 increases the amount of beige adipocytes in WAT depots and reduces obesity in mice kept at temperatures below thermoneutrality, indicating that the cold stimulus is still important for activation of expanded BAT depots [95]. Another member of the BMP family, BMP8B, regulates BAT activity on both a central and a peripheral level [96]. Even though the BAT morphology of BMP8B-deficient mice appears to be normal, the thermogenic function of the tissue is impaired. Central administration of BMP8B increases sympathetic outflow towards BAT, whilst direct treatment of brown adipocytes with the peptide increases their lipolytic capacity in response to norepinephrine [96]. Administration of fibroblast growth factor 21 (FGF21) increases body temperature and BAT thermogenesis in newborn mice and induces expression of thermogenic genes in vitro [97]. Additionally, FGF21 seems to be secreted by BAT in an auto- or paracrine fashion in response to adrenergic activation upon cold exposure [98, 99]. Recently, however, the BAT specificity of FGF21stimulated thermogenesis has been questioned [100]. Recent evidence suggests that natriuretic peptides (NPs) have the capacity to induce browning of WAT depots [101]. The levels of these peptides in plasma increase dramatically in response to cold exposure, and infusion of NPs leads to increased energy expenditure and expression of thermogenic genes in subcutaneous WAT in mice. Mechanistically the effects of NPs are mediated by the NP receptors via the cGMP second messenger pathway and protein kinase G (PKG). The actions of PKG overlap with those of the catecholamine-stimulated protein kinase A (PKA): activation of hormone-sensitive lipase and perilipin induces lipolysis and phosphorylation of p38 MAPK, which in turn triggers phosphorylation of PGC1a, finally leading to transcription of Ucp1 [101]. The signalling pathways of NPs and catecholamines thus seem to converge and act synergistically to activate BAT. Whilst this observation has to be corroborated in humans, it might provide a possible link between heart failure and related cachexia. Recently, Bostr€ om et al. [18] described irisin, a novel peptide from muscle that induced a brown adipocyte phenotype when applied to subcutaneous ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

371

M. E. Lidell et al.

Review: Brown adipose tissue

white adipocytes in culture. Adenovirus-mediated overexpression of irisin in mice led to a beige phenotype in subcutaneous adipose tissue depots. Additionally, increased plasma levels of irisin were detected in humans after muscular exercise. Using the recombinant peptide, others were initially not able to reproduce these effects in vitro [102]. However, a recent study by Zhang et al. [103] convincingly showed that recombinant irisin induces browning of adipocytes isolated from subcutaneous WAT of rats. Importantly, the recombinant peptide also induced browning of subcutaneous fat pads in vivo, and protected mice against weight gain and insulin resistance induced by a high-fat diet. It has been questioned whether the beneficial effects of irisin observed in mice can be translated to humans or not as a mutation (ATG ? ATA) present in the start codon of the human FNDC5 gene, that encodes the precursor of irisin, greatly reduces the translation efficiency into full-length peptide [102]. However, Lee et al. [104] recently confirmed circulating irisin in humans using a mass spectrometry approach. The authors also confirmed that exercise increases the serum levels of irisin. In addition, they demonstrated that cold exposure of healthy humans increased circulating irisin levels and that the secretion of the peptide correlated with shivering intensity. Irisin secretion might thus not necessarily be linked to muscular exercise per se but to repetitive small movements as in shivering. Importantly, Lee et al. [104] also showed that the recombinant peptide induced browning and thermogenesis in human neck adipocytes in vitro. Hence, irisin appears to link shivering thermogenesis to an increase in nonshivering thermogenesis.

The major endocrine factors influencing BAT are listed in Table 3. Increasing BAT energy expenditure as a therapeutic means Enhancing thermogenesis to facilitate weight reduction is not an entirely new concept. During the 1930s 2,4-dinitrophenol (DNP), an artificial uncoupler of mitochondrial respiration, was widely popular as a drug for inducing weight loss. This small lipophilic molecule acts as a shuttle for protons across the inner mitochondrial membrane, thereby short-circuiting the respiratory chain in a manner similar to UCP1, albeit in an unregulated fashion [105]. Compared with UCP1, DNP does not selectively act on brown adipocyte mitochondria, and although it was effective in inducing weight loss, it was withdrawn from the market because of a narrow therapeutic range and severe side effects [106]. Prior to the discovery of metabolically active BAT in human adults, several studies evaluated the use of b3-adrenergic agonists to facilitate weight loss in humans. Although the selective b3-adrenergic agonist L-796568 acutely increased energy expenditure in obese men, it failed to induce weight loss or long-term increased energy expenditure [107, 108]. Administration of another b3-adrenergic agonist, CL 316,243, initially led to increased insulin sensitivity and fat oxidation, but it also failed to alter energy expenditure after 8 weeks of use; this was related to an unexpected decline in the plasma levels of the drug, indicating that the metabolism of the drug might have accelerated over time [109]. However, several other factors may explain why long-term treatment with b3-adrenergic agonists

Table 3 Key endocrine regulators of BAT function and development Norepinephrine

Primary neurotransmitter responsible for acute activation of BAT as well as mitochondrial biogenesis, increased UCP1 expression and tissue hyperplasia.

Thyroid hormones

Increase local sensitivity of brown adipocytes towards catecholamines and expression of UCP1 and PGC1a; local availability of T3 increased by expression of DIO2 in response to adrenergic stimulation.

BMP7/BMP8B Natriuretic peptides

These hormones have recently been shown to enhance the induction of beige adipocytes in WAT depots.

FGF21 Irisin BAT, brown adipose tissue; UCP1, uncoupling protein 1; PGC1a, peroxisome proliferator-activated receptor c coactivator 1a; T3, triiodothyronine; DIO2, type 2 iodothyronine deiodinase; BMP, bone morphogenetic protein; FGF21, fibroblast growth factor 21. 372

ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

M. E. Lidell et al.

has been unsuccessful. Chronic stimulation of adrenergic receptors commonly leads to downregulation of the target receptors over the course of a few days, a phenomenon that is referred to as tachyphylaxis, which is well known in clinical application of catecholamine derivatives. Stimulation of BAT by exogenous catecholamine analogues might also reduce the endogenous activation of the tissue. Furthermore, the prevalence and activity of targetable BAT appears to be lower in obese persons [110, 111]. It should be pointed out that the human studies were carried out before the presence of thermogenically active BAT in adult humans was demonstrated and the tissue’s activity was thus not measured. Given the increased knowledge about BAT in humans and the much improved methods of its assessment, it is still worthwhile to pursue developing new b3-adrenergic agonists with enhanced specificity, oral bioavailability and more constant plasma half-life over time, as such drugs very well could give the desired effects in humans. Stimulation of BAT via the sympathetic nervous system is thought to contribute to the weight loss effects from appetite-reducing drugs such as sibutramine or ephedrine [112–114]. These centrally acting drugs can cause serious side effects such as cardiac arrhythmias or hypertension, however, and have therefore been withdrawn from the market. A more selective way to increase BAT thermogenesis might be to target local conversion of T4 to T3 by stimulation of the G-protein-coupled bile acid receptor TGR5, thereby inducing expression of DIO2, or direct targeting of THRb with selective agonists [115–117]. Pharmacologic interventions that activate and expand BAT would provide a very attractive means for weight reduction, especially in individuals unable to exercise. However, physiological interventions such as intermittent cold stimuli or a reduction of ambient temperature could also provide a safe and affordable alternative, analogous to increasing energy expenditure through muscular exercise. In this respect, it is important to notice that human BAT is activated by mild cold stimuli such as ambient temperatures around 16 °C or by placing a hand in cold water at around 20 °C [14, 118]. Accordingly, recent preclinical research demonstrated an increase in BAT activity and cold-induced thermogenesis as well as a reduced amount of body fat after 6 weeks of repeated short-term mild cold

Review: Brown adipose tissue

exposure (17 °C for 2 h) in healthy men with normal weight [119]. Whether these effects translate to reduced obesity or increased insulin sensitivity needs to be evaluated in future studies. Conclusion From the literature, it is clear that expansion or increased activity of BAT in rodents is associated with a metabolically healthy phenotype. Despite this intriguing association in rodents, BAT was until recently not seen as a potential target for antiobesity drugs in humans because adults, the primary age group that would use such drugs, were not believed to possess significant amounts of metabolically active BAT. Hence, BAT was perceived as a conceptually interesting target without promise for therapeutic use. With the discovery of metabolically active BAT in adults in 2009, however, the tissue has become the subject of intense research. Over the last 5 years, increasing numbers of BATrelated studies have been published that confirm that the association between BAT and a metabolically healthy phenotype also holds true for humans. During the last decade, our knowledge about factors influencing BAT recruitment and function in rodents has increased substantially. Hopefully, this new knowledge can be extrapolated to humans and provide a foundation for studies that will identify and evaluate potential molecular drug targets in the coming years. Acknowledgements The work was supported by grants from the Swedish Research Council (2012-1652 and 2010-3281), the Knut and Alice Wallenberg Foundation, the Sahlgrenska University Hospital (LUA-ALF), the European Union (HEALTH-F2-2011-278373; DIABAT), the IngaBritt and Arne Lundgren Foundation, the S€ oderberg Foundation and the King Gustaf V and Queen Victoria Freemason Foundation. Conflict of interest statement Sven Enerb€ ack is shareholder and consultant to Ember Therapeutics.

References 1 Bonnot E. The interscapular gland. J Anat Physiol 1908; 43: 43–58. ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

373

M. E. Lidell et al.

2 Hatai S. On the presence in human embryos of an interscapular gland corresponding to the so-called hibernating gland in lower animals. Anat Anz 1902; 21: 369–73. 3 Johansson B. Brown fat: a review. Metabolism 1959; 8: 221–40. 4 Polimanti O. Il letargo. Roma: Tipografia del Senato di G. Bardi, 1912. 5 Smith RE, Horwitz BA. Brown fat and thermogenesis. Physiol Rev 1969; 49: 330–425. 6 Silverman WA, Zamelis A, Sinclair JC, Agate FJ. Warm nape of the newborn. Pediatrics 1964; 33: 984–7. 7 Dawkins MJ, Scopes JW. Non-shivering thermogenesis and brown adipose tissue in the human new-born infant. Nature 1965; 206: 201–2. 8 Aherne W, Hull D. Brown adipose tissue and heat production in the newborn infant. J Pathol Bacteriol 1966; 91: 223–34. 9 Heaton JM. The distribution of brown adipose tissue in the human. J Anat 1972; 112: 35–9. 10 Merklin RJ. Growth and distribution of human fetal brown fat. Anat Rec 1974; 178: 637–45. 11 Nedergaard J, Bengtsson T, Cannon B. Unexpected evidence for active brown adipose tissue in adult humans. Am J Physiol Endocrinol Metab 2007; 293: E444–52. 12 Virtanen KA, Lidell ME, Orava J et al. Functional brown adipose tissue in healthy adults. N Engl J Med 2009; 360: 1518–25. 13 Cypess AM, Lehman S, Williams G et al. Identification and importance of brown adipose tissue in adult humans. N Engl J Med 2009; 360: 1509–17. 14 van Marken Lichtenbelt WD, Vanhommerig JW, Smulders NM et al. Cold-activated brown adipose tissue in healthy men. N Engl J Med 2009; 360: 1500–8. 15 Lee P, Greenfield JR, Ho KK, Fulham MJ. A critical appraisal of the prevalence and metabolic significance of brown adipose tissue in adult humans. Am J Physiol Endocrinol Metab 2010; 299: E601–6. 16 Lee P, Zhao JT, Swarbrick MM et al. High prevalence of brown adipose tissue in adult humans. J Clin Endocrinol Metab 2011; 96: 2450–5. 17 Saito M, Okamatsu-Ogura Y, Matsushita M et al. High incidence of metabolically active brown adipose tissue in healthy adult humans: effects of cold exposure and adiposity. Diabetes 2009; 58: 1526–31. 18 Bostrom P, Wu J, Jedrychowski MP et al. A PGC1-alpha-dependent myokine that drives brown-fat-like development of white fat and thermogenesis. Nature 2012; 481: 463–8. 19 Cederberg A, Gronning LM, Ahren B, Tasken K, Carlsson P, Enerback S. FOXC2 is a winged helix gene that counteracts obesity, hypertriglyceridemia, and diet-induced insulin resistance. Cell 2001; 106: 563–73. 20 Kiefer FW, Vernochet C, O’Brien P et al. Retinaldehyde dehydrogenase 1 regulates a thermogenic program in white adipose tissue. Nat Med 2012; 18: 918–25. 21 Zhou Z, Yon Toh S, Chen Z et al. Cidea-deficient mice have lean phenotype and are resistant to obesity. Nat Genet 2003; 35: 49–56. 22 Ouellet V, Routhier-Labadie A, Bellemare W et al. Outdoor temperature, age, sex, body mass index, and diabetic status determine the prevalence, mass, and glucose-uptake activity

374

ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

Review: Brown adipose tissue

23

24 25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

of 18F-FDG-detected BAT in humans. J Clin Endocrinol Metab 2011; 96: 192–9. Pfannenberg C, Werner MK, Ripkens S et al. Impact of age on the relationships of brown adipose tissue with sex and adiposity in humans. Diabetes 2010; 59: 1789–93. Davis TR. Chamber cold acclimatization in man. J Appl Physiol 1961; 16: 1011–5. Davis TR, Johnston DR, Bell FC, Cremer BJ. Regulation of shivering and non-shivering heat production during acclimation of rats. Am J Physiol 1960; 198: 471–5. Golozoubova V, Hohtola E, Matthias A, Jacobsson A, Cannon B, Nedergaard J. Only UCP1 can mediate adaptive nonshivering thermogenesis in the cold. FASEB J 2001; 15: 2048–50. Cannon B, Nedergaard J. Brown adipose tissue: function and physiological significance. Physiol Rev 2004; 84: 277– 359. Fedorenko A, Lishko PV, Kirichok Y. Mechanism of fatty-acid-dependent UCP1 uncoupling in brown fat mitochondria. Cell 2012; 151: 400–13. Krauss S, Zhang CY, Lowell BB. The mitochondrial uncoupling-protein homologues. Nat Rev Mol Cell Biol 2005; 6: 248–61. Orava J, Nuutila P, Lidell ME et al. Different metabolic responses of human brown adipose tissue to activation by cold and insulin. Cell Metab 2011; 14: 272–9. Lowell BB, Spiegelman BM. Towards a molecular understanding of adaptive thermogenesis. Nature 2000; 404: 652–60. Bukowiecki L, Collet AJ, Follea N, Guay G, Jahjah L. Brown adipose tissue hyperplasia: a fundamental mechanism of adaptation to cold and hyperphagia. Am J Physiol 1982; 242: E353–9. Puigserver P, Wu Z, Park CW, Graves R, Wright M, Spiegelman BM. A cold-inducible coactivator of nuclear receptors linked to adaptive thermogenesis. Cell 1998; 92: 829–39. Arrojo EDR, Fonseca TL, Werneck-de-Castro JP, Bianco AC. Role of the type 2 iodothyronine deiodinase (D2) in the control of thyroid hormone signaling. Biochim Biophys Acta 2013; 1830: 3956–64. Wu Z, Puigserver P, Andersson U et al. Mechanisms controlling mitochondrial biogenesis and respiration through the thermogenic coactivator PGC-1. Cell 1999; 98: 115–24. Enerback S, Jacobsson A, Simpson EM et al. Mice lacking mitochondrial uncoupling protein are cold-sensitive but not obese. Nature 1997; 387: 90–4. Lowell BB, V SS, Hamann A et al. Development of obesity in transgenic mice after genetic ablation of brown adipose tissue. Nature 1993; 366: 740–2. Au-Yong IT, Thorn N, Ganatra R, Perkins AC, Symonds ME. Brown adipose tissue and seasonal variation in humans. Diabetes 2009; 58: 2583–7. Betz MJ, Slawik M, Lidell ME et al. Presence of brown adipocytes in retroperitoneal fat from patients with benign adrenal tumors: relationship with outdoor temperature. J Clin Endocrinol Metab 2013; 98: 4097–104. Cousin B, Cinti S, Morroni M et al. Occurrence of brown adipocytes in rat white adipose tissue: molecular and morphological characterization. J Cell Sci 1992; 103: 931– 42. Guerra C, Koza RA, Yamashita H, Walsh K, Kozak LP. Emergence of brown adipocytes in white fat in mice is under

M. E. Lidell et al.

42

43

44

45

46 47

48

49

50 51

52

53

54

55

56

57

58

genetic control. Effects on body weight and adiposity. J Clin Invest 1998; 102: 412–20. Himms-Hagen J, Melnyk A, Zingaretti MC, Ceresi E, Barbatelli G, Cinti S. Multilocular fat cells in WAT of CL-316243-treated rats derive directly from white adipocytes. Am J Physiol Cell Physiol 2000; 279: C670–81. Wilson-Fritch L, Nicoloro S, Chouinard M et al. Mitochondrial remodeling in adipose tissue associated with obesity and treatment with rosiglitazone. J Clin Invest 2004; 114: 1281–9. Petrovic N, Walden TB, Shabalina IG, Timmons JA, Cannon B, Nedergaard J. Chronic peroxisome proliferator-activated receptor gamma (PPARgamma) activation of epididymally derived white adipocyte cultures reveals a population of thermogenically competent, UCP1-containing adipocytes molecularly distinct from classic brown adipocytes. J Biol Chem 2010; 285: 7153–64. Wu J, Bostrom P, Sparks LM et al. Beige adipocytes are a distinct type of thermogenic fat cell in mouse and human. Cell 2012; 150: 366–76. Seale P, Bjork B, Yang W et al. PRDM16 controls a brown fat/skeletal muscle switch. Nature 2008; 454: 961–7. Atit R, Sgaier SK, Mohamed OA et al. Beta-catenin activation is necessary and sufficient to specify the dorsal dermal fate in the mouse. Dev Biol 2006; 296: 164–76. Sharp LZ, Shinoda K, Ohno H et al. Human BAT possesses molecular signatures that resemble beige/brite cells. PLoS ONE 2012; 7: e49452. Lee YH, Petkova AP, Mottillo EP, Granneman JG. In vivo identification of bipotential adipocyte progenitors recruited by beta3-adrenoceptor activation and high-fat feeding. Cell Metab 2012; 15: 480–91. Cannon B, Nedergaard J. Cell biology: neither brown nor white. Nature 2012; 488: 286–7. Lidell ME, Betz MJ, Dahlqvist Leinhard O et al. Evidence for two types of brown adipose tissue in humans. Nat Med 2013; 19: 631–4. Cypess AM, White AP, Vernochet C et al. Anatomical localization, gene expression profiling and functional characterization of adult human neck brown fat. Nat Med 2013; 19: 635–9. Jespersen NZ, Larsen TJ, Peijs L et al. A classical brown adipose tissue mRNA signature partly overlaps with brite in the supraclavicular region of adult humans. Cell Metab 2013; 17: 798–805. Xue B, Rim JS, Hogan JC, Coulter AA, Koza RA, Kozak LP. Genetic variability affects the development of brown adipocytes in white fat but not in interscapular brown fat. J Lipid Res 2007; 48: 41–51. Seale P, Kajimura S, Yang W et al. Transcriptional control of brown fat determination by PRDM16. Cell Metab 2007; 6: 38–54. Kajimura S, Seale P, Tomaru T et al. Regulation of the brown and white fat gene programs through a PRDM16/CtBP transcriptional complex. Genes Dev 2008; 22: 1397–409. Seale P, Conroe HM, Estall J et al. Prdm16 determines the thermogenic program of subcutaneous white adipose tissue in mice. J Clin Invest 2011; 121: 96–105. Ohno H, Shinoda K, Spiegelman BM, Kajimura S. PPARgamma agonists induce a white-to-brown fat conversion through stabilization of PRDM16 protein. Cell Metab 2012; 15: 395–404.

Review: Brown adipose tissue

59 Kajimura S, Seale P, Kubota K et al. Initiation of myoblast to brown fat switch by a PRDM16-C/EBP-beta transcriptional complex. Nature 2009; 460: 1154–8. 60 Lin J, Wu PH, Tarr PT et al. Defects in adaptive energy metabolism with CNS-linked hyperactivity in PGC-1alpha null mice. Cell 2004; 119: 121–35. 61 Uldry M, Yang W, St-Pierre J, Lin J, Seale P, Spiegelman BM. Complementary action of the PGC-1 coactivators in mitochondrial biogenesis and brown fat differentiation. Cell Metab 2006; 3: 333–41. 62 Kleiner S, Mepani RJ, Laznik D et al. Development of insulin resistance in mice lacking PGC-1alpha in adipose tissues. Proc Natl Acad Sci USA 2012; 109: 9635–40. 63 Ohno H, Shinoda K, Ohyama K, Sharp LZ, Kajimura S. EHMT1 controls brown adipose cell fate and thermogenesis through the PRDM16 complex. Nature 2013; 504: 163–7. 64 Qiang L, Wang L, Kon N et al. Brown remodeling of white adipose tissue by SirT1-dependent deacetylation of Ppargamma. Cell 2012; 150: 620–32. 65 Hallberg M, Morganstein DL, Kiskinis E et al. A functional interaction between RIP140 and PGC-1alpha regulates the expression of the lipid droplet protein CIDEA. Mol Cell Biol 2008; 28: 6785–95. 66 Christian M, Kiskinis E, Debevec D, Leonardsson G, White R, Parker MG. RIP140-targeted repression of gene expression in adipocytes. Mol Cell Biol 2005; 25: 9383–91. 67 Leonardsson G, Steel JH, Christian M et al. Nuclear receptor corepressor RIP140 regulates fat accumulation. Proc Natl Acad Sci USA 2004; 101: 8437–42. 68 Picard F, Gehin M, Annicotte J et al. SRC-1 and TIF2 control energy balance between white and brown adipose tissues. Cell 2002; 111: 931–41. 69 Coste A, Louet JF, Lagouge M et al. The genetic ablation of SRC-3 protects against obesity and improves insulin sensitivity by reducing the acetylation of PGC-1{alpha}. Proc Natl Acad Sci USA 2008; 105: 17187–92. 70 Lerin C, Rodgers JT, Kalume DE, Kim SH, Pandey A, Puigserver P. GCN5 acetyltransferase complex controls glucose metabolism through transcriptional repression of PGC-1alpha. Cell Metab 2006; 3: 429–38. 71 Rothwell NJ, Stock MJ. A role for brown adipose tissue in diet-induced thermogenesis. Nature 1979; 281: 31–5. 72 Dulloo AG, Miller DS. Energy balance following sympathetic denervation of brown adipose tissue. Can J Physiol Pharmacol 1984; 62: 235–40. 73 Feldmann HM, Golozoubova V, Cannon B, Nedergaard J. UCP1 ablation induces obesity and abolishes diet-induced thermogenesis in mice exempt from thermal stress by living at thermoneutrality. Cell Metab 2009; 9: 203–9. 74 Kopecky J, Clarke G, Enerback S, Spiegelman B, Kozak LP. Expression of the mitochondrial uncoupling protein gene from the aP2 gene promoter prevents genetic obesity. J Clin Invest 1995; 96: 2914–23. 75 Kim JK, Kim HJ, Park SY et al. Adipocyte-specific overexpression of FOXC2 prevents diet-induced increases in intramuscular fatty acyl CoA and insulin resistance. Diabetes 2005; 54: 1657–63. 76 Collins S, Daniel KW, Petro AE, Surwit RS. Strain-specific response to b3-adrenergic receptor agonist treatment of diet-induced obesity in mice. Endocrinology 1997; 138: 405–13.

ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

375

M. E. Lidell et al.

77 Collins S, Daniel KW, Rohlfs EM, Ramkumar V, Taylor IL, Gettys TW. Impaired expression and functional activity of the beta 3- and beta 1-adrenergic receptors in adipose tissue of congenitally obese (C57BL/6J ob/ob) mice. Mol Endocrinol 1994; 8: 518–27. 78 Ouellet V, Labbe SM, Blondin DP et al. Brown adipose tissue oxidative metabolism contributes to energy expenditure during acute cold exposure in humans. J Clin Invest 2012; 122: 545–52. 79 Bartelt A, Bruns OT, Reimer R et al. Brown adipose tissue activity controls triglyceride clearance. Nat Med 2011; 17: 200–5. 80 Geerling JJ, Boon MR, van der Zon GC et al. Metformin lowers plasma triglycerides by promoting VLDL-triglyceride clearance by brown adipose tissue in mice. Diabetes 2014; 63: 880–91. 81 Gasparetti AL, de Souza CT, Pereira-da-Silva M et al. Cold exposure induces tissue-specific modulation of the insulin-signalling pathway in Rattus norvegicus. J Physiol 2003; 552: 149–62. 82 Ghorbani M, Himms-Hagen J. Appearance of brown adipocytes in white adipose tissue during CL 316,243-induced reversal of obesity and diabetes in Zucker fa/fa rats. Int J Obes Relat Metab Disord 1997; 21: 465–75. 83 Umekawa T, Yoshida T, Sakane N, Saito M, Kumamoto K, Kondo M. Anti-obesity and anti-diabetic effects of CL316,243, a highly specific beta 3-adrenoceptor agonist, in Otsuka Long-Evans Tokushima Fatty rats: induction of uncoupling protein and activation of glucose transporter 4 in white fat. Eur J Endocrinol 1997; 136: 429–37. 84 Lean ME, James WP, Jennings G, Trayhurn P. Brown adipose tissue in patients with phaeochromocytoma. Int J Obes 1986; 10: 219–27. 85 Ricquier D, Nechad M, Mory G. Ultrastructural and biochemical characterization of human brown adipose tissue in pheochromocytoma. J Clin Endocrinol Metab 1982; 54: 803–7. 86 Wang Q, Zhang M, Ning G et al. Brown adipose tissue in humans is activated by elevated plasma catecholamines levels and is inversely related to central obesity. PLoS ONE 2011; 6: e21006. 87 Silva JE, Larsen PR. Adrenergic activation of triiodothyronine production in brown adipose tissue. Nature 1983; 305: 712–3. 88 Ribeiro MO, Carvalho SD, Schultz JJ et al. Thyroid hormone–sympathetic interaction and adaptive thermogenesis are thyroid hormone receptor isoform–specific. J Clin Invest 2001; 108: 97–105. 89 Ribeiro MO, Bianco SD, Kaneshige M et al. Expression of uncoupling protein 1 in mouse brown adipose tissue is thyroid hormone receptor-beta isoform specific and required for adaptive thermogenesis. Endocrinology 2010; 151: 432–40. 90 Wulf A, Harneit A, Kroger M, Kebenko M, Wetzel MG, Weitzel JM. T3-mediated expression of PGC-1alpha via a far upstream located thyroid hormone response element. Mol Cell Endocrinol 2008; 287: 90–5. 91 Lahesmaa M, Orava J, Schalin-Jantti C et al. Hyperthyroidism increases brown fat metabolism in humans. J Clin Endocrinol Metab 2013; 99: E28–35. 92 Watanabe M, Houten SM, Mataki C et al. Bile acids induce energy expenditure by promoting intracellular thyroid hormone activation. Nature 2006; 439: 484–9.

376

ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

Review: Brown adipose tissue

93 Thomas C, Gioiello A, Noriega L et al. TGR5-mediated bile acid sensing controls glucose homeostasis. Cell Metab 2009; 10: 167–77. 94 Tseng YH, Kokkotou E, Schulz TJ et al. New role of bone morphogenetic protein 7 in brown adipogenesis and energy expenditure. Nature 2008; 454: 1000–4. 95 Boon MR, van den Berg SA, Wang Y et al. BMP7 activates brown adipose tissue and reduces diet-induced obesity only at subthermoneutrality. PLoS ONE 2013; 8: e74083. 96 Whittle AJ, Carobbio S, Martins L et al. BMP8B increases brown adipose tissue thermogenesis through both central and peripheral actions. Cell 2012; 149: 871–85. 97 Hondares E, Rosell M, Gonzalez FJ, Giralt M, Iglesias R, Villarroya F. Hepatic FGF21 expression is induced at birth via PPARalpha in response to milk intake and contributes to thermogenic activation of neonatal brown fat. Cell Metab 2010; 11: 206–12. 98 Hondares E, Iglesias R, Giralt A et al. Thermogenic activation induces FGF21 expression and release in brown adipose tissue. J Biol Chem 2011; 286: 12983–90. 99 Chartoumpekis DV, Habeos IG, Ziros PG, Psyrogiannis AI, Kyriazopoulou VE, Papavassiliou AG. Brown adipose tissue responds to cold and adrenergic stimulation by induction of FGF21. Mol Med 2011; 17: 736–40. 100 Camporez JP, Jornayvaz FR, Petersen MC et al. Cellular mechanisms by which FGF21 improves insulin sensitivity in male mice. Endocrinology 2013; 154: 3099–109. 101 Bordicchia M, Liu D, Amri EZ et al. Cardiac natriuretic peptides act via p38 MAPK to induce the brown fat thermogenic program in mouse and human adipocytes. J Clin Invest 2012; 122: 1022–36. 102 Raschke S, Elsen M, Gassenhuber H et al. Evidence against a Beneficial Effect of Irisin in Humans. PLoS ONE 2013; 8: e73680. 103 Zhang Y, Li R, Meng Y et al. Irisin stimulates browning of white adipocytes through mitogen-activated protein kinase p38 MAP kinase and ERK MAP kinase signaling. Diabetes 2014; 63: 514–25. 104 Lee P, Linderman JD, Smith S et al. Irisin and FGF21 Are Cold-Induced Endocrine Activators of Brown Fat Function in Humans. Cell Metab 2014; 19: 302–9. 105 Berg JM, Tymoczko JL, Stryer L. Biochemistry. New York: W. H. Freeman, 2007. 106 Harper JA, Dickinson K, Brand MD. Mitochondrial uncoupling as a target for drug development for the treatment of obesity. Obes Rev 2001; 2: 255–65. 107 van Baak MA, Hul GB, Toubro S et al. Acute effect of L-796568, a novel beta 3-adrenergic receptor agonist, on energy expenditure in obese men. Clin Pharmacol Ther 2002; 71: 272–9. 108 Larsen TM, Toubro S, van Baak MA et al. Effect of a 28-d treatment with L-796568, a novel beta(3)-adrenergic receptor agonist, on energy expenditure and body composition in obese men. Am J Clin Nutr 2002; 76: 780–8. 109 Weyer C, Tataranni PA, Snitker S, Danforth E Jr, Ravussin E. Increase in insulin action and fat oxidation after treatment with CL 316,243, a highly selective beta3adrenoceptor agonist in humans. Diabetes 1998; 47: 1555–61. 110 Orava J, Nuutila P, Noponen T et al. Blunted metabolic responses to cold and insulin stimulation in brown adipose tissue of obese humans. Obesity 2013; 21: 2279–87.

M. E. Lidell et al.

111 Vijgen GH, Bouvy ND, Teule GJ, Brans B, Schrauwen P, Van Marken Lichtenbelt WD. Brown adipose tissue in morbidly obese subjects. PLoS ONE 2011; 6: e17247. 112 Liu YL, Heal DJ, Stock MJ. Mechanism of the thermogenic effect of Metabolite 2 (BTS 54 505), a major pharmacologically active metabolite of the novel anti-obesity drug, sibutramine. Int J Obes Relat Metab Disord 2002; 26: 1245–53. 113 Baba S, Tatsumi M, Ishimori T, Lilien DL, Engles JM, Wahl RL. Effect of nicotine and ephedrine on the accumulation of 18F-FDG in brown adipose tissue. J Nucl Med 2007; 48: 981–6. 114 Dulloo AG, Seydoux J, Girardier L. Peripheral mechanisms of thermogenesis induced by ephedrine and caffeine in brown adipose tissue. Int J Obes 1991; 15: 317–26. 115 Watanabe M, Morimoto K, Houten SM et al. Bile acid binding resin improves metabolic control through the induction of energy expenditure. PLoS ONE 2012; 7: e38286.

Review: Brown adipose tissue

116 Amorim BS, Ueta CB, Freitas BC et al. A TRbeta-selective agonist confers resistance to diet-induced obesity. J Endocrinol 2009; 203: 291–9. 117 Villicev CM, Freitas FR, Aoki MS et al. Thyroid hormone receptor beta-specific agonist GC-1 increases energy expenditure and prevents fat-mass accumulation in rats. J Endocrinol 2007; 193: 21–9. 118 Symonds ME, Henderson K, Elvidge L et al. Thermal imaging to assess age-related changes of skin temperature within the supraclavicular region co-locating with brown adipose tissue in healthy children. J Pediatr 2012; 161: 892–8. 119 Yoneshiro T, Aita S, Matsushita M et al. Recruited brown adipose tissue as an antiobesity agent in humans. J Clin Invest 2013; 123: 3404–8. Correspondence: Sven Enerb€ ack, Department of Medical and Clinical Genetics, Institute of Biomedicine, University of Gothenburg, P.O. Box 440, SE 40530 Gothenburg, Sweden. (fax: +46 31 416108; e-mail: [email protected]).

ª 2014 The Association for the Publication of the Journal of Internal Medicine Journal of Internal Medicine, 2014, 276; 364–377

377

Brown adipose tissue and its therapeutic potential.

Obesity and related diseases are a major cause of human morbidity and mortality and constitute a substantial economic burden for society. Effective tr...
590KB Sizes 3 Downloads 4 Views