Anda di halaman 1dari 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/265173825

Endocannabinoid signaling and food addiction

Article  in  Neuroscience & Biobehavioral Reviews · August 2014


DOI: 10.1016/j.neubiorev.2014.08.008 · Source: PubMed

CITATIONS READS

45 698

8 authors, including:

Claudio d'addario Maria Vittoria Micioni Di Bonaventura


Università degli Studi di Teramo University of Camerino
95 PUBLICATIONS   1,330 CITATIONS    59 PUBLICATIONS   584 CITATIONS   

SEE PROFILE SEE PROFILE

Mariangela Pucci Adele Romano


Università degli Studi di Teramo Sapienza University of Rome
42 PUBLICATIONS   335 CITATIONS    39 PUBLICATIONS   818 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

any kind of projects of interest for cosmetic application View project

Vulnerability factors for adolescent alcohol consumption, as assessed via animal rat models View project

All content following this page was uploaded by Maria Vittoria Micioni Di Bonaventura on 22 January 2019.

The user has requested enhancement of the downloaded file.


Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

Contents lists available at ScienceDirect

Neuroscience and Biobehavioral Reviews


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

Review

Endocannabinoid signaling and food addiction


C. D’Addario a,e,∗ , M.V. Micioni Di Bonaventura b , M. Pucci a , A. Romano c , S. Gaetani c ,
R. Ciccocioppo b , C. Cifani b , M. Maccarrone d,f,∗∗
a
Faculty of Bioscience and Technology for Food, Agriculture and Environment, University of Teramo, Piazza A Moro 45, 64100 Teramo, Italy
b
School of Pharmacy, Pharmacology Unit, University of Camerino, Via Madonna delle Carceri 9, 62032 Camerino, MC, Italy
c
Department of Physiology and Pharmacology “Vittorio Erspamer”, Sapienza University of Rome, Piazzale Aldo Moro 5, 00185 Rome, Italy
d
School of Medicine and Center of Integrated Research, Campus Bio-Medico University of Rome, Via Alvaro del Portillo 21, 00128 Rome, Italy
e
Department of Clinical Neuroscience, Center for Molecular Medicine, Karolinska Institutet, SE-171 77 Stockholm, Sweden
f
European Center for Brain Research (CERC)/Santa Lucia Foundation, Via del Fosso di Fiorano 64, 00143 Rome, Italy

a r t i c l e i n f o a b s t r a c t

Article history: Overeating, frequently linked to an increasing incidence of overweight and obesity, has become epidemic
Received 3 May 2014 and one of the leading global health problems. To explain the development of this eating behavior, new
Received in revised form 28 July 2014 hypotheses involve the concept that many people might be addicted to food by losing control over their
Accepted 18 August 2014
ability to regulate food intake. Among the different neurotransmitter networks that partake in the reward
Available online 27 August 2014
circuitry within the brain, a large body of evidence supports the involvement of the endocannabinoid
system. Indeed, its dysfunctions might contribute to food addiction, by regulating appetite and food
Keywords:
preference through central and peripheral mechanisms. Here, we review and discuss the role of endo-
Endocannabinoid signaling
Food addiction
cannabinoid signaling in the reward circuitry, and the possible therapeutic exploitation of strategies
Food reward mechanisms based on its fine regulation.
Food intake regulation and energy balance © 2014 Elsevier Ltd. All rights reserved.

Contents

1. The endocannabinoid system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204


1.1. The endocannabinoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
1.2. eCBs metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
1.3. Molecular targets and signaling pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
2. Role of eCB system in food intake regulation and energy balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
3. Regulation of endocannabinoid levels by the diet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

Abbreviations: 2-AG, 2-arachidonoylglycerol; AA, arachidonic acid; ABA, activity-based anorexia; ACC1, acetyl coenzyme-A carboxylase-1; AEA, N-
arachidonoylethanolamine; AN, anorexia nervosa; ARC, arcuate nucleus; BBB, blood-brain barrier; BED, binge eating disorder; BN, bulimia nervosa; CB, cannabinoid; CB1, CB
receptor subtype 1; CB2, CB receptor subtype 2; CCK, colecistokinin; CRH, corticotrophin-releasing hormone; DA, dopamine; DAG, diacylglycerol; DAGL, diacylglycerol lipase;
DHA, docosahexanoic acid; DMH, dorsomedial hypothalamus; DMS-5, Diagnostic and Statistical Manual of Mental Disorders, Fifth Edition; DOP, ␦-opioid receptor; eCB, endo-
cannabinoid; EDI-2, the Eating Disorder Inventory-2; EMA, European Medicines Agency; EMT, eCB membrane transporter; EPA, eicosapentaenoic paraventricular nucleus
acid; FAAH, fatty acid amide hydrolase; FAA, fatty acid amide; FAS, fatty acid synthase; Fa, fatty acid; FDA, Food and Drug Adminstration; GABA, ␥-aminobutyric acid; GI,
gastro-intestinal; GPR119, G protein coupled receptor 119; GPR55, G protein-coupled receptor 55; HFD, high fat diet; Icv, intracerebroventricular; LA, linoleic acid; LH, lateral
hypothalamus; MAGL, monoacylglycerol lipase; MAG, monoacylglycerol; MCH, melanin-concentrating hormone; MOP, ␮-opioid receptor; NAc, nucleus accumbens; NADA,
N-arachidonoyldopamine; NAE, N-acylethanolamine; NAPE, N-acylphosphatidyl-ethanolamines; NAT, N-acyltransferase; NPY, neuropeptide Y; OEA, N-oleoylethanolamine;
OP, obesity-prone; PE, phosphatidylethanolamine; PEA, palmitoylethanolamine; PET, positron emission tomography; PI3K, phosphoinositide 3-kinase; PKA, protein kinase A;
PLC, phospholipase C; PLD, phospholipase D; POMC, proopiomelanocortin; PPAR, peroxisome proliferator-activated receptor; PUFA, polyunsaturated fatty acid; PYY, peptide
YY; SEA, N-stearoylethanolamine; THC, 9-tetrahydrocannabinol; TRPV1, transient receptor potential vanilloid 1; VTA, ventral tegmental area; WAT, white adipose tissue;
YFAS, Yale Food Addiction Scale.
∗ Corresponding author at: University of Teramo, Faculty of Bioscience and Technology for Food, Agriculture and Environment, Piazza A. Moro, 45, 64100 Teramo, Italy.
Tel.: +39 0861 266877; fax: +39 0737 403343.
∗∗ Corresponding author at: School of Medicine and Center of Integrated Research, Campus Bio-Medico University of Rome, Via Alvaro del Portillo 21, 00128 Rome, Italy.
Tel.: +39 06 2254 19169; fax: +39 06 2254 1456.
E-mail addresses: cdaddario@unite.it (C. D’Addario), m.maccarrone@unicampus.it (M. Maccarrone).

http://dx.doi.org/10.1016/j.neubiorev.2014.08.008
0149-7634/© 2014 Elsevier Ltd. All rights reserved.
204 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

4. Neurobiology of food addiction and food reward mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209


5. Cannabinoids and reward mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
6. eCB system-based drugs in eating disorders and obesity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
6.1. eCB system in eating disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
6.2. eCB system in obesity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217

1. The endocannabinoid system (NAPE-PLD) hydrolyzes NAPEs to release NAEs (included AEA), and
phosphatidic acid (Okamoto et al., 2004).
Since the isolation and characterization of the active phyto- 2-AG is mainly synthesized from AA-containing membrane
cannabinoids in the cannabis (Cannabis sativa) plant, including phospholipids through the action of phospholipase C (PLC), lead-
the most psychoactive ingredient 9 -tetrahydrocannabinol (THC) ing to the formation of diacylglycerol (DAG), and then through one
(Mechoulam and Gaoni, 1965; Mechoulam, 1970), and after the of two diacylglycerol lipases (DAGLs), DAGL␣ and DAGL␤ (Bisogno
identification and cloning of the target of THC, the type-1 cannabi- et al., 2003).
noid receptor (Devane et al., 1988; Herkenham et al., 1991; Matsuda Inactivation of eCB signaling is achieved by two steps: a
et al., 1990), and of its endogenous counterparts, collectively rapid removal from molecular targets and subsequent hydroly-
termed “endocannabinoids” (Devane et al., 1992; Mechoulam et al., sis by specific enzymatic systems. eCBs rapid diffusion through
1995; Sugiura et al., 1995), many efforts have been profused to the plasma membrane has been shown to be mediated by a
study the different physiological functions modulated by what is selective and saturable transporter, the eCB membrane trans-
known as “endocannabinoid system”. porter (EMT), responsible for both AEA and 2-AG uptake by
several cells (Di Marzo et al., 1994; Hájos et al., 2004; Chicca
et al., 2012). Moreover, an intracellular accumulation of AEA
1.1. The endocannabinoids in adiposomes has been also reported (Oddi et al., 2008). Still
under debate is whether or not eCB transport and degra-
The endocannabinoids (eCBs) are derivatives of arachidonic acid dation are coupled or independent processes (Fegley et al.,
(AA), resembling other lipid transmitters such as prostaglandins or 2004).
leukotrienes. They are conjugated with ethanolamine to form fatty Once taken up by cells, AEA and 2-AG can be degraded by fatty
acid amides (FAAs), or with glycerol to form monoacylglycerols acid amide hydrolase (FAAH), which breaks the amide or ester
(MAGs), N-arachidonoylethanolamine (anandamide, AEA) and 2- bond, to release AA and ethanolamine or glycerol, respectively
arachidonoylglycerol (2-AG) so far representing the best studied (Cravatt et al., 1996). Yet, the main responsible for 2-AG metabolism
and most active members of each class, respectively (Devane et al., is monoacylglycerol lipase (MAGL) (Dinh et al., 2002). The degrada-
1992; Mechoulam et al., 1995). There are also other endoge- tion products are then recycled into the membrane phospholipids,
nous FAAs, like the appetite-suppressor N-oleoylethanolamine where they produce de novo the two eCBs (Bisogno et al., 2005).
(OEA) (Fu et al., 2008), the antiinflammatory, anticonvulsant and Moreover, AEA can be hydrolysed also by other enzymes with an
antiproliferative N-palmitoylethanolamine (PEA) (Lambert et al., “amidase signature”, like FAAH-2 (Wei et al., 2006), or that belong
2001), and the immunomodulator N-stearoylethanolamine (SEA) to the choloylglycine hydrolase family, like N-acylethanolamine-
(Maccarrone et al., 2002). The latter are called “eCB-like” com- hydrolysing acid amidase (Tsuboi et al., 2005). When MAGL or FAAH
pounds, since they do not activate cannabinoid receptors directly activity is suppressed, AEA and 2-AG might become substrates for
but have an “entourage effect” (Ben-Shabat et al., 1998; Lambert lipoxygenases (Van der Stelt et al., 2002), cyclooxygenase-2 (Rouzer
and Di Marzo, 1999; De Petrocellis et al., 2004). Moreover, 2- and Marnett, 2011) or cytochrome P450 (Snider et al., 2010), gen-
arachidonoylglycerolether (noladin ether) (Hanus et al., 2001), erating oxidative derivatives endowed with their own biological
N-arachidonoyldopamine (NADA) (Bisogno et al., 2000), and the activities. It has been also reported that AEA and possibly other
“inverted anandamide” virodhamine (Porter et al., 2002) also eCBs are transported intracellularly by distinct carriers, like fatty
belong to the ever-growing eCBs family. The chemical structures acid binding proteins (Kaczocha et al., 2009), albumin and heat
of these substances are shown in Fig. 1. shock protein 70 (Oddi et al., 2009), or a truncated FAAH termed
“FAAH-1-like AEA transporter” (Fu et al., 2011; Leung et al., 2013).
Therefore, the intracellular trafficking of eCBs might represent a
1.2. eCBs metabolism
new dimension to drive distinct signaling cascades of these com-
pounds in the CNS and at the periphery (Maccarrone et al., 2010a,b;
eCBs are produced by multiple synthetic pathways from lipid
Kaczocha et al., 2012).
precursors present in cell membranes “on demand”, that is when
and where needed following physiological or pathological stimuli.
However, AEA has been shown to have the potential to accumu- 1.3. Molecular targets and signaling pathways
late in intracellular stores called adiposomes (or lipid droplets)
(Oddi et al., 2008), whereas 2-AG might be pre-formed and AEA and 2-AG activate different signaling pathways by binding
sequestered in distinct intracellular pools until needed (Alger and to (with different affinities) and activating two well-characterized
Kim, 2011). 7-transmembrane G protein-coupled cannabinoid (CB) receptor
The best known biosynthetic pathway of AEA and other subtypes: type-1 (CB1 ) (Herkenham et al., 1991) and type-2 (CB2 )
N-acylethanolamines (NAEs) occurs in two steps: first a calcium- (Munro et al., 1993). In humans, CB1 is localized preferentially
dependent transacylase (NAT, N-acyltransferase) transfers an acyl in the terminals of central and peripheral neurons and glial cells
group from a membrane phospholipid to the N-position of (Egertová et al., 2003), but is also expressed in peripheral tis-
phosphatidylethanolamine (PE), to generate N-acylphosphatidyl- sues like heart, uterus, testis, liver and small intestine, as well
ethanolamines (NAPEs); then, a NAPE-selective phospholipase D as in immune cells (Maccarrone et al., 2001; Nong et al., 2001;
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 205

Fig. 1. Chemical structures of the major endocannabinoids and related compounds.

Klein et al., 2003) and adipose tissue (Spoto et al., 2006). Pheripheral potential for the treatment of diet-induced obesity. Indeed, these
tissues also express CB2 (Roche et al., 2006), that is predominantly substances decrease food intake and body weight gain, not only
found at high levels in leukocytes and spleen, and to a lower extent in animal models but also in humans (Cota, 2007; Di Marzo and
in muscle, liver, intestine and testis (Liu et al., 2009). Interestingly, Desprès, 2009). Earlier studies had shown that pharmacological
CB2 is also present in the brain (Nunez et al., 2004; Van Sickle et al., stimulation of CB1 by systemic administration of plant-derived,
2005), where it is markedly up-regulated upon various stressing synthetic, or endogenous cannabinoids stimulate eating and pro-
conditions (Viscomi et al., 2009). duce anabolic effects (Williams et al., 1998; Williams and Kirkham,
Besides the well-characterized CB1 and CB2 receptors, pharma- 1999; Hao et al., 2000). Moreover, using mice lacking CB1 , it was
cological evidence suggests the existence of two additional GPCRs documented that eCBs actions on food intake and body weight
as eCBs-binding receptors: G protein-coupled receptor 55 (GPR55), depend on the functional expression and activity of CB1 (Cota et al.,
and G protein coupled receptor 119 (GPR119). GPR55 is a purported 2003).
“type-3” (“CB3 ”) cannabinoid receptor, that shares low sequence In the brain, eCB system tonically reinforces the motivation to
homology (10%–15%) with CB1 and CB2 (Ross, 2009), and has been find and consume food, by interacting with the mesolimbic path-
found in both CNS (Sawzdargo et al., 1999) and peripheral tissues ways engaged in reward mechanisms, whereas it is activated “on
(Brown and Robin Hiley, 2009). Also GPR119 has been found in var- demand” in the hypothalamus to transiently regulate levels and/or
ious mammalian species (Fredriksson et al., 2003), and in humans actions of other orexigenic and anorexigenic mediators (Di Marzo
is expressed in pancreas, liver and gastrointestinal tract (Overton and Matias, 2005). eCBs levels in both hypothalamus and limbic
et al., 2008). AEA, but not 2-AG, also activates at an intracellular site forebrain are highest during food deprivation and lowest during
the transient receptor potential vanilloid 1 (TRPV1) ion channel (Di food consumption (Kirkham et al., 2002). When directly injected
Marzo and De Petrocellis, 2010), expressed in peripheral sensory into the hypothalamus or the NAc shell, eCBs induce food intake
fibers and also in several nuclei of the CNS (Marinelli et al., 2003). even in satiated animals (Kirkham et al., 2002; Jamshidi and Taylor,
Finally, the peroxisome proliferator-activated receptor (PPAR) sub- 2001). In agreement with this observation, local administration
types ␣ and ␥ are also activated by eCBs for the regulation of of N-arachidonoylserotonin (an inhibitor of eCBs hydrolysis by
lipid and glucose metabolism, as well as of inflammatory responses FAAH) in the NAc shell also stimulates food intake in parallel to
(Gasperi et al., 2007; Pagano et al., 2008). All the major elements of c-Fos expression in the arcuate nucleus (ARC) and paraventricular
the eCB system are schematically depicted in Fig. 2. nucleus (PVN), as well as in the dorsomedial (DMH) and lateral (LH)
hypothalamus (Soria et al., 2006).
2. Role of eCB system in food intake regulation and energy Exogenous cannabinoids, via CB1 stimulation, inhibit gluta-
balance matergic signaling in pro-opiomelanocortin (POMC) neurons, thus
reducing the release of melanocortin, an anorexigenic media-
The role of eCB system in maintaining energy balance has tor (Hentges et al., 2005). This finding is further supported by
received great attention since when synthetic compounds, like additional evidence showing that intracerebroventricular (icv)
the CB1 antagonist/inverse agonist SR141716A (rimonabant), also administration of CB1 antagonists can attenuate the orexigenic
known as Acomplia® and developed by Sanofi-Aventis, have shown effect of melanocortin antagonists (Verty et al., 2004). It seems
206 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

Fig. 2. Overview of the eCB system indicating the most likely targets of eCBs, their anabolic and catabolic pathways, and intracellular localization. MAGs, Monoacylglycerol;
2-AG, 2-Arachidonoylglycerol; FAAs, Fatty acid amides; AEA, Anandamide; OEA, N-Oleoylethanolamine; SEA, N-stearoylethanolamine; PEA, N-Palmitoylethanolamine; AA,
Arachidonic acid; CB1/2 , type 1/2 cannabinoid receptors; GPR55/119, G protein-coupled receptor 55/119; TRPV1, transient receptor potential vanilloid 1 channel; EMT, endo-
cannabinoid membrane transporter; NAPE-PLD, N-acylphosphatidylethanolamine-selective phospholipase D; DAGL, diacylglycerol lipase; FAAH, fatty acid amide hydrolase;
NAAA, N-acylethanolamine-hydrolyzing acid amidase; MAGL, monoacylglycerol lipase; COX-2, cyclooxygenase-2; LOXs, lipoxygenases; eCBs, endocannabinoids; PPAR␥/␣,
Peroxisome proliferator-activated receptor ␥ and ␣; PG-G PG-EA See text for details.

that CB1 acts downstream of melanocortin receptors, because neurons, inducing a reduction in local inhibitory transmission and
melanocortin agonists administered icv do not alter hypothalamic thus mediating a hypophagic action (Bellocchio et al., 2010).
AEA nor 2-AG levels (Matias et al., 2008a,b), while central admin- The modulation of the expression of several anorexigenic and
istration of melanocortin antagonists leads to a delayed increase orexigenic peptides by eCB system in the hypothalamus is often
of hypothalamic eCBs (Matias et al., 2008a,b). In addition, AEA counterbalanced by opposite actions mediated by the adipose-
was shown to increase the secretion of the orexigenic media- derived hormone leptin, another key player in the regulation of
tor neuropeptide Y (NPY) (Gamber et al., 2005). However, CB1 energy intake and expenditure (included appetite and hunger,
is not expressed by NPY-secreting neurons of the hypothalamus metabolism and behavior). eCB levels in the hypothalamus seemed
(Cota et al., 2003), and rimonabant is able to reduce food intake to inversely correlate with leptin plasma levels (Di Marzo et al.,
equally well in wild-type and NPY null-mice, making it unlikely 2001). Moreover, leptin directly inhibits eCB synthesis by reducing
that eCBs can modulate food intake through NPY (Bermudez-Silva both intracellular calcium levels in MCH neurons (Jo et al., 2005)
et al., 2012). Within the PVN, eCBs can be released “on demand” and glucocorticoid-mediated eCB release in the PVN (Malcher-
upon activation of the orexigenic substance glucocorticoid (Di Lopes et al., 2006). In the same context, it should be recalled that
et al., 2003), which inhibits (among others) the anorexigenic medi- in humans leptin can also promote eCBs hydrolysis by enhancing
ator corticotrophin-releasing hormone (CRH). In keeping with FAAH gene expression and activity via binding to a distinct site
these observations, CRH levels were found to be higher in CB1 - of the FAAH promoter (Maccarrone et al., 2003a,b). Such a tight
deficient mice (Cota et al., 2003). Moreover, CB1 agonists inhibit cross-talk between eCB system and leptin seems to affect also the
orexin neurons (Huang et al., 2007), while exciting the orexigenic activity of the brain reward system. Obese rats with defective lep-
melanin-concentrating hormone (MCH) neurons in LH (Morrison tin signaling show increased CB1 expression and binding activity
and Berthoud, 2007). Interestingly, in vitro studies have shown that in reward brain structures (Thanos et al., 2008). The interaction
CB1 and orexin-1 receptors are present as heterodimers/oligomers between eCB system and leptin seems noteworthy also outside
in intracellular vesicles, and that treatment with rimonabant blocks the CNS, in metabolically relevant organs like the white adipose
the activation by orexin A of intracellular pathways under control tissue (WAT). Indeed, it was shown that the administration of
of the orexin receptor (Ellis et al., 2006). It has been also proposed a leptin in the medio-basal hypothalamus can inhibit WAT lipoge-
bimodal regulation of food intake by eCB system involving the con- nesis by engaging hypothalamic phosphoinositide 3-kinase (PI3K)
trol by CB1 , on glutamatergic transmission, possibly responsible for signaling, thus activating sympathetic innervations on the adipose
the well-known orexigenic role, and on ventrostriatal GABAergic tissue and locally inhibiting AEA production (Buettner et al., 2008).
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 207

It has been reported that fatty acid (FA) sensing in the CNS soleus muscle, pharmacological blockade of CB1 was shown to
is associated with metabolic traits such as energy expenditure, improve both basal and insulin-stimulated glucose transport activ-
plasma glucose and substrate utilization in both rodents (Lam et al., ity, while CB1 activation had the opposite effect (Lindborg et al.,
2005) and humans (Jumpertz et al., 2012). Besides the hypothala- 2010).
mus and the reward system, eCB system regulates food intake and Several studies have demonstrated the presence of both CB1
energy balance also at the level of the vagus nerve, that connects the and CB2 in rodent and human pancreatic islets (Li et al., 2011a),
gastro-intestinal (GI) tract to key brainstem nuclei and contributes where eCBs seem to play a role in the regulation of endocrine
to the so called “gut-to-brain axis”, where a number of signals origi- secretion (Bermúdez-Silva et al., 2008). This hypothesis was ini-
nating from the GI tract are implicated too. Among these, hormones tially suggested by in vitro studies, in which a down-regulation of
like colecistokinin (CCK), peptide YY (PYY), and ghrelin exert impor- pancreatic basal insulin hypersecretion was observed when islets
tant functions by regulating meal duration and termination, and by cultured from obese Zucker rats were exposed to rimonabant for
inducing satiety (Strader and Woods, 2005). It has been shown that 24 h (Getty-Kaushik et al., 2009). As yet, conflicting results have
CB1 expression in vagal afferent neurons is increased in fasting and been obtained in human studies, where both activation and block-
decreased after refeeding, under the control of CCK (Burdyga et al., ade of CB1 can modulate insulin secretion in pancreatic tissue (Li
2004). Interestingly, the down-regulation of CB1 expression during et al., 2011b).
refeeding is also prevented by the administration of ghrelin, sug- Apart from the role played by the eCB system in the modula-
gesting a role for the latter hormone in blocking CCK action on CB1 tion of GI hormones and of peripheral organs involved in energy
expression (Burdyga et al., 2006, 2010). Furthermore, the effects of metabolism, increasing evidence suggests that another key aspect
CCK on CB1 expression can also be blocked by co-administration of of the regulation of energy metabolism by eCB system might be
AEA, while the fasting-induced increase of CB1 mRNA can be pre- linked to gut microbiota, that is known to influence whole body
vented by the CB1 antagonist AM281 (Burdyga et al., 2010). Besides metabolism and energy balance (Turnbaugh et al., 2006). Mucci-
CCK, there might also be a potential link between eCB system and oli and colleagues have made observations in a colonic epithelial
PYY, since expression of Y2R, the receptor subtype that mediates monolayer cell model upon CB1 activation, demonstrating that gut
the anorectic action of PYY (Batterham et al., 2002), and CB1 mRNAs microbiota control intestinal eCB system tone, which in turn mod-
are oppositely regulated in the vagal afferent neurons (Burdyga ulates gut permeability (Muccioli et al., 2010).
et al., 2010). In this context, Y2R mRNA expression is increased On a final note, it is known that orosensory positive feedback
by refeeding or CCK treatment, while high doses of AEA block CCK- mechanisms play a pivotal role in the rewarding properties of food
induced increase of Y2R expression (Burdyga et al., 2010). (Greenberg and Smith, 1996). Recent evidence demonstrated that
Among gut hormones, ghrelin is known to exert an orexigenic CB1 receptors are present in cells of the fungiform papillae of the
action (Wiedmer et al., 2007). In rats, a sub-anorectic dose of mouse tongue, where they mainly colocalize with type-1 taste
rimonabant is able to abolish the orexigenic effect of an intra-PVN receptor 3, a putative sweet receptor (Montmayeur et al., 2001;
injection of ghrelin, demonstrating for the first time a cross-talk Max et al., 2001). Indeed, pharmacological activation of the latter
between ghrelin and eCB system (Tucci et al., 2004). A further study receptors enhances the neural responses to sweet foods, but not
showed that the release of ghrelin could depend on eCB system by to bitter, umami, salty, or sour substances (Yoshida et al., 2010).
activation, since systemic administration of rimonabant reduces This effect was observed both in vivo and in vitro after application
circulating ghrelin levels (Cani et al., 2004). Later on, further stud- of CB1 agonists to isolated taste cells (Yoshida et al., 2010). Inter-
ies have demonstrated that ghrelin increases hypothalamic eCB estingly, recent observations demonstrated that the oral exposure
content (Kola et al., 2008), and that the administration of CB1 ago- to dietary fats mobilizes eCBs in the rat proximal small intestine
nists or of methanandamide (a non-hydrolyzable AEA analog) in (DiPatrizio et al., 2011), while intraduodenal infusion of rimona-
rats increases plasma ghrelin levels and ghrelin secretion from bant inhibit fat intake. Altogether, these findings support the view
gastric X/A-like cells (named X cells for their unknown functions that eCBs in the gut may play a major role in stimulating the intake
and A-like cells for the similarity with pancreatic A-cells) (Zbucki of fatty meals (DiPatrizio et al., 2011), thus acting as a crucial com-
et al., 2008). In addition to their role in food intake, forebrain eCBs ponent of the positive orosensory feedback mechanism that drives
regulate energy homeostasis by modifying also the activity of the fat intake. Finally, it was also observed that CB1 increases odor
sympathetic nervous system (Quarta et al., 2010; Jung et al., 2012). detection and food intake in fasted mice through cortical feedback
Indeed, it has been recently reported that mice selectively lacking projections to the main olfactory bulb (Soria-Gómez et al., 2014).
CB1 expression in forebrain neurons, and showing a 60% reduction Table 1 summarizes the above-reported studies.
of the expression of the same receptor in the cervical sympathetic
ganglia, have a lean phenotype and are resistant to diet-induced
obesity (Quarta et al., 2010). This is due to an increase in lipid oxi- 3. Regulation of endocannabinoid levels by the diet
dation and thermogenesis caused by an enhanced sympathetic tone
(Quarta et al., 2010). Essential FAs, such as ␣-linolenic and linoleic acids, are con-
Conversely, it has been shown that activation of CB1 stimu- verted respectively to ␻-3 and ␻-6 polyunsaturated fatty acids
lates fat deposition by facilitating adipocyte differentiation, and (PUFAs), that include the AEA and 2-AG moiety AA (Fig. 1). Since
by increasing expression of adipogenic enzymes and the activity dietary fats are the only source of PUFAs required for eCBs biosyn-
of lipoprotein lipase (Bensaid et al., 2003; Cota et al., 2003; Matias thesis, their intake clearly affects the modulation of eCB system
et al., 2006; Muccioli et al., 2010). functions (Maccarrone et al., 2010a,b; Bisogno and Maccarrone,
In concert with the action of the eCB system in the adipose 2014). Moreover, eCBs distribution in different tissues depends on
tissue, it was described that activation of CB1 expressed by hepa- the amounts of ingested PUFAs, also because diet per se is only a
tocytes can induce the expression of lipogenic enzymes, such as minor source of preformed eCBs (Berger et al., 2003; Bisogno and
acetyl coenzyme-A carboxylase-1 (ACC1) and fatty acid synthase Maccarrone, 2014). Increased levels of eCBs have been found in spe-
(FAS), which in turn increase de novo fatty acid synthesis and favor cific brain regions of newborn piglets fed milk supplemented with
the development of liver steatosis, during exposure to HFD (Osei- AA and docosahexanoic acid (DHA) (Berger et al., 2001). Similar
Hyiaman et al., 2005, 2008). results have been observed in whole brain of newborn mice fed up
Much less investigated is the specific role of the eCB system to day 58 an AA-rich diet (Berger et al., 2001; Berger and Roberts,
within the skeletal muscle and the endocrine pancreas. In isolated 2005). A decrease in hypothalamic AEA levels has been reprtedin
208 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

Table 1
link between CB1 activation, CB1 antagonism/deletion and central/peripheral signals involved in the control of food intake and metabolism.

Effects of CB1 activation Effects of CB1 antagosnim/genetic deletion Effects of other signals on eCB system elements

Melanocortin system
↓ Release of Hentges et al. (2005) ↓ orexigenic effect of Verty et al. (2004) melanocortin antagonists ↑ Matias et al. (2008a,b)
hypothalamic melanocortin hypothalamic eCBs synthesis
melanocortin antagonists

Leptin system
↑ Plasma leptin Monteleone et al. ↓ adipose tissue leptin Mastinu et al. (2012) Leptin ↑ eCBs synthesis both Di Marzo et al. (2001)
concentrations in (2005) expression centrally and peripherally Maccarrone et al.
patients with AN ↑ CB1 expression and binding (2003a,b)
activity in reward brain Buettner et al. (2008)
structures of obese rats with Thanos et al. (2008)
defective leptin signaling Di Marzo et al. (2001)
↑ hypothalamic eCBs levels in
ob/ob mice

CCK system
↓ CCK effects on CB1 Burdyga et al. (2010) ↓ fasting-induced Burdyga et al. (2010) CCK controls CB1 expression in Burdyga et al. (2004)
expression increase of CB1 mRNA vagal afferent neurons
↓CCK-induced increase Burdyga et al. (2010) expression of Y2R and CB1 Burdyga et al. (2010)
of Y2R expression mRNAs are oppositely
regulated in the vagal afferent
neurons

Other
↑ Plasma ghrelin levels Zbucki et al. (2008) ↓ orexigenic effect of Tucci et al. (2004) ghrelin ↑ hypothalamic eCB Kola et al. (2008)
and ghrelin secretion an intra-PVN injection Cani et al. (2004) content
from gastric X/A-like of ghrelin
cells. ↓ circulating ghrelin
levels
↑ CRH hypothalamic Glucocorticoid ↑ eCBs release Di et al. (2003)
levels in the PVN, while decreasing Maccarrone (2004)
CRH levels
↓ Activation of orexin Huang et al. (2007) ↓ activation of Ellis et al. (2006)
neurons in LH intracellular pathways
under control of the
orexin A receptor
↓ Activation of MCH Morrison and Berthoud
neurons in LH (2007)
↑ Hypothalamic Gamber et al. (2005) ↓ food intake equally Bermudez-Silva et al.
secretion of well in wild-type and (2012)
neuropeptide Y NPY null-mice

pups until weaning, but not in adults, when dams where under food in adipose tissue of obese women (Engeli et al., 2005). Incidentally,
restriction during gestation and/or lactation (Matias et al., 2003). human adipose tissue has been shown to possess a fully functional
In another study, 4 week-old mice fed an ␻-3 PUFA-deficient diet eCB system (Spoto et al., 2006). In addition, HFD induced an increase
showed higher brain levels of 2-AG when compared to the ␻-3 of AEA in liver, due to reduced degradation by FAAH (Osei-Hyiaman
PUFA-sufficient diet group, whereas a 6 weeks supplementation of et al., 2005). Another study reported higher eCB levels from mice
DHA-rich fish oil reduced brain 2-AG levels. These data suggested with diet-induced obesity, compared to lean mice, in epididymal fat
that in the developing mice manipulation of dietary ␻-3 PUFAs reg- and pancreas (Matias et al., 2006). In visceral adipose tissue, strong
ulates CNS physiological functions of eCBs (Watanabe et al., 2003). expression of FAAH and CB1 receptors was observed when com-
Consistently, a 2 weeks diet with DHA supplementation induced pared to subcutaneous adipose tissue, leading to the suggestion of
alterations in the eCB metabolome in adult mice plasma and brain, potential beneficial effects of CB1 blockade on metabolic syndrome
plasma being more responsive than brain to dietary manipulation (Bluher et al., 2006). In the same study, an increase in circulating
(Wood et al., 2010). It can be anticipated that longer times are 2-AG was reported in visceral fat of obese subjects (Bluher et al.,
required to affect brain eCBs, because FA composition of brain is 2006). Consistently, another investigation showed that only plasma
more stable than other tissues to changes in the diet (Lauritzen 2-AG levels, among all the eCBs measured, markedly increased
et al., 2001; Hulbert, 2003). Since adult humans do not have a high in subjects with intra-abdominal adiposity accumulation (Cota,
intake of fish oil for a prolonged time, it is unlikely that brain levels 2007).
of 2-AG and AEA could be affected. It was also reported that circulating eCBs resulted to be unaf-
Many studies have documented an eCB system overactivity in fected by isocaloric changes in dietary fat intake do not affect in
animals fed a high fat diet (HFD), as well as in obese patients lean and in obese subjects. However, HFD selectively reduced CB1
(Monteleone et al., 2005; Engeli et al., 2005; Osei-Hyiaman et al., and MAGL genes expression in skeletal muscle (Engeli et al., 2014).
2005; Matias et al., 2006; Bluher et al., 2006; Cota, 2007), at multiple Moreover, the same authors also observed the up-regulation of
levels: plasma, hypothalamus, organs with endocrine function or DAGL␣ mRNA and the down-regulation of FAAH and MAGL mRNAs
involved in energy expenditure or affected by the consequences of in obese subjects in subcutaneous adipose tissue (Engeli et al.,
metabolic disorders (e.g., heart and kidney) (Matias et al., 2008a,b). 2014). In the stomach of HFD mice, OEA levels increased, in parallel
In human studies, an increase in AEA levels was observed in the with an up-regulation of NAPE-PLD and a down-regulation of FAAH
blood of women with anorexia nervosa (AN) and binge eatind dis- gene expression (Aviello et al., 2008). Interestingly, alterations of
order (BED) (Monteleone et al., 2005), along with an increase of AEA eCB system elements were found in adipose tissue, pancreas, liver
and 2-AG levels and a reduction of FAAH and CB1 gene expression and skeletal muscle of HFD-induced obese mice, and thus eCB
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 209

system overactivation was supposed to contribute to the reduced liver, an effect that is possibly related to the observed diet-
nutrient oxidation and oxygen consumption that sustain obesity induced obesity via hepatic CB1 activation (Osei-Hyiaman et al.,
development in this model (Osei-Hyiaman et al., 2005; Pagotto 2005).
et al., 2006; Starowicz et al., 2008). Dietary guidelines recommend the shift from saturated to
As mentioned in the previous section, just as yet one clinical polyunsaturated fat, in order to reduce the incidence of cardio-
study investigated the regulation of eCB levels after hedonic eating vascular diseases (Harris et al., 2009). This recommendation has
(Monteleone et al., 2012). The authors observed that, in satiated led to increased consumption of vegetable oils. Soybean oil is one
normal-weight healthy subjects, consumption of food for pleasure, of the major dietary sources of linoleic acid (LA), an AA precur-
increased plasma levels of 2-AG and ghrelin, but not of AEA, OEA sor that is able to modulate eCBs biosynthesis (Salem et al., 1999).
and PEA (Monteleone et al., 2012). In agreement with these data, 2- Human and animal studies showed that high LA diets promote obe-
AG resulted increased in the hypothalamic reward system of mice sity (Massiera et al., 2003, 2010; Alvheim et al., 2012), and in mice
showing HFD preference (Higuchi et al., 2011). Later on, the same induce an increase of 2-AG and AEA levels (Alvheim et al., 2012).
research group showed that there are two different stages, induc- More recently, Alvehim and colleagues also observed that replac-
tion and maintenance, that involve a transient and a long-lasting ing fish oil with soyabean oil in feed for Atlantic salmon evokes
increase of 2-AG, respectively (Higuchi et al., 2012). In the cen- high dietary levels of LA and elevates AA, eCB activity and fat accu-
tral amygdala of rats during abstinence from palatable food, an mulation in the salmon liver. Moreover, mice consuming Atlantic
elevation of 2-AG and CB1 levels was observed, and was possi- salmon fed soyabean oil have higher weight gain, adipose tissue
bly needed to counteract the negative emotional state produced inflammation and liver levels of AA, AEA and 2-AG, when compared
by withdrawal (Blasio et al., 2013). Lower expression of CB1 and to mice fed fish oil salmon diet (Alvheim et al., 2013). In a recent
FAAH were observed in the hippocampus of Fischer rats, when investigation, rats receiving LA diet showed a 2-fold accumulation
compared to Wistar rats, the first strain expressing a lower reward of jejunal eCBs, as well as a strong preference to this type of food
sensitivity than the second toward a palatable food reward (Brand when compared to other unsaturated fats (DiPatrizio et al., 2013).
et al., 2012). CB1 mRNA levels were also found to be differentially All the studies described so far support a regulatory role for
regulated depending on the access to high palatable food: they eCB system in modulation of appetite, control of food consumption
increased in the nucleus of the solitary tract of rats with contin- and regulation of the hedonic aspects of eating (see summary in
uous access to highly palatable sweet for 6 weeks, and decreased Table 2). However, eCB system is also implicated in the homeostatic
in the cingulate cortex of rats with intermittent feeding schedules. regulation of eating, as described in the next section.
Moreover, CB1 binding densities were reduced in NAc shell of rats
with access to a highly palatable sweet-fat food (Bello et al., 2012). 4. Neurobiology of food addiction and food reward
Down-regulation of CB1 expression was observed also in the NAc of mechanisms
rats fed a highly palatable chow for 10 weeks (Harrold et al., 2002),
as well as in other rat brain regions (cingulate cortex, ventromedial The concept of food addiction has been suggested for the first
hypothalamic nucleus, descending/autonomic divisions of the par- time by Randolph in 1956 (Randolph, 1956), but only recently it
vocellular hypothalamic nucleus, ventrolateral parvocellular, and has received due attention, mainly because of its correlation to the
dorsomedial parvocellular) of animals fed a high palatable, high increasing rate of obesity; a condition that has reached pandemic
energy food for 13 weeks (Timofeeva et al., 2009). proportions, and that is increasing exponentially if considering that
Selective alterations of eCB levels have been reported in liver overweight/obese people are estimated to become ∼58% of the
and small intestine, but also in the brain, of adult rats fed 5 differ- world population by the year 2030 (Skidmore and Yarnell, 2004;
ent dietary fats for only 1 week (Artmann et al., 2008). For instance, Kelly et al., 2008).
DHA and eicosapentaenoic acid (EPA) administration reduced PEA Obesity is a complex multifactorial disease, not limited to
levels in rats (Artmann et al., 2008), even though PEA levels are genetic or environmental factors, nor to an unbalanced energy
not yet believed to be affected by nutrient intake, nor by starva- uptake and expenditure. Current knowledge supports a relation-
tion/refeeding protocols (Fu et al., 2007). ship with neurobiological as well as psychological aspects of
EPA/DHA supplementation induced a decrease in 2-AG (Banni overeating, including mood disturbances, altered reward percep-
et al., 2011; Batetta et al., 2009; Di Marzo et al., 2010) and AEA tion and motivation, and possibly addictive behavior (Jauch-Chara
(Batetta et al., 2009; Balvers et al., 2012; Wood et al., 2010) levels in and Oltmanns, 2014).
obese subjects (both humans and animals) at peripheral and central Hence obesity should no longer be viewed only as a metabolic
levels, paralleled by a reduction in NAPE-PLD, FAAH and CB2 gene disease but also as result of eating behavior disturbances charac-
expression (Hutchins et al., 2011). Overall, these data indicate a terized by a significant CNS component, with important impact not
reduction in eCB system signaling linked to decreased appetite and only on physical health but also on psychosocial functioning (APA,
food intake. 2013).
The primary FA in olive oil, oleic acid, is the precursor of Binge eating is a proto-typical eating-related maladaptive
OEA that, when administered orally as part of a HFD, increases behavior that may determine fluctuations in body weight and in
gene expression of FAAH (Thabuis et al., 2010) and PPAR␣ (Cluny some instance may cause obesity (Hudson et al., 2007; Swanson
et al., 2009). OEA levels are, instead, reduced by a ketogenic diet, et al., 2011). It also represents a central feature of bulimia nervosa
that is a HFD used for the treatment of refractory epilepsy in (BN) and BED and as described by the Diagnostic and Statistical
children (Freeman et al., 2007). Because of its effects on OEA Manual of Mental Disorders, Fifth Edition (DMS-5) (APA, 2013) it
and congeners, such a diet could also lead to overconsump- is a condition in which the subject consumes an amount of food
tion and obesity (Hansen et al., 2009). The role of OEA in the that is definitely larger than what most individuals would eat in
induction of satiety is well-established (Fu et al., 2007), and it a similar period and it is characterized by a subjective sense of
has been proposed that at the periphery, OEA increase due to loss of control over food consumption. Over the recent years it
NAPE-PLD overexpression, reduces across-meal satiety in rats has been recognized that BN and BED share important common-
(Fu et al., 2008). Consistently, high intake of any type of fats alities with substance abuse. In fact, like substance abuse, these
reduces intestinal levels of OEA and congeners (Artmann et al., eating disorders are characterized by an uncontrollable urge to
2008). At the peripheral level, it has also been observed that a seek the reward and to cosume it in excessive amounts despite the
HFD (60% energy) for 14 weeks increased AEA levels in mice perceived negative consequences (Gold et al., 2003). Additionally,
210 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

Table 2
Overview of studies examining the regulation of eCB levels by the diet.

Experimental system Diet ECS elements Reference

Newborn piglets and Milk supplemented with AA and ↑ eCBs Berger et al. (2001)
mice AA-rich diet
Mice ␻ 3-PUFA-deficient diet ↑ 2-AG Watanabe et al. (2003)
DHA-rich fish oil ↓ 2-AG
Adult mice Diet DHA-rich (for 6 weeks) Alteration eCB metabolome Wood et al. (2010)
CNS Mice HFD (414 kcal/100 g) ↑ 2-AG Higuchi et al. (2011,
2012)
Mice obese Diet supplemented with EPA/DHA ↓ AEA Balvers et al. (2012),
Wood et al. (2010)
Mice and Rats Highly palatable and high-energy food ↓ CB1 gene expression Timofeeva et al. (2009)
(13 weeks)
Rats Abstinence from palatable food ↑ 2-AG Blasio et al. (2013)
↑CB1 gene expression
Rats Continuous access to highly palatable ↑ CB1 gene expression Bello et al. (2012)
sweet-fat food for 6 weeks) ↓ CB1 gene expression
Intermittent feeding schedules ↓ CB1 gene expression
Highly palatable sweet-fat food
Rats Highly palatable chow (10 weeks) ↓ CB1 binding Harrold et al. (2002)
Adult rats 5 different dietary fats (1 weeks) ↓ PEA Artmann et al. (2008)
↓ OEA
Obese Zucker rats Diet supplemented with EPA/DHA ↓ 2-AG Di Marzo et al. (2010)
Women with AN and HFD ↑ AEA Monteleone et al.
BED (2005)
Obese women HFD ↑ AEA and 2-AG Engeli et al. (2005)
↓ FAAH and CB1 gene
expression
Obese humans Diet supplemented with EPA/DHA ↓ 2-AG Banni et al. (2011)
Adult mice Diet DHA-rich (for 6 weeks) Alteration eCB metabolome Wood et al. (2010)
Pheripheral
Mice HFD (33.5% fat) ↑ AEA Osei-Hyiaman et al.
↓ FAAH activity (2005)
Mice Diet-induced obesity ↑ eCB Matias et al. (2006)
Mice HFD (25.5% fat) ↑ OEA Aviello et al. (2008)
↑ NAPE-PLD and ↓ FAAH
activity
Mice obese Diet supplemented with EPA/DHA ↓ AEA Balvers et al. (2012);
Wood et al. (2010)
Mice Diet supplemented with EPA/DHA ↓ NAPE-PLD, FAAH and CB2 Hutchins et al. (2011)
gene expression
Mice HFD enriched with oleic acid (the ↑ FAAH and PPAR␣ gene Thabuis et al. (2010)
precursor of OEA) expression Cluny et al. (2009)
Mice Diet supplemented with Atlantic ↑ AEA, 2-AG and AA Alvheim et al. (2013)
salmon fed a high LA diet
Adult rats 5 different dietary fats (1 weeks) ↓ PEA Artmann et al. (2008)
↓ OEA
Obese Zucker rats Diet supplemented with EPA/DHA ↓ 2-AG Batetta et al. (2009)
Rats LA diet ↑ eCB DiPatrizio et al. (2013)

binge eating episodes usually occur in secrecy, involve intense crav- these include naloxone (Boggiano et al., 2005), naltrexone, baclofen
ings associated with stress and sense of guilt (Jastreboff et al., 2013; (Buda-Levin et al., 2005; Corwin et al., 2012), topiramate (Cifani
McAleavey, 2001), and are associated with high rate of recidivism et al., 2009), Rhodiola rosea and Hypericum perforatum extracts
(Bergh et al., 2002; Fairburn et al., 2000). (Cifani et al., 2010; Micioni Di Bonaventura et al., 2012a), adeno-
Unlike drugs of abuse, food is necessary for survival and humans sine A2A receptor agonists (Micioni Di Bonaventura et al., 2012b),
have developed evolutionary mechanisms to select and prefer orexin 1 receptor antagonists (Piccoli et al., 2012) and CRF1 recep-
calory reach highly palatable food. The constant and easy access tor antagonists (Micioni Di Bonaventura et al., 2014). The effecicacy
to these foods leads to their unnatural consumption that is no of some of these drugs on binge eating have been described also in
longer aimed at maintaining health and satiety but rather to clinical studies (Broft et al., 2007; Meyer, 2008; Arbaizar et al., 2008;
achieve reward. Moreover, refined carbohydrate, fat and salt rich McElroy et al., 2004).
of highly palatable food has an important impact on the regulation In addition to these pharmacological data, molecular and neu-
of mood and some individuals may overconsume these nutrients rochemical evidence suggest that overconsumption of palatable
to self-medicate from negative affective conditions such as anxi- foods is accompained by the stimulation of brain reward dopa-
ety, depression, mental fatigue (Ifland et al., 2009). In this respect minergic and opioid systems. Notably, in binge eaters and in
highly palatable foods may evoke alterations in the CNS in ways the obese population these systems are subjected to a profound
similar to drugs of abuse and reinforce behaviors with mechanisms readaptation leading to a state of reward hyposensitivity and to
resembling addiction (Spring et al., 2008; Gearhardt et al., 2011a,b; the development of compulsive-like eating (Volkow et al., 2009a;
Kenny, 2011), and through regulation of common neurobiological Avena et al., 2008b; Kenny et al., 2013; Thamotharan et al.,
substrates (Avena et al., 2008a; Oswald et al., 2011; Rossetti et al., 2013). Evidence also suggests that the reward deficit observed
2013). in obese patients and in rodents fed with high fat food may be
This view is indirectly supported by pharmacological evidence linked to dysregulation of striatal D2 receptors signaling (Volkow
showing that binge eating and excessive food consumption are et al., 2008; De Weijer et al., 2011; Geiger et al., 2009; Johnson
attenuated by drugs that influence also the intake of drugs of abuse; and Kenny, 2010). In overweight rats it has been proposed that
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 211

Fig. 3. Main features of food addiction.

reward deficit, associated with blunted D2 mediated activity may levels of DA receptors are related to both obesity and drug depend-
reflect counteradaptive decreases in the baseline sensitivity of ence (Volkow et al., 2009a).
brain reward circuitries to oppose their overstimulation by palat- Furthermore, individuals with a variety of eating disturbances
able food (Johnson and Kenny, 2010). In another study, Alsio report craving and symptoms they percieve as withdrawal, and
et al. (2010) showed that D1 and D2 receptors expression in the escalating patterns of eating that might be viewed as evidence
nucleus accumbens (NAc) is downregulated in obesity-prone (OP) for tolerance and dependence (Pelchat, 2002). Likewise, binge
rats, in comparison to obesity-resistant rats, under high-fat and eating patients report withdrawal-like feelings of panic and dis-
high-sugar diets. These findings strongly involve both dopamine tress and a sense of abstinence when their binge eating behaviors
(DA) and opioid receptors in the OP phenotype, but they also were delayed (Huebner, 1993). In other studies, removal of certain
suggest longer term changes for the DA circuitry. Interestingly, foods from diet triggered some symptoms of withdrawal such as
similar disruption occurs after intravenous cocaine or heroin self- headaches, sweats, irritability, and panic (McAleavey, 2001). Rats
administration (Ahmed et al., 2002; Markou and Koob, 1991; Kenny fed sucrose solution and lab chow on a schedule that induced bing-
et al., 2006), and supports other studies suggesting that obesity ing, after several weeks exhibited behaviors similar to those found
and drug addiction share neuroadaptive responses in brain reward in addicted individuals, like symptoms resembling tolerance, with-
circuitries, to alleviate the persistent state of reduced reward drawal, and craving (Avena et al., 2008b). Accordingly, Parylak
(Stice et al., 2008; Teegarden and Bale, 2007; Avena et al., 2008a; et al. (2011) suggested that the “side” of addiction, described by
Volkow et al., 2006; Volkow and Wise, 2005; Volkow and O’Brien, Koob and Le Moal (2005) and Le Moal and Koob (2007) as devel-
2007). opment of the aversive emotional state that drives the negative
Human brain imaging data showed compromised striatal DA reinforcement of addiction, occurred also in food addiction under
signaling, particularly in subjects carrying genetic polymorphisms certain instances of obesity or eating disorders. According to this
associated with long-term weight gain (Stice et al., 2008). Imag- view, like for drug addiction (Koob and Le Moal, 1997), stress
ing studies also showed that increased DA in addicted subjects and negative moods (depression, anxiety) can trigger compulsive
predicts the intensity of cue-induced cravings (Volkow et al., food intake in humans, through a three-stage cycle characterized
2006) and similarly, in obese individuals with BED, higher DA by binge/intoxication, withdrawal/negative affect and preoccupa-
release upon exposure to food cues was associated with the sever- tion/anticipation.
ity of the disorder (Wang et al., 2011). Functional neuroimaging Although excessive consumption of different nutrients can
has also demonstrated in obese individuals that pleasant looking, induce behaviors associated with addiction, caution is still neces-
smelling, and tasting foods have reinforcing properties similar to sary before drawing conclusions. Many groups, first of all those led
drugs of abuse (Zhang et al., 2011). Positron emission tomographic by the pioneers Bartley G. Hoebel (Hoebel and Teitelbaum, 1962)
(PET) imaging studies have further provided evidence that reduced and Nicole M. Avena (Avena et al., 2008a), have interrogated the
212 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

Fig. 4. Interactions between the endocannabinoid, opioid and dopaminergic systems in the mesolimbic reward circuit. VTA, ventral tegmental area; NAc, Nucleus Accumbens;
AM, amygdala; FC, Frontal cortices; GABA, GABAergic neurons, GLU, glutamatergic neurons; DA, dopaminergic neurons; Opioid; Opioid neurons; CB1 , type 1 cannabinoid
receptors.

validity of the food addiction construct (Avena et al., 2011, 2012) and depression, emotion dysregulation, eating disorder psychopathol-
investigated different angles of this phenomenon also by dissecting ogy and lower self-esteem it has been observed, when compared
the neurobiology of appetite from hedonic overeating (Bocarsly to subjects that were not diagnosed with food addiction (Gearhardt
et al., 2011; Avena and Gold, 2011). et al., 2011b). Similarly, Davis et al. (2011) showed that food addic-
It has been noted that food addiction behavior is related to the tion was linked to higher levels of depression and trait impulsivity
nutritional composition of the food consumed under certain envi- in obese adults. More recently, Gearhardt et al. (2013) indicated
ronmental conditions (Corwin, 2011). Rats with a fat-rich food did that ∼42% of obese patients with BED met the food addiction
not show signs of opiate-like withdrawal when precipitated by the threshold, whereas Pedram et al. (2013) reported that ∼80–89%
opiate antagonist naloxone or during fasting, as previously shown of subjects, recruited in a population of food-addicted individuals,
with sugar (Avena et al., 2008b; Colantuoni et al., 2002), suggesting were overweight/obese, further providing evidence that food addic-
that the brain opioid system can be differently affected by overeat- tion contributes to obesity. Finally, in patients undergoing weight
ing fat foods versus sugar. Moreover, neither obesity (Blundell and loss surgery, more than 40% met the criteria for food addiction
Finlayson, 2011) nor BED (Gearhardt et al., 2011b) should be con- (Meule, 2011).
sidered synonymous of food addiction, that is neither necessary nor Although the concept of food addiction remains highly debated
sufficient for obesity. Indeed, among adults with BED, the preva- (Ziauddeen et al., 2012; Salamone and Correa, 2013), these find-
lence of obesity is 42%, that is only ∼8% higher than that seen in the ings call for additional investigations to identify clinically relevant
general population (Wonderlich et al., 2009). subgroups within obesity or eating disorder, that show features (or
In order to further validate the construct of food addiction, and symptoms) described in Fig. 3. The ultimate goal is to define more
to explore its prevalence, the Yale Food Addiction Scale (YFAS) personalized prevention and treatment strategies.
has been designed in college students and binge eaters (Gearhardt The proposal, originally made by Holden (2001, 2010), and then
et al., 2009), and then used in eating disorder patients (Gearhardt the inclusion of food addiction in the DSM-V strongly suggest that
et al., 2011b) and obese subjects (Davis et al., 2011). YFAS is a the status of some individuals with eating disorders resembles that
systematic tool to distinguish between individuals with or with- typically endorsed for individuals with substance use disorders.
out addictive-like eating patterns. This scale may be a useful tool This observation might reflect the engagement of the same neu-
for the detection of “food addiction” in humans, using the cri- ral systems, including those implicated in regulatory self-control
teria identified for substance dependence in the fifth edition of and reward (APA, 2013).
the Diagnostic and Statistical Manual of Mental Disorders (DSM- On a final note, it should be recalled that human and animal
V). studies are pivotal to further our understanding of the factors that
In BED patients with high prevalence of food addiction (51%), contribute to the escalating rates of obesity, and to develop effective
a relationship between diagnosis of this trait and more severe approaches not only to restrain hunger and eating behavior, but also
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 213

to study the next generation of obesity treatments, as suggested by the dorsal and ventral striatum, and that these increases are asso-
Avena et al. (2013). Among potential neuropathways and neuro- ciated with the subjective rating of drug reward (Bossong et al.,
transmitters involved in the reinforcement and reward associated 2009).
with food intake, this review focuses on the eCB system. Moreover, in a study on healthy voluntiers it has been shown
that treatment with rimonabant reduces striatal brain activity dur-
ing reward processing (Horder et al., 2010).
5. Cannabinoids and reward mechanisms THC, is the major psychoactive ingredient of cannabis and is
the main responsible of cannabis effects on the reward system
eCBs affect the release of various neurotransmitters, including (Pertwee, 2008).
GABA (␥-aminobutyric acid), glutamate, and DA (Pertwee, 2008; For example in Wistar rats it has been shown that THC stimu-
Lüscher and Malenka, 2011; Szabo et al., 2002) and plays an lates reward as measured by either conditioned place preference
important function in the neuroplasticity associated with chronic or intracranial operant self-administration in a free-choice proce-
exposure to drug of abuse (Fig. 4). dure, yet only within a limited dose range (Braida et al., 2004).
High density of eCB receptors is found in several regions of the Squirrel monkeys lever-press for intravenous injections of THC
basal ganglia (Ameri et al., 1999; Gardner, 2005) and their impor- (Justinova et al., 2005). Moreover, THC decreases electrical brain-
tant role in the regulation of the cortico-mesolimbic DA system stimulation reward thresholds (Gardner, 2005), and increases
that originates from the ventral tegmental area (VTA), and project the firing rate of VTA dopaminergic neurons that project to the
to the NAc, the amygdala and the frontal and limbic cortices has NAc (French et al., 1997; Gifford et al., 1997). A PET study in
been described (Solinas et al., 2006; Oleson et al., 2014; Massi et al., humans in which glucose metabolism was monitored it has been
2008; Melis et al., 2004, 2012). shown that, after chronic THC, marijuana abusers have increased
For instance, eCBs modulate the glutamatergic excitatory and orbitofrontal cortex, medial superior prefrontal cortex and striatum
GABAergic inhibitory synaptic inputs into the dopaminergic neu- metabolic activation compared to controls, whereas cerebellar
rons of the VTA and the glutamate transmission in the NAc, acting metabolism resulted increased in both abusers and controls. Also
as retrograde messengers on CB1 . The latter is the primary tar- tactile marijuana-related cues versus neutral cues were shown
get of eCBs to control emotional reactivity, motivated behaviors to increase fMRI activation in VTA, supporting the involvement
and energy homeostasis (Fride, 2002; Piomelli, 2005; Steiner and of striatal networks in abstinent marijuana users (Filbey et al.,
Wotjak, 2008). 2009). Finally, 69% of the abnormal activation clusters in studies
The activation of CB1 , present on axon terminals of GABAergic on marijuana effects on brain function were located in regions
neurons in the VTA, inhibits GABA transmission and removing this functionally connected to the striatum (Tomasi and Volkow, 2013).
inhibitory input on DA neurons, it leads to an increase in the fir- These findings suggest that striatal networks are involved in mari-
ing patterns of these cells (Riegel and Lupica, 2004). With similar juana addiction (Volkow et al., 1996).
mechanism, the activation of CB1 decrease excitatory glutamater- Besides DA, other neurochemicals reciprocally modulate
gic transmission in the VTA and NAc, mainly regulating the activity cannabinoid-induced reward, like noradrenaline, serotonine,
of neurons projecting from the prefrontal cortex (Melis et al., 2004). acetylcholine, adenosine and opioids (for a review, see Maldonado
Therefore, the final effect of eCBs on the modulation of DA activity et al., 2011). The best studied cross-talk occurs with the brain opi-
depends on the functional balance between the inhibitory GABAer- oid peptide system (Gardner, 1997, 2000; Robledo et al., 2008;
gic and excitatory glutamatergic inputs to the VTA, with the latter Parolaro et al., 2010). For example, it has been also shown that
that predominates (Maldonado et al., 2006). Hence, eCBs do not ␦-opioid (DOP) and CB1 receptor agonists synergistically activate
directly depolarize DA neurons but acting indirectly via pre- and the mesolimbic DA system, increasing intracellular cAMP and pro-
post-synaptic inhibition of interneurons (Szabo et al., 2002), they moting protein kinase A (PKA)-dependent C␣ translocation (Yao
may ultimately activate mesolimbic DA transmission (Gessa et al., et al., 2003). Moreover, in a microdialysis study in the rat it has
1998). been shown that THC stimulates DA release in the shell of the NAc
Although DA is not a direct measure of reward, it is considered an through mechanisms involving VTA MOP receptors (Tanda et al.,
important neurochemical correlate of rewarding effects of abused 1997). Noteworthy the interaction the opioid systems seem to be
drugs, and a potential mechanism for these effects. All addictive important not only for the reinforcing and rewarding effects of
drugs display ability to increase DA, particularly in the NAc, under- cannabinoids (Tanda et al., 1997) but also for the expression of
ling their rewarding effects (Koob, 1992; Wise, 2004). Like for other withdrawal (Navarro et al., 1998; for reviews, see Justinova et al.,
addictive drugs, animal (Gardner et al., 2005; Solinas et al., 2007) 2005; Scavone et al., 2013).
and human (Glass et al., 1997; Terry et al., 2009; Van Hell et al., More recently, the hypocretin (Flores et al., 2013) has also
2012) studies have indicated that cannabinoids affect brain reward been linked to cannabinoid-induced reward. The link between
processes and reward-related behaviors, through their ability to cannabinoid system with this neuropeptidergic circuitry playing a
enhance extracellular DA levels in the NAc (Fadda et al., 2006; Lecca central role in caloric homeostasis, food consumption and reward
et al., 2006). processing (Martin-Fardon et al., 2010; Aston-Jones et al., 2010;
For instance, Gardner et al. (2005) were the first to report Piccoli et al., 2012) may be suggestive of potential mechanisms
that cannabinoids enhance extracellular DA overflow in brain through wich the eCB system may impact on food reward and may
reward axon terminal loci (Ng Cheong Ton and Gardner, 1986); regulate its non-homeostatic consumption (Di Marzo et al., 2009).
whereas direct agonism at CB1 leads to rewarding effects in ani- Foods that are rich in sugars and fat are potent reward stimuli,
mals (Gardner, 2005; Solinas et al., 2007). Consistent with a role and can promote over-eating (Lenoir et al., 2007). As discussed
of CB1 in reward modulation other studies have shown that recep- above, they activate brain reward circuits with release of DA, eCBs
tor antagonists like rimonabant reduce the reinforcing effects of all and opiates, while inhibiting that of satiety mediators. Indeed,
major drugs of abuse (Economidou et al., 2006; De Vries et al., 2003; limbic levels of DA and eCBs correlate with craving for tasty
Cohen et al., 2002). Reduction in substance of abuse self adminis- food, indicating their involvement in reward-driven palatable food
tration has been also reported following the selective inhibition of eating behavior (Tanda and Goldberg, 2003; DiPatrizio, 2014). Con-
the AEA transporter (Gamaleddin et al., 2013; Cippitelli et al., 2007). sistent with a role of cannabinoid neurotransmission in food reward
In line with preclinical evidence, human PET studies have shown pharmacological data with rimonabant have shown that its admin-
that like other addictive substances, cannabis increase DA levels in istration leads to a reduction in the conditioned place preference for
214 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

palatable food, such as sucrose or chocolate (Chaperon et al., 1998). studies on AEA levels, these observations were made on patients
Moreover, they dose-dependently reduce the motivation for food who were classified as recovered from the pathology, thus possibly
in a progressive ratio schedule of food self-administration in rats excluding the limit that the altered nutritional status might have
(Gallate and Mc Gregor, 1999). CB1 antagonism has been shown also affected the plasma levels of the investigated lipids (Gaetani et al.,
to counteract the increase of extracellular DA release, induced by 2008).
novel and highly palatable foods (Melis and Pistis, 2007); while CB1 Higher blood levels of CB1 mRNA were detected in women suf-
agonists do the opposite (Higgs et al., 2005; Solinas and Goldberg, fering from AN and BN (Frieling et al., 2009), whereas decreased
2005; Wakley and Rasmussen, 2009; Ward and Dykstra, 2005). For peripheral CB1 mRNA expression was observed in patients with
instance, THC increases sucrose-induced hedonic activity and DA more severe forms of the disorders, such as those with self-
release into the NAc (De Luca et al., 2012). Furthermore, CB1 and injurious behaviors (Schroeder et al., 2012). Further observations
MOP blockade synergize to reduce food intake and body weight on altered CB1 expression came from PET studies, demonstrating
in rodents (Lockie et al., 2011; Tallett et al., 2009). Accordingly, increased receptor levels in the insula and inferior frontal and tem-
CB1 knock-out mice show a reduced motivation to work for food, poral cortex of underweight AN patients, as well as in the insula
compared to wild-type mice (Sanchis-Segura et al., 2004). In addi- of symptomatic BN subjects (Gérard et al., 2011). Finally, human
tion, intra-NAc injections of AEA enhance reward associated with genetic studies reported a positive association between eating dis-
food palatability (Mahler et al., 2007). Thus, CB1 appears to be an orders and specific polymorphisms of genes encoding for different
important component of the neural substrate that mediates the components of the eCB system, such as CB1 (Siegfried et al., 2004;
reinforcing and motivational properties of a highly palatable food Monteleone et al., 2008), CB2 (Ishiguro et al., 2010) and FAAH
(Maccioni et al., 2008). In line with this, CB1 mRNA expression (Monteleone et al., 2009), although previous contrasting findings
is down-regulated in areas of the limbic forebrain of rats fed a showing no association were also reported (Muller et al., 2008).
palatable diet (Harrold et al., 2002), an effect that can be consid- It has been hypothesized that the deregulated eCB tone of AN
ered as part of a compensatory mechanism aimed at counteracting and BN patients may represent an adaptive response aimed at
increased levels of eCBs resulting from the consumption of a fat- maintaining energy balance by potentiating internal orexigenic
rich palatable food. On the other hand it has been also shown that signals, and hence stimulating food ingestion. Yet, it might also con-
fasting leads to an increase of limbic of AEA and 2-AG levels, that tribute to facilitate the rewarding properties of the aberrant eating
return to normal after refeeding. This effect occurs selectively in behaviors (Monteleone and Maj, 2013).
brain areas that are not involved in the regulation of feeding behav- Given the appetite-stimulating and antiemetic properties of
ior (Kirkham et al., 2002). This overactivation of the eCBs systems THC, research has focused on its role in patients with AN. How-
during fasting may be part of a physiological mechanism aimed at ever, two small trials exploring the effects of THC in underweight
enhancing the motivation for food. Furthermore, injection of 2-AG AN patients did not give the expected positive results (Gross et al.,
directly in the NAc shell, which shows high CB1 expression, trigg- 1983; Berry, 2006). In particular, five years before the discovery of
ers increased feeding also in satiated rodents, via CB1 activation CB1 receptors, Gross et al. (1983) evaluated the effects of synthetic
(Kirkham et al., 2002). THC treatment in 11 patients with primary AN, that caused a 25%
Interestingly, in healthy volunteers consumption of favorite loss of their body weight. The study was conducted in a 4-week
food, as compared to normal food, is accompanied by elevated crossover trial, testing quite high doses of THC (from 7.5 mg/day
plasma 2-AG levels, which correlate positively with plasma ghrelin to a maximum dose of 30 mg/day), against diazepam as an active
levels (Monteleone et al., 2012). Altogether these evidence demon- placebo. Moreover, the medication was discontinued during the
strate that the eCBs system plays a major role in regulating weekend, possibly to avoid addiction. The daily caloric intake was
reward related mechanisms in response to not only substances of controlled, and occasionally tube feeding was used. THC treatment
abuse but also calory reach highly palatable food. This strongly corresponded to a slightly higher weight gain (about 1 kg), as com-
supports its role in the regulation of food-related addictive-like pared to that observed during diazepam treatment; three patients
behaviors. (27%) withdrew after experiencing severe dysphoric reactions and
the authors concluded that THC was not an effective drug for the
treatment of primary AN. This study raised several concerns, mainly
6. eCB system-based drugs in eating disorders and obesity due to the use of high doses of THC and of diazepam as a con-
trol, despite the evidence from both animal (Naruse et al., 1991)
6.1. eCB system in eating disorders and human (Hein and Huyser, 2010) studies that benzodiazepines
may also increase food intake. Several years later a pilot study was
The widespread role of the eCB system as a modulator of both conducted by Berry (2006) in 9 ambulatory AN patients treated
homeostatic and hedonic aspects of eating prompted intensive with THC, who did not show any effect on body weight gain, but a
investigations into possible defects of this system in eating disor- significant improvement of depression and perfectionism scores.
ders. In particular, increased blood levels of AEA (with unaltered A very recent publication reported the results of an add-on,
levels of 2-AG) were found in patients with AN or BE disorder, prospective, randomized, double blind, controlled cross-over study,
but not in women affected by BN (Monteleone et al., 2005). In aimed at testing the effects of dronabinol (a synthetic THC for oral
both healthy controls and AN women, blood AEA levels appeared administration, 2.5 mg twice a day for 4 weeks) on body weight
inversely correlated with plasma leptin concentrations. Since a gain, as well as on the Eating Disorder Inventory-2 (EDI-2) score
negative modulation of leptin on eCB mobilization was previously and leptin plasma levels (Andries et al., 2014). The results of this
demonstrated (Di Marzo et al., 2001), this observation suggested study demonstrated that dronabinol therapy induced a small yet
that the decreased leptin signaling of underweight AN patients significant weight gain in the absence of severe adverse events.
could be at least in part involved in the increase of AEA lev- Conversely, no effects were noticed on EDI-2 score change dur-
els. A subsequent study demonstrated that AEA is not the only ing treatment, while leptin levels rose to 15% above placebo, but
acylethanolamide altered in eating disorders. Indeed, also OEA was returned to baseline after the last dose of dronabinol. In addition,
associated with eating disorders. In particular, both plasma and weight records collected up to one year after the end of the trial
cerebrospinal fluid levels of this endogenous lipid were increased indicated a persistent improvement of patient nutritional status
in bulimic-anorexic women, with respect to restrictor anorexic and without signs of addiction or withdrawal, suggesting that dron-
control women (Gaetani et al., 2008). Differently from the previous abinol effects were longlasting and safe in these patients, thus
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 215

encouraging more research on the positive effects of cannabinoids Interestingly, body weight loss in these patients after a 12-weeks
in eating disorders. lifestyle program, significantly decreased salivary AEA levels. Based
A certain degree of positive effects induced by THC treat- on their findings, the authors suggest that salivary AEA might rep-
ment was also reported in a recent preclinical study that used resent a useful biomarker in obesity.
the activity-based anorexia (ABA) model of AN. In this model, Similar alterations in the blood and adipose tissues of obese sub-
THC treatment attenuated the weight loss associated with the jects and in different animal models of obesity were also reported
development of ABA, by increasing high-palatable food intake and for other eCB-related acylethanolamides, such as OEA and PEA.
moderately decreasing running wheel activity (Verty et al., 2011). In particular, the latter compounds increased in the blood and
The stimulation of eCB system actiivity has been clearly shown to be subcutaneous adipose tissue of subjects with both obesity and
of success in diseases in which weight gain is needed (i.e., cachexia type 2 diabetes (Annuzzi et al., 2010), and in the saliva of obese
in AIDS patients), however only a few studies claim so far the effi- subjects (Matias et al., 2012). Conversely, the intestinal levels of
cacy in ED and the development of new clinical trials will be of OEA appeared to follow an opposite deregulation in obese ani-
relevance (Marco et al., 2012). mals, being down-regulated by the prolonged exposure to HFD. The
latter observation raised the possibility that excessive fat consump-
6.2. eCB system in obesity tion might promote overeating, by suppressing (at least in part)
the satiating effects of gut-derived OEA (Hansen et al., 2012; Diep
Several preclinical and clinical observations (reviewed by et al., 2011; Artmann et al., 2008; Tellez et al., 2013; Aviello et al.,
Matias and Di Marzo, 2007; Bermudez-Silva et al., 2010) sug- 2008).
gest that there is a close positive association between obesity Based on the observations of a pathological overactivation of the
and the hyperactivity of the eCB system manifested as either/both eCB system in overweight and obese subjects, restoring a normal
an overproduction of eCBs or/and an upregulation of cannabi- eCB tone by drugs that interfere with eCB signaling was considered
noid receptors in central and peripheral tissues involved in energy as a potential therapeutic approach to arrest both the development
homeostasis. At the central level, the first evidence for the hyper- and the maintenance of obesity and obesity-related co-morbidities.
activity of the eCB system came from studies on genetically obese Initial preclinical studies conducted with rimonabant found evi-
mice (ob/ob), whose hypothalamic eCBs levels were higher with dence to support this hypothesis (Cota et al., 2003).
respect to lean animals, and were significantly reduced by the Several clinical trials followed the initial experimental findings
exogenous administration of leptin (Di Marzo et al., 2001). Sub- and they were grouped in the four rimonabant in Obesity (RIO)
sequently, an elevated CB1 expression was observed in the white studies (Van Gaal et al., 2005; Després et al., 2005; Pi-Sunyer et al.,
adipose tissue of another rodent model of obesity, caused by 2006; Scheen et al., 2006).
defective leptin signaling as compared to lean controls (Bensaid The very promising therapeutic benefits of rimonabant seemed
et al., 2003). An up-regulation of CB1 was observed also dur- to extend even beyond obesity and obesity co-morbidities, as
ing the differentiation of mouse 3T3-F442A (Bensaid et al., 2003) it turned to be efficacious also for addictive disorders such as
and 3T3-L1 (Gasperi et al., 2007) fibroblasts into adipocytes, and alcoholism (Gutierrez-Lopez et al., 2010; Getachew et al., 2011;
increased 2-AG levels were found in mature and hypertrophic Economidou et al., 2006), and nicotine dependence (Rigotti et al.,
adipocytes, as well as in the epididymal fat pads of obese animals 2009; Diergaarde et al., 2008; Shoaib, 2008; Bhatti et al., 2009).
compared with those of lean mice (Matias et al., 2006). Remark- The latter aspect was considered particularly interesting because
ably, in differentiated 3T3-L1 adipocytes AEA increased by ∼2-fold of its negative effect on weight gain (that is often associated with
insulin-stimulated glucose uptake, according to a CB1 -dependent smoking cessation).
mechanism that involves nitric oxide synthase (Gasperi et al., Rimonabant was approved and introduced into the market as
2007). Elevated levels of both 2-AG and AEA were detected in pan- an anti-obesity agent in several countries, including the European
creatic ␤-cells cultured under high glucose conditions, and were Union. However, the USA Food and Drug Adminstration (FDA) asked
further increased by insulin application, suggesting that hyper- further evidence about the safety of rimonabant before approving
glycemia and insulinemia may cause an overactivation of the eCB its marketing in the United States. Even though psychiatric disor-
system in obesity-related type 2 diabetes (Matias et al., 2007). Dys- ders were used as exclusion criteria, a meta-analysis of these four
regulation of the eCB system in obese human subjects has also RIO studies suggested that patients treated with Rimonabant were
been reported, and the first evidence came from studies in over- 2.5- and 3.0-times more likely to experience psychiatric adverse
weight/obese women with a BED (Monteleone et al., 2005) and in effects, such as anxiety and depression or depressive symptoms,
obese postmenopausal women (Engeli et al., 2005). In clinical stud- than patients receiving placebo (Christensen et al., 2007).
ies obesity-related elevations of eCBs were coupled to a decreased Further investigations were conducted on rimonabant, in which
activity of FAAH, and consistently specific polymorphisms of the psychiatric disorders were not regarded as exclusion criteria
faah gene were associated with overweight and obesity (Sipe et al., (Rosenstock et al., 2008; Rigotti et al., 2009; Nissen et al., 2008;
2005) or with lower insulin sensitivity (de Luis et al., 2010a,b; de Després et al., 2009). The authors showed that adverse psychiatric
Luis et al., 2012). effects were more pronounced in patients treated with rimon-
When obese subjects were classified on the basis of fat distri- abant versus placebo controls. The European Union Committee
bution by Blüher et al. (2006), higher levels of 2-AG in adipose for Medicinal Products for Human Use definitely concluded that
tissue samples were found in patients with elevated abdominal fat rimonabant doubled the risk of psychiatric disorders, and the Euro-
distribution, as compared to those observed in patients with sub- pean Medicines Agency (EMA) suspended the license of this drug.
cutaneous fat or in lean controls. Such an increase was shown to Sanofi-Aventis withdrew the product from sale worldwide, and also
positively correlate with major cardiometabolic risk factors, such as stopped clinical development.
visceral fat mass and fasting plasma insulin levels, and to correlate These events profoundly influenced the pharmaceutical indus-
negatively with glucose infusion rates during the clamp procedure, try. The development of several other CB1 antagonists that were
a method for assessing insulin sensitivity in humans (Blüher et al., being tested in clinical trials, such as otenabant (Pfizer), ibipin-
2006; Côté et al., 2007). abant (Solvay/Bristol-Myers Squibb), surinabant (Sanofi-Aventis),
A recent report by Matias et al. (2012) has shown that AEA lev- and taranabant (Merck) were interrupted and their further devel-
els increased in the saliva of obese subjects, and directly correlate opment was not supported because of adverse psychiatric side
with Body Mass Index, waist circumference and fasting insulin. effects (Addy et al., 2008a,b; Kipnes et al., 2010).
216 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

Table 3
modulators of the eCB system for the treatment of obesity.

Drug Mechanism Reduction of Reduction of Reduction References


food intake body weight of liver fat

Synthetic cannabinoid antagonist/inverse agonist


AM6545 Selective CB1 Yes Yes Yes Tam et al. (2012)
peripheral antagonist
LH21 Peripheral CB1 Yes Yes No Alonso et al. (2012),
antagonist Pavón et al. (2008)
JD5037 CB1 antagonist/inverse Yes Yes Yes Tam et al. (2012)
agonist
URB447 Mixed CB1 Yes Yes Not yet LoVerme et al. (2009)
antagonist/CB2 agonist assessed
PSNCBAM-1 CB1 allosteric Yes Yes Not yet Horswill et al. (2007)
antagonist assessed

Polyphenols
Catechins Antagonists/inverse Yes Yes Yes Korte et al. (2009,
(epigallocatechin-3-O- agonists of human 2010) and Bose et al.
gallate cannabinoid receptors (2008)
(EGCG))

Inhibitors of 2-AG biosynthesis


O-5596 di- acylglycerol lipase Yes Not yet Not yet Bisogno et al. (2009,
(DAGL) inhibitors assessed assessed 2012)

Compounds alternative to cannabinoid receptor antagonists


OEA PPAR-␣ Yes Yes Yes Piomelli (2013)

New neutral CB1 antagonists and/or peripherally restricted CB1 Among the peripherally targeted CB1 antagonists, only LH-21
antagonists or inverse agonists unable to cross the blood–brain has been investigated with regards to psycho-behavioral effects
barrier (BBB) have been developed, in the attempt to preserve (Griebel et al., 2005). Therefore, the potential adverse behavioral
rimonabant positive effects and to avoid the unwanted psy- effects of these new peripherally targeted CB1 antagonists should
chiatric side-effects (for a review, see Silvestri and Di Marzo, be addressed in future studies.
2012). Many of these new compounds have already provided Finally, it is noteworthy that some natural dietary antioxidant
very promising results at the preclinical level (Tam et al., 2010, polyphenols display significant affinity for human cannabinoid
2012) and are reported schematically in Table 3. Examples of receptors, acting as antagonists/inverse agonists (Gertsch et al.,
this kind of drugs are AM4113, AM6545, LH-21, JD5037, or 2008; Korte et al., 2009, 2010).
URB447 (for a review, see O’Keefe et al., 2014; Bermudez-Silva Blockade of CB1 may induce a feedback loop that drives eCB
et al., 2010). AM4113 (N-piperidin-1-yl-2,4-dichlorophenyl-1 levels even higher than those derived from the activation of other
h-pyrazole-3-carboxamide analog) in Sprague-Dawley rats fed cannabinoid receptors (Bermudez-Silva et al., 2010). Moreover,
either a normal diet or a diet high in carbohydrates or HFD, pharmacological modulation of eCBs synthesis, rather than CB1
reinforced positive appetite control crossing the BBB and binding blockade, may provide a more physiological approach to treat
to neural CB1 (Sink et al., 2008). AM6545 (5-(4-(4-cyanobut- obesity, since excessive eCB levels, rather than excessive CB1
1-ynyl)phenyl)-1-(2,4-dichlorophenyl)-4-methyl-N-(1,1-dioxo- expression, appear to be the primary alteration associated with
thiomorpholino)-1H-pyrazole-3-carboxamide) is a high affinity obesity (Matias and Di Marzo, 2007). Evidence for the efficacy of this
and selective CB1 peripheral antagonist, that showed high BBB strategy was obtained by the manipulation of 2-AG levels in two dif-
permeability resistance (Tam et al., 2012). In diet-induced obese ferent ways: the systemic administration of DAGL inhibitors, that
mice it displayed the same effects as rimonabant, by ameliorating was able to inhibit palatable or HFD intake in mice (Bisogno et al.,
the fatty liver and improving lipid profile (Tam et al., 2012). LH-21 2009, 2012); the overexpression of MAGL, that increased energy
(5-(4-Chlorophenyl)-1-(2,4-Dichlorophenyl)-3-Hexyl-1H-1,2,4- expenditure and decreased weight gain on a HFD regimen (Jung
Triazole) displays poor brain penetration and peripheral CB1 et al., 2012).
targeting in in vitro studies (Alonso et al., 2012). Diet-induced Another potential strategy is related to the allosteric modulation
obese rats showed reduced food intake, body weight gain and gene of CB1 , suggested by the finding that PSNCBAM-1, a novel allosteric
expression of leptin, FAS and Stearoyl-CoA desaturase 1 (Alonso antagonist of this receptor, inhibited appetite and produced weight
et al., 2012), following subchronic treatment with LH-21. The same loss in rats (Horswill et al., 2007). Finally, other pharmacological
drug, however, did not inhibit fatty acid accumulation in the liver, opportunities for the development of new therapies linked to the
nor it improved metabolic parameters in Zucker rats, suggesting eCB system may derive from the involvement of other receptor-
that lack of inverse agonism, or poor penetration in the brain, sand/or other eCB congeners, such as OEA.
might limit the metabolic benefits of CB1 blockade (Pavón et al., For example, Deveaux et al. (2009) showed that CB2 potentiates
2008). obesity-associated inflammation, insulin resistance and fatty liver
JD5037 is aCB1 antagonist/inverse agonist able to block in mice fed a HFD. Therefore, CB2 antagonism may open a novel
peripheral CB1 upon both acute and chronic administration in diet- therapeutic approach for the management of obesity-associated
induced obese mice, where it reduced body weight and food intake, metabolic disorders. Furthermore, the pharmacological blockade
reversed hepatic steatosis and improved insulin sensitivity (Tam of TRPV1 receptor was reported to enhance insulin release from
et al., 2012). islet ␤ cells, and to improve glucose tolerance in diabetic Zucker
URB447 has been identified as a mixed CB1 antagonist/CB2 ago- rats (Suri and Szallasi, 2008). In mice, a similar effect was linked
nist with anorectic actions and devoid of CNS effects (LoVerme et al., also to the activation of GPR119 receptors, that caused an increase
2009). in insulin secretion as well as a decrease in food intake and body
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 217

weight gain (Overton et al., 2008). In this context, it should be Bermúdez-Silva, F.J., Suárez, J., Rodríguez de Fonseca, F., Pavón, F.J., 2012. Anti-
recalled that GPR119 has been recently proposed as a preferen- obesity efficacy of LH-21, a cannabinoid CB(1) receptor antagonist with poor
brain penetration, in diet-induced obese rats. Br. J. Pharmacol. 165, 2274–2291.
tial fat sensor (Hansen et al., 2012). OEA has also emerged as a Alsio, J., Olszewski, P.K., Norback, A.H., Gunnarsson, Z.E., Levine, A.S., Pickering,
possible pharmacological alternative to CB1 antagonists for the C., Schioth, H.B., 2010. Dopamine D1 receptor gene expression decreases in
treatment of obesity, and for limiting excessive food intake in the nucleus accumbens upon long-term exposure to palatable food and dif-
fers depending on diet-induced obesity phenotype in rats. Neuroscience 171,
rodents (Romano et al., 2014). OEA is a gut-derived satiety signal 779–787.
(for a review, see Dipasquale et al., 2010; Piomelli, 2013), released Alvheim, A.R., Malde, M.K., Osei-Hyiaman, D., Lin, Y.H., Pawlosky, R.J., Madsen, L.,
from enterocytes and acting on local PPAR-␣ receptor to modulate Kristiansen, K., Frøyland, L., Hibbeln, J.R., 2012. Dietary linoleic acid elevates
endogenous 2-AG and anandamide and induces obesity. Obesity 20, 1984–1994.
absorption and oxidation of dietary fats. OEA also induces satiety by Alvheim, A.R., Torstensen, B.E., Lin, Y.H., Lillefosse, H.H., Lock, E.J., Madsen, L.,
prolonging the inter-meal interval and activating brain oxytociner- Hibbeln, J.R., Malde, M.K., 2013. Dietary linoleic acid elevates endogenous 2-
gic and histaminergic neurons (Gaetani et al., 2010; Romano et al., arachidonoylglycerol and anandamide in Atlantic salmon (Salmo salar L.) and
mice, and induces weight gain and inflammation in mice. Br. J. Nutr. 109,
2013a,b; Provensi et al., 2014). Decreased intestinal content of OEA
1508–1517.
was observed in rodent models of obesity (Artmann et al., 2008; Ameri, A., Wilhelm, A., Simmet, T., 1999. Effects of the endogeneous cannabinoid,
Aviello et al., 2008; Diep et al., 2011), where exogenous adminis- anandamide, on neuronal activity in rat hippocampal slices. Br. J. Pharmacol.
tration of OEA decreased hyperphagia and body weight gain, by 126, 1831–1839.
American Psychiatric Association, 2013. Diagnostic and Statistical Manual of Mental
increasing lipolysis, and decreasing hypertriglyceridemia, hyper- Disorders, 5th ed. American Psychiatric Association, Arlington.
cholesterolemia and liver steatosis (Rodriguez de Fonseca et al., Andries, A., Frystyk, J., Flyvbjerg, A., Støving, R.K., 2014. Dronabinol in severe, endur-
2001; Lo Verme et al., 2005; Fu et al., 2005; Newberry et al., 2012; ing anorexia nervosa: a randomized controlled trial. Int. J. Eat. Disord. 47, 18–23.
Annuzzi, G., Piscitelli, F., Di Marino, L., Patti, L., Giacco, R., Costabile, G., Bozzetto, L.,
Thabuis et al., 2010; Serrano et al., 2008). It also re-established a Riccardi, G., Verde, R., Petrosino, S., Rivellese, A.A., Di Marzo, V., 2010. Differential
normal response of the brain reward system to an intraduodenal alterations of the concentrations of endocannabinoids and related lipids in the
infusion of lipids, which seemed to be altered in diet-induced obese subcutaneous adipose tissue of obese diabetic patients. Lipids Health Dis. 9, 43.
Arbaizar, B., Gómez-Acebo, I., Llorca, J., 2008. Efficacy of topiramate in bulimia ner-
animals (Tellez et al., 2013). vosa and binge-eating disorder: a systematic review. Gen. Hosp. Psychiatry 30,
471–475.
Artmann, A., Petersen, G., Hellgren, L.I., Boberg, J., Skonberg, C., Nellemann, C.,
7. Conclusions Hansen, S.H., Hansen, H.S., 2008. Influence of dietary fatty acids on endocannabi-
noid and N-acylethanolamine levels in rat brain, liver and small intestine.
Accumulated evidence places eCB system components at the Biochim. Biophys. Acta 1781, 200–212.
Aston-Jones, G., Smith, R.J., Sartor, G.C., Moorman, D.E., Massi, L., Tahsili-Fahadan, P.,
heart of communications between the brain and peripheral organs Richardson, K.A., 2010. Lateral hypothalamic orexin/hypocretin neurons: a role
and tissues. Based on its widespread role in the multifaceted aspects in reward-seeking and addiction. Brain Res. 16 (1314), 74–90.
of food intake and energy balance, the eCB system appears as a Avena, N.M., Bocarsly, M.E., Hoebel, B.G., Gold, M.S., 2011. Overlaps in the nosol-
ogy of substance abuse and overeating: the translational implications of “food
unique modulatory system that integrates different neurotrans-
addiction”. Curr. Drug Abuse Rev. 4, 133–139.
mitter and hormonal signals. As such, it offers an array of novel Avena, N.M., Bocarsly, M.E., Rada, P., Kim, A., Hoebel, B.G., 2008b. After daily binge-
therapeutic opportunities to combat pathologies linked to aberrant ing on a sucrose solution, food deprivation induces anxiety and accumbens
dopamine/acetylcholine imbalance. Physiol. Behav. 94, 309–315.
feeding patterns and/or altered metabolic pathways, such as eating
Avena, N.M., Gearhardt, A.N., Gold, M.S., Wang, G.J., Potenza, M.N., 2012. Tossing
disorders and obesity, as well as disorders related to altered moti- the baby out with the bathwater after a brief rinse? The potential downside of
vation, hedonic value perception and/or activation of the reward dismissing food addiction based upon limited data. Nat. Rev. Neurosci. 13, 514.
system In particular, peripherally restricted CB1 antagonists might Avena, N.M., Gold, M.S., 2011. Food and addiction-sugars, fats and hedonic overeat-
ing. Addiction 5, 106–1214.
be useful in controllong overeating of fatty foods, thus possibly Avena, N.M., Murray, S., Gold, M.S., 2013. The next generation of obesity treatments:
opening the avenue to next-generation therapeutics able to con- beyond suppressing appetite. Front. Psychol. 9, 4–721.
trol excessive food intake and energy balance. The mechanisms Avena, N.M., Rada, P., Hoebel, B.G., 2008a. Evidence for sugar addiction: behav-
ioral and neurochemical effects of intermittent, excessive sugar intake. Neurosci.
engaged span from those involving the hedonic aspects of food Biobehav. Rev. 32, 20–39.
consumption, to those depending on storage of its energy content. Aviello, G., Matias, I., Capasso, R., Petrosino, S., Borrelli, F., Orlando, P., Romano, B.,
Capasso, F., Di Marzo, V., Izzo, A.A., 2008. Inhibitory effect of the anorexic com-
pound oleoylethanolamide on gastric emptying in control and overweight mice.
Acknowledgements J. Mol. Med. 86, 413–422.
Balvers, M.G., Verhoeckx, K.C., Bijlsma, S., Rubingh, C.M., Meijerink, J., Wortelboer,
H.M., Witkamp, R.F., 2012. Fish oil and inflammatory status alter the n-3 to n-6
The work was supported by the Italian Ministry of Education, balance of the endocannabinoid and oxylipin metabolomes in mouse plasma
University and Research under grants FIRB-RBFR12DELS to CDA, and tissues. Metabolomics 8, 1130–1147.
CC and SG and PRIN2010-11 to MM. Banni, S., Carta, G., Murru, E., Cordeddu, L., Giordano, E., Sirigu, A.R., Berge, K., Vik,
H., Maki, K.C., Di Marzo, V., Griinari, M., 2011. Krill oil significantly decreases
2-arachidonoylglycerol plasma levels in obese subjects. Nutr. Metab. 8, 7.
References Batetta, B., Griinari, M., Carta, G., Murru, E., Ligresti, A., Cordeddu, L., Giordano,
E., Sanna, F., Bisogno, T., Uda, S., Collu, M., Bruheim, I., Di Marzo, V., Banni, S.,
Addy, C., Li, S., Agrawal, N., Stone, J., Majumdar, A., Zhong, L., Li, H., Yuan, J., Maes, 2009. Endocannabinoids may mediate the ability of (n-3) fatty acids to reduce
A., Rothenberg, P., Cote, J., Rosko, K., Cummings, C., Warrington, S., Boyce, ectopic fat and inflammatory mediators in obese Zucker rats. J. Nutr. 139 (8),
M., Gottesdiener, K., Stoch, A., Wagner, J., 2008a. Safety, tolerability, pharma- 1495–1501.
cokinetics, and pharmacodynamic properties of taranabant, a novel selective Batterham, R.L., Cowley, M.A., Small, C.J., Herzog, H., Cohen, M.A., Dakin, C.L., Wren,
cannabinoid-1 receptor inverse agonist, for the treatment of obesity: results A.M., Brynes, A.E., Low, M.J., Ghatei, M.A., Cone, R.D., Bloom, S.R., 2002. Gut
from a double-blind, placebo-controlled, single oral dose study in healthy vol- hormone PYY (3-36) physiologically inhibits food intake. Nature 418, 650–654.
unteers. J. Clin. Pharmacol. 48, 418–427. Bello, N.T., Coughlin, J.W., Redgrave, G.W., Ladenheim, E.E., Moran, T.H., Guarda, A.S.,
Addy, C., Rothenberg, P., Li, S., Majumdar, A., Agrawal, N., Li, H., Zhong, L., Yuan, J., 2012. Dietary conditions and highly palatable food access alter rat cannabinoid
Maes, A., Dunbar, S., Cote, J., Rosko, K., Van Dyck, K., De Lepeleire, I., de Hoon, receptor expression and binding density. Physiol. Behav. 105, 720–726.
J., Van Hecken, A., Depré, M., Knops, A., Gottesdiener, K., Stoch, A., Wagner, Bellocchio, L., Lafenêtre, P., Cannich, A., Cota, D., Puente, N., Grandes, P., Chaouloff,
J., 2008b. Multiple-dose pharmacokinetics, pharmacodynamics, and safety of F., Piazza, P.V., Marsicano, G., 2010. Bimodal control of stimulated food intake
taranabant, a novel selective cannabinoid-1 receptor inverse agonist, in healthy by the endocannabinoid system. Nat. Neurosci. 13, 281–283.
male volunteers. J. Clin. Pharmacol. 48, 734–744. Ben-Shabat, S., Fride, E., Sheskin, T., Tamiri, T., Rhee, M.H., Vogel, Z., Bisogno, T., De
Ahmed, S.H., Kenny, P.J., Koob, G.F., Markou, A., 2002. Neurobiological evidence Petrocellis, L., Di Marzo, V., Mechoulam, R., 1998. An entourage effect: inactive
for hedonic allostasis associated with escalating cocaine use. Nat. Neurosci. 5, endogenous fatty acid glycerol esters enhance 2-arachidonoyl-glycerol cannabi-
625–626. noid activity. Eur. J. Pharmacol. 353, 23–31.
Alger, B.E., Kim, J., 2011. Supply and demand for endocannabinoids. Trends Neurosci. Bensaid, M., Gary-Bobo, M., Esclangon, A., Maffrand, J.P., Le Fur, G., Oury-Donat, F.,
34, 304–515. Soubrié, P., 2003. The cannabinoid CB1 receptor antagonist SR141716 increases
Alonso, M., Serrano, A., Vida, M., Crespillo, A., Hernandez-Folgado, L., Jagerovic, N., Acrp30 mRNA expression in adipose tissue of obese fa/fa rats and in cultured
Goya, P., Reyes-Cabello, C., Perez-Valero, V., Decara, J., Macías-González, M., adipocyte cells. Mol. Pharmacol. 63, 908–914.
218 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

Berger, A., Crozier-Willi, G., Di Marzo, V., 2003. Anandamides and diet. A new pot Brand, T., Spanagel, R., Schneider, M., 2012. Decreased reward sensitivity in rats
of nutritional research is simmering. In: Neeser, J.R., German, B.J. (Eds.), Bio- from the Fischer344 strain compared to Wistar rats is paralleled by differences
processes and Biotechnology for Functional Foods and Nutraceuticals. Marcel in endocannabinoid signaling. PLoS ONE 7, e31169.
Dekker, New York, pp. 225–261. Broft, I., Spanos, A., Corwin, R.L., Mayer, L., Steinglass, J., Devlin, M.J., Attia, E., Walsh,
Berger, A., Crozier, G., Bisogno, T., Cavaliere, P., Innis, S., Di Marzo, V., 2001. Anan- B.T., 2007. Baclofen for binge eating: an open-label trial. Int. J. Eat. Disord. 40,
damide and diet: inclusion of dietary arachidonate a.nd docosahexaenoate leads 687–691.
to increased brain levels of the corresponding N-acylethanolamines in piglets. Brown, A.J., Robin Hiley, C., 2009. Is GPR55 an anandamide receptor? Vitam. Horm.
Proc. Natl. Acad. Sci. U.S.A. 98, 6402–6406. 81, 111–113.
Berger, A., Roberts, M.A., 2005. Dietary effects of arachidonate-rich fungal oil and fish Buda-Levin, A., Wojnicki, F.H., Corwin, R.L., 2005. Baclofen reduces fat intake under
oil on murine hippocampal gene expression. In: Berger, A., Roberts, M.A. (Eds.), binge-type conditions. Physiol. Behav. 86, 176–184.
Unraveling Lipid Metabolism with Microarrays. Marcel Dekker, New York, pp. Buettner, C., Muse, E.D., Cheng, A., Chen, L., Scherer, T., Pocai, A., Su, K., Cheng, B., Li, X.,
69–100. Harvey-White, J., Schwartz, G.J., Kunos, G., Rossetti, L., Buettner, C., 2008. Leptin
Bergh, C., Brodin, U., Lindberg, G., Södersten, P., 2002. Randomized controlled trial controls adipose tissue lipogenesis via central, STAT3-independent mechanisms.
of a treatment for anorexia and bulimia nervosa. Proc. Natl. Acad. Sci. U.S.A. 99, Nat. Med. 14, 667–675.
9486–9491. Burdyga, G., Lal, S., Varro, A., Dimaline, R., Thompson, D.G., Dockray, G.J., 2004.
Bermudez-Silva, F.J., Cardinal, P., Cota, D., 2012. The role of the endocannabinoid Expression of cannabinoid CB1 receptors by vagal afferent neurons is inhibited
system in the neuroendocrine regulation of energy balance. J. Psychopharmacol. by cholecystokinin. J. Neurosci. 24, 2708–2715.
26, 114–124. Burdyga, G., Varro, A., Dimaline, R., Thompson, D.G., Dockray, G.J., 2006. Ghrelin
Bermúdez-Silva, F.J., Suárez, J., Baixeras, E., Cobo, N., Bautista, D., Cuesta-Muñoz, receptors in rat and human nodose ganglia: putative role in regulating CB-1
A.L., Fuentes, E., Juan-Pico, P., Castro, M.J., Milman, G., Mechoulam, R., Nadal, A., and MCH receptor abundance. Am. J. Physiol. Gastrointest. Liver Physiol. 290,
Rodríguez de Fonseca, F., 2008. Presence of functional cannabinoid receptors in G1289–G1297.
human endocrine pancreas. Diabetologia 51, 476–487. Burdyga, G., Varro, A., Dimaline, R., Thompson, D.G., Dockray, G.J., 2010. Expression
Bermudez-Silva, F.J., Viveros, M.P., McPartland, J.M., Rodriguez de Fonseca, F., 2010. of cannabinoid CB1 receptors by vagal afferent neurons: kinetics and role in
The endocannabinoid system, eating behavior and energy homeostasis: the end influencing neurochemical phenotype. Am. J. Physiol. Gastrointest. Liver Physiol.
or a new beginning? Pharmacol. Biochem. Behav. 95, 375–382. 299, G63–G69.
Berry, E.M., 2006. Pilot Study of THC (2.5 mg 2) in 9 ambulatory AN patients Cani, P.D., Montoya, M.L., Neyrinck, A.M., Delzenne, N.M., Lambert, D.M., 2004. Poten-
(abstract). In: The sixth nordic congress on eating disorders, Aarhus, Denmark. tial modulation of plasma ghrelin and glucagon-like peptide-1 by anorexigenic
Bhatti, A.S., Aydin, C., Oztan, O., Ma, Z., Hall, P., Tao, R., Isgor, C., 2009. Effects of cannabinoid compounds, SR141716A (rimonabant) and oleoylethanolamide. Br.
a cannabinoid receptor (CB) 1 antagonist AM251 on behavioral sensitization to J. Nutr. 92, 757–761.
nicotine in a rat model of novelty-seeking behavior: correlation with hippocam- Chaperon, F., Soubrié, P., Puech, A.J., Thiébot, M.H., 1998. Involvement of central
pal 5HT. Psychopharmacology (Berl.) 203, 23–32. cannabinoid (CB1) receptors in the establishment of place conditioning in rats.
Bisogno, T., Maccarrone, M., 2014. Endocannabinoid signaling and its regulation by Psychopharmacology (Berl.) 135, 324–332.
nutrients. Biofactors 40, 373–380. Chicca, A., Marazzi, J., Nicolussi, S., Gertsch, J., 2012. Evidence for bidirectional endo-
Bisogno, T., Burston, J.J., Rai, R., Allarà, M., Saha, B., Mahadevan, A., Razdan, R.K., cannabinoid transport across cell membranes. J. Biol. Chem. 287, 34660–34682.
Wiley, J.L., Di Marzo, V., 2009. Synthesis and pharmacological activity of a potent Christensen, R., Kristensen, P.K., Bartels, E.M., Bliddal, H., Astrup, A., 2007. Efficacy
inhibitor of the biosynthesis of the endocannabinoid 2-arachidonoylglycerol. and safety of the weight-loss drug rimonabant: a meta-analysis of randomised
ChemMedChem 4, 946–950. trials. Lancet 370, 1706–1713.
Bisogno, T., Howell, F., Williams, G., Minassi, A., Cascio, M.G., Ligresti, A., Matias, I., Cifani, C., Micioni Di Bonaventura, M.V., Vitale, G., Ruggieri, V., Ciccocioppo, R., Massi,
Schiano-Moriello, A., Paul, P., Williams, E.J., Gangadharan, U., Hobbs, C., Di Marzo, M., 2010. Effect of salidroside, active principle of Rhodiola rosea extract, on binge
V., Doherty, P., 2003. Cloning of the first sn1-DAG lipases points to the spatial eating. Physiol. Behav. 101, 555–562.
and temporal regulation of endocannabinoid signaling in the brain. J. Cell Biol. Cifani, C., Polidori, C., Melotto, S., Ciccocioppo, R., Massi, M., 2009. A preclinical model
163, 463–468. of binge eating elicited by yo-yo dieting and stressful exposure to food: effect
Bisogno, T., Ligresti, A., Di Marzo, V., 2005. The endocannabinoid signalling system: of sibutramine, fluoxetine, topiramate, and midazolam. Psychopharmacology
biochemical aspects. Pharmacol. Biochem. Behav. 81, 224–238. (Berl.) 204, 113–125.
Bisogno, T., Mahadevan, A., Coccurello, R., Chang, J.W., Allarà, M., Chen, Y., Gia- Cippitelli, A., Bilbao, A., Gorriti, M.A., Navarro, M., Massi, M., Piomelli, D., Ciccocioppo,
covazzo, G., Lichtman, A., Cravatt, B., Moles, A., Di Marzo, V., 2012. A novel R., Rodríguez de Fonseca, F., 2007. The anandamide transport inhibitor AM404
fluorophosphonate inhibitor of the biosynthesis of the endocannabinoid 2- reduces ethanol self-administration. Eur. J. Neurosci. 26, 476–486.
arachidonoylglycerol with potential anti-obesity effects. Br. J. Pharmacol. 169, Cluny, N.L., Keenan, C.M., Lutz, B., Piomelli, D., Sharkey, K.A., 2009. The identifica-
784–793. tion of peroxisome proliferator-activated receptor alpha-independent effects of
Bisogno, T., Melck, D., Bobrov, M.Y., Gretskaya, N.M., Bezuglov, V.V., De oleoylethanolamide on intestinal transit in mice. Neurogastroenterol. Motil. 21,
Petrocellis, L., Di Marzo, V., 2000. N-acyl-dopamines: novel synthetic 420–429.
CB1 cannabinoid-receptor ligands and inhibitors of anandamide inactiva- Cohen, C., Perrault, G., Voltz, C., Steinberg, R., Soubrié, P., 2002. SR141716, a
tion with cannabimimetic activity in vitro and in vivo. Biochem. J. 351, central cannabinoid (CB(1)) receptor antagonist, blocks the motivational and
817–824. dopamine-releasing effects of nicotine in rats. Behav. Pharmacol. 13, 451–463.
Blasio, A., Iemolo, A., Sabino, V., Petrosino, S., Steardo, L., Rice, K.C., Orlando, P., Ian- Colantuoni, C., Rada, P., McCarthy, J., Patten, C., Avena, N.M., Chadeayne, A., Hoebel,
notti, F.A., Di Marzo, V., Zorrilla, E.P., Cottone, P., 2013. Rimonabant precipitates B.G., 2002. Evidence that intermittent, excessive sugar intake causes endogenous
anxiety in rats withdrawn from palatable food: role of the central amygdala. opioid dependence. Obes. Res. 10, 478–488.
Neuropsychopharmacology 38, 2498–2507. Corwin, R.L., 2011. The face of uncertainty eats. Curr. Drug Abuse Rev. 4, 174–181.
Bluher, M., Engeli, S., Klöting, N., Berndt, J., Fasshauer, M., Bátkai, S., Pacher, P., Schön, Corwin, R.L., Boan, J., Peters, K.F., Ulbrecht, J.S., 2012. Baclofen reduces binge eat-
M.R., Jordan, J., Stumvoll, M., 2006. Dysregulation of the peripheral and adi- ing in a double-blind,placebo-controlled,crossover study. Behav. Pharmacol. 23,
pose tissue endocannabinoid system in human abdominal obesity. Diabetes 55, 616–625.
3053–3060. Cota, D., 2007. CB1 receptors: emerging evidence for central and peripheral mech-
Blundell, J.E., Finlayson, G., 2011. Food addiction not helpful: the hedonic component anisms that regulate energy balance, metabolism, and cardiovascular health.
implicit wanting is important. Addiction 106, 1216–1218. Diabetes Metab. Res. Rev. 23, 507–517.
Blüher, M., Engeli, S., Klöting, N., Berndt, J., Fasshauer, M., Bátkai, S., Pacher, P., Schön, Cota, D., Marsicano, G., Tschöp, M., Grübler, Y., Flachskamm, C., Schubert, M., Auer,
M.R., Jordan, J., Stumvoll, M., 2006. Dysregulation of the peripheral and adi- D., Yassouridis, A., Thöne-Reineke, C., Ortmann, S., Tomassoni, F., Cervino, C.,
pose tissue endocannabinoid system in human abdominal obesity. Diabetes 55, Nisoli, E., Linthorst, A.C., Pasquali, R., Lutz, B., Stalla, G.K., Pagotto, U., 2003. The
3053–3060. endogenous cannabinoid system affects energy balance via central orexigenic
Bocarsly, M.E., Berner, L.A., Hoebel, B.G., Avena, N.M., 2011. Rats that binge eat drive and peripheral lipogenesis. J. Clin. Invest. 112, 423–431.
fat-rich food do not show somatic signs or anxiety associated with opiate-like Cravatt, B.F., Giang, D.K., Mayfield, S.P., Boger, D.L., Lerner, R.A., Gilula, N.B., 1996.
withdrawal: implications for nutrientspecific food addiction behaviors. Physiol. Molecular characterization of an enzyme that degrades neuromodulatory fatty-
Behav. 104, 865–872. acid amides. Nature 384, 83–87.
Boggiano, M.M., Chandler, P.C., Viana, J.B., Oswald, K.D., Maldonado, C.R., Wauford, Côté, M., Matias, I., Lemieux, I., Petrosino, S., Alméras, N., Després, J.P., Di Marzo,
P.K., 2005. Combined dieting and stress evoke exaggerated responses to opioids V., 2007. Circulating endocannabinoid levels, abdominal adiposity and related
in binge-eating rats. Behav. Neurosci. 119, 1207–1214. cardiometabolic risk factors in obese men. Int. J. Obes. 31, 692–699.
Bose, M., Lambert, J.D., Ju, J., Reuhl, K.R., Shapses, S.A., Yang, C.S., 2008. The Davis, C., Curtis, C., Levitan, R.D., Carter, J.C., Kaplan, A.S., Kennedy, J.L., 2011. Evidence
major green tea polyphenol, (−)-epigallocatechin-3-gallate, inhibits obesity, that “food addiction” is a valid phenotype of obesity. Appetite 57, 711–717.
metabolic syndrome, and fatty liver disease in high-fat-fed mice. J. Nutr. 138, De Luca, M.A., Solinas, M., Bimpisidis, Z., Goldberg, S.R., Di Chiara, G., 2012.
1677–1683. Cannabinoid facilitation of behavioral and biochemical hedonic taste responses.
Bossong, M., Van Berckel, B., Boellaard, R., Zuurman, L., Schuit, R., Windhorst, Neuropharmacology 63, 161–168.
A., Van Gerven, J., Ramsey, N., Lammertsma, A., Kahn, R., 2009. Delta 9- de Luis, D.A., Aller, R., Izaola, O., Conde, R., Sagrado, M.G., Primo, D., Castro, M.J., 2012.
tetrahydrocannabinol induces dopamine release in the human striatum. Relationship among metabolic syndrome, C358A polymorphism of the endo-
Neuropsychopharmacology 34, 759–766. cannabinoid degrading enzyme fatty acid amide hydrolase (FAAH) and insulin
Braida, D., Iosuè, S., Pegorini, S., Sala, M., 2004. D9-Tetrahydrocannabinol-induced resistance. J. Diabetes Complicat. 26, 328–332.
conditioned place preference and intracerebroventricular self-administration in de Luis, D.A., Gonzalez Sagrado, M., Aller, R., Izaola, O., Conde, R., Romero, E., 2010a.
rats. Eur. J. Pharmacol. 506, 63–69. C358A missense polymorphism of the endocannabinoid-degrading enzyme
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 219

fatty acid amide hydrolase (FAAH) and visfatin levels in obese females. Int. J. evidence of a widespread role for fatty acid amide hydrolase in regulation of
Obes. 34, 511–515. endocannabinoid signaling. Neuroscience 119, 481–496.
de Luis, D.A., Sagrado, M.G., Aller, R., Izaola, O., Conde, R., Romero, E., 2010b. C358A Ellis, J., Pediani, J.D., Canals, M., Milasta, S., Milligan, G., 2006. Orexin-1 receptor-
missense polymorphism of the endocannabinoid degrading enzyme fatty acid cannabinoid CB1 receptor heterodimerization results in both ligand-dependent
amide hydrolase (FAAH) and insulin resistance in patients with diabetes mellitus and -independent coordinated alterations of receptor localization and function.
type 2. Diabetes Res. Clin. Pract. 88, 76–80. J. Biol. Chem. 281, 38812–38824.
De Petrocellis, L., Cascio, M.G., Di Marzo, V., 2004. The endocannabinoid system: a Engeli, S., Böhnke, J., Feldpausch, M., Gorzelniak, K., Janke, J., Bátkai, S., Pacher,
general view and latest additions. Br. J. Pharmacol. 141, 765–774. P., Harvey-White, J., Luft, F.C., Sharma, A.M., Jordan, J., 2005. Activation
De Vries, T.J., Homberg, J.R., Binnekade, R., Raasø, H., Schoffelmeer, A.N., 2003. of the peripheral endocannabinoid system in human obesity. Diabetes 54,
Cannabinoid modulation of the reinforcing and motivational properties of 2838–2843.
heroin and heroin-associated cues in rats. Psychopharmacology (Berl.) 168, Engeli, S., Lehmann, A.C., Kaminski, J., Haas, V., Janke, J., Zoerner, A.A., Luft, F.C., Tsikas,
164–169. D., Jordan, J., 2014. Influence of dietary fat intake on the endocannabinoid system
De Weijer, B.A., Van de Giessen, E., Van Amelsvoort, T.A., Boot, E., Braak, B., Janssen, in lean and obese subjects. Obesity (Silver Spring) 22, E70–E76.
I.M., van de Laar, A., Fliers, E., Serlie, M.J., Booij, J., 2011. Lower striatal dopamine Fadda, P., Scherma, M., Spano, M.S., Salis, P., Melis, V., Fattore, L., et al., 2006. Cannabi-
D2/3 receptor availability in obese compared with non-obese subjects. EJNMMI noid self-administration increases dopamine release in the nucleus accumbens.
Res. 1, 37. Neuroreport 17, 1629–1632.
Després, J.P., Golay, A., Sjostrom, L., 2005. Effects of rimonabant on metabolic Fairburn, C.G., Cooper, Z., Doll, H.A., Norman, P., O’Connor, M., 2000. The natural
risk factors in overweight patients with dyslipidemia. N. Engl. J. Med. 353, course of bulimia nervosa and binge eating disorder in young women. Arch.
2121–2134. Gen. Psychiatry 57, 659–665.
Després, J.P., Ross, R., Boka, G., Alméras, N., Lemieux, I., 2009. Effect of rimonabant Fegley, D., Kathuria, S., Mercier, R., Li, C., Goutopoulos, A., Makriyannis, A., Piomelli,
on the hightriglyceride/low-HDL-cholesterol dyslipidemia, intraabdominal adi- D., 2004. Anandamide transport is independent of fatty-acid amide hydrolase
posity, and liver fat: the ADAGIO-lipids trial. Arterioscler. Thromb. Vasc. Biol. 29, activity and is blocked by the hydrolysis-resistant inhibitor AM1172. Proc. Natl.
416–423. Acad. Sci. U.S.A. 101, 8756–8761.
Devane, W.A., Dysarz, F.A., Johnson, M.R., Melvin, L.S., Howlett, A.C., 1988. Deter- Filbey, F., Schacht, J., Myers, U., Chavez, R., Hutchison, K., 2009. Marijuana craving in
mination and characterization of a cannabinoid receptor in rat brain. Mol. the brain. Proc. Natl. Acad. Sci. U.S.A. 106, 13016–13021.
Pharmacol. 34, 605–613. Flores, Á., Maldonado, R., Berrendero, F., 2013. The hypocretin/orexin receptor
Devane, W.A., Hanus, L., Breuer, A., Pertwee, R.G., Stevenson, L.A., Griffin, G., Gibson, 1 as a novel target to modulate cannabinoid reward. Biol. Psychiatry 75,
D., Mandelbaum, A., Etinger, A., Mechoulam, R., 1992. Isolation and structure 499–507.
of a brain constituent that binds to the cannabinoid receptor. Science 258, Fredriksson, R., Höglund, P.J., Gloriam, D.E., Lagerström, M.C., Schiöth, H.B., 2003.
1946–1949. Seven evolutionarily conserved human rhodopsin G protein-coupled receptors
Deveaux, V., Cadoudal, T., Ichigotani, Y., Teixeira-Clerc, F., Louvet, A., Manin, S., Nhieu, lacking close relatives. FEBS Lett. 554, 381–388.
J.T., Belot, M.P., Zimmer, A., Even, P., Cani, P.D., Knauf, C., Burcelin, R., Bertola, A., Freeman, J.M., Kossoff, E.H., Hartman, A.L., 2007. The ketogenic diet: one decade
Le Marchand-Brustel, Y., Gual, P., Mallat, A., Lotersztajn, S., 2009. Cannabinoid later. Pediatrics 119, 535–543.
CB2 receptor potentiates obesity-associated inflammation, insulin resistance French, E.D., Dillon, K., Wu, X., 1997. Cannabinoids excite dopamine neurons in the
and hepatic steatosis. PLoS ONE 4, e5844. ventral tegmentum and substantia nigra. Neuroreport 8, 649–652.
Di Marzo, V., De Petrocellis, L., 2010. Endocannabinoids as regulators of tran- Fride, E., 2002. Endocannabinoids in the central nervous system – an overview.
sient receptor potential (TRP) channels: a further opportunity to develop new Prostaglandins Leukot. Essent. Fatty Acids 66, 221–233.
endocannabinoid-based therapeutic drugs. Curr. Med. Chem. 17, 1430–1449. Frieling, H., Albrecht, H., Jedtberg, S., Gozner, A., Lenz, B., Wilhelm, J., Hillemacher,
Di Marzo, V., Desprès, J.P., 2009. CB1 antagonists for obesity – what lessons have we T., de Zwaan, M., Kornhuber, J., Bleich, S., 2009. Elevated cannabinoid 1 receptor
learned from rimonabant? Nat. Rev. Endocrinol. 5, 633–638. mRNA is linked to eating disorder related behavior and attitudes in females with
Di Marzo, V., Fontana, A., Cadas, H., Schinelli, S., Cimino, G., Schwartz, J.C., Piomelli, eating disorders. Psychoneuroendocrinology 34, 620–624.
D., 1994. Formation and inactivation of endogenous cannabinoid anandamide Fu, J., Astarita, G., Gaetani, S., Kim, J., Cravatt, B.F., Mackie, K., Piomelli, D., 2007. Food
in central neurons. Nature 372, 686–691. intake regulates oleoylethanolamide formation and degradation in the proximal
Di Marzo, V., Goparaju, S.K., Wang, L., Liu, J., Bátkai, S., Járai, Z., Fezza, F., Miura, G.I., small intestine. J. Biol. Chem. 282, 1518–1528.
Palmiter, R.D., Sugiura, T., Kunos, G., 2001. Leptin-regulated endocannabinoids Fu, J., Bottegoni, G., Sasso, O., Bertorelli, R., Rocchia, W., Masetti, M., Guijarro, A.,
are involved in maintaining food intake. Nature 410, 822–825. Lodola, A., Armirotti, A., Garau, G., Bandiera, T., Reggiani, A., Mor, M., Cavalli,
Di Marzo, V., Griinari, M., Carta, G., Murru, E., Ligrestia, A., Cordeddu, L., Giordano, A., Piomelli, D., 2011. A catalytically silent FAAH-1 variant drives anandamide
E., Bisogno, T., Collu, M., Batetta, B., Uda, S., Berge, K., Bannic, S., 2010. Dietary transport in neurons. Nat. Neurosci. 15, 64–69.
krill oil increases docosahexaenoic acid and reduces 2-arachidonoylglycerol but Fu, J., Kim, J., Oveisi, F., Astarita, G., Piomelli, D., 2008. Targeted enhancement of
not N-acylethanolamine levels in the brain of obese Zucker rats. Int. Dairy J. 20, oleoylethanolamide production in proximal small intestine induces across-meal
231–235. satiety in rats. Am. J. Physiol. Regul. Integr. Comp. Physiol. 295, R45–R50.
Di Marzo, V., Ligresti, A., Cristino, L., 2009. The endocannabinoid system as a link Fu, J., Oveisi, F., Gaetani, S., Lin, E., Piomelli, D., 2005. Oleoylethanolamide, an endoge-
between homoeostatic and hedonic pathways involved in energy balance reg- nous PPAR alpha agonist, lowers body weight and hyperlipidemia in obese rats.
ulation. Int. J. Obes. 33, 18–24. Neuropharmacology 48, 1147–1153.
Di Marzo, V., Matias, I., 2005. Endocannabinoid control of food intake and energy Gaetani, S., Fu, J., Cassano, T., Dipasquale, P., Romano, A., Righetti, L., Cianci, S.,
balance. Nat. Neurosci. 8, 585–589. Laconca, L., Giannini, E., Scaccianoce, S., Mairesse, J., Cuomo, V., Piomelli, D.,
Di, S., Malcher-Lopes, R., Halmos, K.C., Tasker, J.G., 2003. Nongenomic glucocorticoid 2010. The fat-induced satiety factor OEA suppresses feeding through central
inhibition via endocannabinoid release in the hypothalamus: a fast feedback release of oxytocin. J. Neurosci. 30, 8096–8101.
mechanism. J. Neurosci. 23, 4850–4857. Gaetani, S., Kaye, W.H., Cuomo, V., Piomelli, D., 2008. Role of endocannabinoids and
Diep, T.A., Madsen, A.N., Holst, B., Kristiansen, M.M., Wellner, N., Hansen, S.H., their analogues in obesity and eating disorders. Eat. Weight Disord. 13, e42–e48.
Hansen, H.S., 2011. Dietary fat decreases intestinal levels of the anorectic lipids Gallate, J.E., Mc Gregor, I.S., 1999. The motivation for beer in rats: effects of ritanserin,
through a fat sensor. FASEB J. 25, 765–774. naloxone and SR141716. Psychopharmacology (Berl.) 142, 302–308.
Diergaarde, L., de Vries, W., Raasø, H., Schoffelmeer, A.N., De Vries, T.J., 2008. Gamaleddin, I., Guranda, M., Scherma, M., Fratta, W., Makriyannis, A., Vadivel, S.K.,
Contextual renewal of nicotine seeking in rats and its suppression by Goldberg, S.R., Le Foll, B., 2013. AM404 attenuates reinstatement of nicotine
the cannabinoid-1 receptor antagonist Rimonabant (SR141716A). Neuro- seeking induced by nicotine-associated cues and nicotine priming but does not
pharmacology 55, 712–716. affect nicotine- and food-taking. J. Psychopharmacol. 27, 564–571.
Dinh, T.P., Carpenter, D., Leslie, F.M., Freund, T.F., Katona, I., Sensi, S.L., Kathuria, S., Gamber, K.M., Macarthur, H., Westfall, T.C., 2005. Cannabinoids augment the
Piomelli, D., 2002. Brain monoglyceride lipase participating in endocannabinoid release of neuropeptide Y in the rat hypothalamus. Neuropharmacology 49,
inactivation. Proc. Natl. Acad. Sci. U.S.A. 99, 10819–10824. 646–652.
Dipasquale, P., Romano, A., Cianci, S., Righetti, L., Gaetani, S., 2010. Gardner, E.L., 1997. Brain reward mechanisms. In: Lowinson, J.H., Ruiz, P., Millman,
Oleoylethanolamide: a new player in energy metabolism control. Role in R.B., Langrod, J.G. (Eds.), Substance Abuse: A Comprehensive Textbook. , 3rd ed.
food intake Drug Discovery Today. Dis. Mech. 7, e169–e174. Williams & Wilkins, Baltimore, pp. 51–85.
DiPatrizio, N.V., 2014. Is fat taste ready for primetime? Physiol Behav., pii: S0031- Gardner, E.L., 2000. What we have learned about addiction from animal models of
9384(14)00136-X. drug self-administration. Am. J. Addict. 9, 285–313.
DiPatrizio, N.V., Astarita, G., Schwartz, G., Li, X., Piomelli, D., 2011. Endocannabinoid Gardner, E.L., 2005. Endocannabinoid signaling system and brain reward: emphasis
signal in the gut controls dietary fat intake. Proc. Natl. Acad. Sci. U.S.A. 108, on dopamine. Pharmacol. Biochem. Behav. 81, 263–284.
12904–12908. Gasperi, V., Fezza, F., Pasquariello, N., Bari, M., Oddi, S., Finazzi-Agrò, A., Maccarrone,
DiPatrizio, N.V., Joslin, A., Jung, K.M., Piomelli, D., 2013. Endocannabinoid signaling M., 2007. Endocannabinoids in adipocytes during differentiation and their role
in the gut mediates preference for dietary unsaturated fats. FASEB J. 27, in glucose uptake. Cell. Mol. Life Sci. 64, 219–229.
2513–2520. Gearhardt, A.N., Corbin, W.R., Brownell, K.D., 2009. Preliminary validation of the Yale
Economidou, D., Mattioli, L., Cifani, C., Perfumi, M., Massi, M., Cuomo, V., Trabace, Food Addiction Scale. Appetite 52, 430–436.
L., Ciccocioppo, R., 2006. Effect of the cannabinoid CB1 receptor antagonist SR- Gearhardt, A.N., Davis, C., Kuschner, R., Brownell, K.D., 2011a. The addiction potential
141716A on ethanol self-administration and ethanol-seeking behaviour in rats. of hyperpalatable foods. Curr. Drug Abuse 4, 140–145.
Psychopharmacology (Berl.) 183, 394–403. Gearhardt, A.N., White, M.A., Masheb, R.M., Grilo, C.M., 2013. An examination of
Egertová, M., Cravatt, B.F., Elphick, M.R., 2003. Comparative analysis of fatty acid food addiction in a racially diverse sample of obese patients with binge eating
amide hydrolase and cb(1) cannabinoid receptor expression in the mouse brain: disorder in primary care settings. Compr. Psychiatry 54, 500–505.
220 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

Gearhardt, A.N., White, M.A., Masheb, R.M., Morgan, P.T., Crosby, R.D., Grilo, C.M., Increment of hypothalamic 2-arachidonoylglycerol induces the preference for
2011b. An examination of the food addiction construct in obese patients with a high-fat diet via activation of cannabinoid 1 receptors. Behav. Brain Res. 216,
binge eating disorder. Int. J. Eat. Disord. 45, 657–663. 477–480.
Geiger, B.M., Haburcak, M., Avena, N.M., Moyer, M.C., Hoebel, B.G., Pothos, E.N., Hoebel, B.G., Teitelbaum, P., 1962. Hypothalamic control of feeding and self-
2009. Deficits of mesolimbic dopamine neurotransmission in rat dietary obesity. stimulation. Science 135, 375–377.
Neuroscience 159, 1193–1199. Holden, C., 2001. Behavioral’ addictions: do they exist? Science 294, 980–982.
Gérard, N., Pieters, G., Goffin, K., Bormans, G., Van Laere, K., 2011. Brain type 1 Holden, C., 2010. Behavioral addictions debut in proposed DSM-V. Science 327, 935.
cannabinoid receptor availability in patients with anorexia and bulimia nervosa. Horder, J., Harmer, C.J., Cowen, P.J., McCabe, C., 2010. Reduced neural response
Biol. Psychiatry 70, 777–784. to reward following 7 days treatment with the cannabinoid CB(1) antag-
Gertsch, J., Leonti, M., Raduner, S., Racz, I., Chen, J.Z., Xie, X.Q., Altmann, K.H., Karsak, onist rimonabant in healthy volunteers. Int. J. Neuropsychopharmacol. 13,
M., Zimmer, A., 2008. Beta-caryophyllene is a dietary cannabinoid. Proc. Natl. 1103–1113.
Acad. Sci. U.S.A. 105, 9099–9104. Horswill, J.G., Bali, U., Shaaban, S., Keily, J.F., Jeevaratnam, P., Babbs, A.J., Reynet, C.,
Gessa, G.L., Melis, M., Muntoni, A.L., Diana, M., 1998. Cannabinoids activate mesolim- Wong Kai In, P., 2007. PSNCBAM-1, a novel allosteric antagonist at cannabi-
bic dopamine neurons by an action on cannabinoid CB1 receptors. Eur. J. noid CB1 receptors with hypophagic effects in rats. Br. J. Pharmacol. 152,
Pharmacol. 341, 39–44. 805–814.
Getachew, B., Hauser, S.R., Dhaher, R., Katner, S.N., Bell, R.L., Oster, S.M., McBride, W.J., Huang, H., Acuna-Goycolea, C., Li, Y., Cheng, H.M., Obrietan, K., van den Pol, A.N.,
Rodd, Z.A., 2011. CB1 receptors regulate alcohol-seeking behavior and alcohol 2007. Cannabinoids excite hypothalamic melanin-concentrating hormone but
self-administration of alcohol-preferring (P) rats. Pharmacol. Biochem. Behav. inhibit hypocretin/orexin neurons: implications for cannabinoid actions on food
97, 669–675. intake and cognitive arousal. J. Neurosci. 27, 4870–4881.
Getty-Kaushik, L., Richard, A.M., Deeney, J.T., Krawczyk, S., Shirihai, O., Corkey, B.E., Hudson, J.I., Hiripi, E., Pope Jr., H.G., Kessler, R.C., 2007. The prevalence and corre-
2009. The CB1 antagonist rimonabant decreases insulin hypersecretion in rat lates of eating disorders in the National Comorbidity Survey Replication. Biol.
pancreatic islets. Obesity 17, 1856–1860. Psychiatry 61, 348–358.
Gifford, A.N., Gardner, E.L., Ashby Jr., C.R., 1997. The effect of intravenous administra- Huebner, H., 1993. Endorphins: Eating Disorders and Other Addictive Behaviors. W.
tion of delta-9-tetrahydrocannabinol on the activity of A10 dopamine neurons W. Norton and Company, New York.
recorded in vivo in anesthetized rats. Neuropsychobiology 36, 96–99. Hulbert, A.J., 2003. Life, death and membrane bilayers. J. Exp. Biol. 206 (Pt 14),
Glass, M., Dragunow, M., Faull, R.L., 1997. Cannabinoid receptors in the human 2303–2311.
brain: a detailed anatomical and quantitative autoradiographic study in the fetal, Hutchins, H.L., Li, Y., Hannon, K., Watkins, B.A., 2011. Eicosapentaenoic acid
neonatal and adult human brain. Neuroscience 77, 299–318. decreases expression of anandamide synthesis enzyme and cannabinoid recep-
Gold, M.S., Frost-Pineda, K., Jacobs, W.S., 2003. Overeating, binge eating, and eating tor 2 in osteoblast-like cells. J. Nutr. Biochem. 22, 195–200.
disorders as addictions. Psychiatr. Ann. 33, 117–122. Ifland, J., Preuss, H., Marcus, M., Rourke, K., Taylor, W., Burau, K., Jacobs, W.S., Kadish,
Greenberg, D., Smith, G.P., 1996. The controls of fat intake. Psychosom. Med. 58, W., Manso, G., 2009. Refined food addiction: a classic substance use disorder.
559–569. Med. Hypotheses 72, 518–526.
Griebel, G., Stemmelin, J., Scatton, B., 2005. Effects of the cannabinoid CB1 recep- Ishiguro, H., Carpio, O., Horiuchi, Y., Shu, A., Higuchi, S., Schanz, N., Benno, R., Ari-
tor antagonist rimonabant in models of emotional reactivity in rodents. Biol. nami, T., Onaivi, E.S., 2010. A non synonymous polymorphism in cannabinoid
Psychiatry 57, 261–267. CB2 receptor gene is associated with eating disorders in humans and food intake
Gross, H., Ebert, M.H., Faden, V.B., Goldberg, S.C., Kaye, W.H., Caine, E.D., Hawks, is modified in mice by its ligands. Synapse 64, 92–96.
R., Zinberg, N., 1983. A double-blind trial of delta 9-tetrahydrocannabinol in Jamshidi, N., Taylor, D.A., 2001. Anandamide administration into the ventromedial
primary anorexia nervosa before and during weight recovery. Am. J. Clin. Psy- hypothalamus stimulates appetite in rats. Br. J. Pharmacol. 134, 1151–1154.
chopharmacol. 3, 165–171. Jauch-Chara, K., Oltmanns, K.M., 2014. Obesity A neuropsychological disease? Sys-
Gutierrez-Lopez, M.D., Llopis, N., Feng, S., Barrett, D.A., O’Shea, E., Colado, M.I., tematic review and neuropsychological model. Prog. Neurobiol. 114C, 84–101.
2010. Involvement of 2-arachidonoyl glycerol in the increased consumption of Jastreboff, A.M., Sinha, R., Lacadie, C., Small, D.M., Sherwin, R.S., Potenza, M.N., 2013.
and preference for ethanol of mice treated with neurotoxic doses of metham- Neural correlates of stress- and food cue-induced food craving in obesity: asso-
phetamine. Br. J. Pharmacol. 160, 772–783. ciation with insulin levels. Diabetes Care 36, 394–402.
Hájos, N., Kathuria, S., Dinh, T., Piomelli, D., Freund, T.F., 2004. Endocannabinoid Jo, Y.H., Chen, Y.J., Chua Jr., S.C., Talmage, D.A., Role, L.W., 2005. Integration of endo-
transport tightly controls 2-arachidonoyl glycerol actions in the hippocampus: cannabinoid and leptin signaling in an appetite-related neural circuit. Neuron
effects of low temperature and the transport inhibitor AM404. Eur. J. Neurosci. 48, 1055–1066.
19, 2991–2996. Johnson, P.M., Kenny, P.J., 2010. Dopamine D2 receptors in addiction like reward
Hansen, H.S., Rosenkilde, M.M., Holst, J.J., Schwartz, T.W., 2012. GPR119 as a fat dysfunction and compulsive eating in obese rats. Nat. Neurosci. 13, 635–641.
sensor. Trends Pharmacol. Sci. 33, 374–381. Jumpertz, R., Guijarro, A., Pratley, R.E., Mason, C.C., Piomelli, D., Krakoff, J., 2012.
Hansen, S.L., Nielsen, A.H., Knudsen, K.E., Artmann, A., Petersen, G., Kristiansen, Associations of fatty acids in cerebrospinal fluid with peripheral glucose con-
U., Hansen, S.H., Hansen, H.S., 2009. Ketogenic diet is antiepileptogenic in centrations and energy metabolism. PLoS ONE 7, e41503.
pentylenetetrazole kindled mice and decrease levels of N-acylethanolamines Jung, K.M., Clapper, J.R., Fu, J., D’Agostino, G., Guijarro, A., Thongkham, D., Avane-
in hippocampus. Neurochem. Int. 54, 199–204. sian, A., Astarita, G., DiPatrizio, N.V., Frontini, A., Cinti, S., Diano, S., Piomelli, D.,
Hanus, L., Abu-Lafi, S., Fride, E., Breuer, A., Vogel, Z., Shalev, D.E., Kustanovich, I., 2012. 2-Arachidonoylglycerol signaling in forebrain regulates systemic energy
Mechoulam, R., 2001. 2-Arachidonyl glyceryl ether, an endogenous agonist of metabolism. Cell Metab. 15, 299–310.
the cannabinoid CB1 receptor. Proc. Natl. Acad. Sci. U.S.A. 98, 3662–3665. Justinova, Z., Goldberg, S.R., Heishman, S.J., Tanda, G., 2005. Self-administration of
Hao, S., Avraham, Y., Mechoulam, R., Berry, E.M., 2000. Low dose anandamide affects cannabinoids by experimental animals and human marijuana smokers. Pharma-
food intake, cognitive function, neurotransmitter and corticosterone levels in col. Biochem. Behav. 81, 285–299.
diet-restricted mice. Eur. J. Pharmacol. 392, 147–156. Kaczocha, M., Glaser, S.T., Deutsch, D.G., 2009. Identification of intracellular car-
Harris, W.S., Mozaffarian, D., Rimm, E., Kris-Etherton, P., Rudel, L.L., Appel, L.J., Engler, riers for the endocannabinoid anandamide. Proc. Natl. Acad. Sci. U.S.A. 106,
M.M., Engler, M.B., Sacks, F., 2009. Omega-6 fatty acids and risk for cardiovascu- 6375–6380.
lar disease: a science advisory from the American Heart Association Nutrition Kaczocha, M., Vivieca, S., Sun, J., Glaser, S.T., Deutsch, D.G., 2012. Fatty acid-binding
Subcommittee of the Council on Nutrition, Physical Activity, and Metabolism; proteins transport N-acylethanolamines to nuclear receptors and are targets of
Council on Cardiovascular Nursing; and Council on Epidemiology and Preven- endocannabinoid transport inhibitors. J. Biol. Chem. 287, 3415–3424.
tion. Circulation 119, 902–907. Kelly, T., Yang, W., Chen, C.S., Reynolds, K., He, J., 2008. Global burden of obesity in
Harrold, J.A., Elliott, J.C., King, P.J., Widdowson, P.S., Williams, G., 2002. Down- 2005 and projections to 2030. Int. J. Obes. 32, 1431–1437.
regulation of cannabinoid-1 (CB-1) receptors in specific extrahypothalamic Kenny, P.J., 2011. Reward mechanisms in obesity: new insights and future directions.
regions of rats with dietary obesity: a role for endogenous cannabinoids in Neuron 69, 664–679.
driving appetite for palatable food? Brain Res. 952, 232–238. Kenny, P.J., Chen, S.A., Kitamura, O., Markou, A., Koob, G.F., 2006. Conditioned with-
Hein, I.M., Huyser, C., 2010. Olanzapine in the treatment of adolescents with anorexia drawal drives heroin consumption and decreases reward sensitivity. J. Neurosci.
nervosa. Tijdschr. Psychiatr. 52, 417–421. 26, 5894–5900.
Hentges, S.T., Low, M.J., Williams, J.T., 2005. Differential regulation of synaptic inputs Kenny, P.J., Voren, G., Johnson, P.M., 2013. Dopamine D2 receptors and striatopallidal
by constitutively released endocannabinoids and exogenous cannabinoids. J. transmission in addiction and obesity. Curr. Opin. Neurobiol. 23, 535–538.
Neurosci. 25, 9746–9751. Kipnes, M.S., Hollander, P., Fujioka, K., Gantz, I., Seck, T., Erondu, N., Shentu, Y., Lu,
Herkenham, M., Lynn, A.B., Johnson, M.R., Melvin, L.S., de Costa, B.R., Rice, K.C., K., Suryawanshi, S., Chou, M., Johnson-Levonas, A.O., Heymsfield, S.B., Shapiro,
1991. Characterization and localization of cannabinoid receptors in rat brain: D., Kaufman, K.D., Amatruda, J.M., 2010. A one-year study to assess the safety
a quantitative in vitro autoradiographic study. J. Neurosci. 11, 563–583. and efficacy of the CB1R inverse agonist taranabant in overweight and obese
Higgs, S., Barber, D.J., Cooper, A.J., Terry, P., 2005. Differential effects of two cannabi- patients with type 2 diabetes. Diabetes Obes. Metab. 12, 517–531.
noid receptor agonists on progressive ratio responding for food and free-feeding Kirkham, T.C., Williams, C.M., Fezza, F., Di Marzo, V., 2002. Endocannabinoid levels
in rats. Behav. Pharmacol. 16, 389–393. in rat limbic forebrain and hypothalamus in relation to fasting, feeding and sati-
Higuchi, S., Irie, K., Yamaguchi, R., Katsuki, M., Araki, M., Ohji, M., Hayakawa, K., ation: stimulation of eating by 2-arachidonoyl glycerol. Br. J. Pharmacol. 136,
Mishima, S., Akitake, Y., Matsuyama, K., Mishima, K., Mishima, K., Iwasaki, K., 550–557.
Fujiwara, M., 2012. Hypothalamic 2-arachidonoylglycerol regulates multistage Klein, T.W., Newton, C., Larsen, K., Lu, L., Perkins, I., Nong, L., Friedman, H., 2003. The
process of high-fat diet preferences. PLoS ONE 7, e38609. cannabinoid system and immune modulation. J. Leukoc. Biol. 74, 486–496.
Higuchi, S., Ohji, M., Araki, M., Furuta, R., Katsuki, M., Yamaguchi, R., Akitake, Y., Kola, B., Farkas, I., Christ-Crain, M., Wittmann, G., Lolli, F., Amin, F., Harvey-White,
Matsuyama, K., Irie, K., Mishima, K., Mishima, K., Iwasaki, K., Fujiwara, M., 2011. J., Liposits, Z., Kunos, G., Grossman, A.B., Fekete, C., Korbonits, M., 2008. The
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 221

orexigenic effect of ghrelin is mediated through central activation of the endoge- Mahler, S.V., Smith, K.S., Berridge, K.C., 2007. Endocannabinoid hedonic hots pot for
nous cannabinoid system. PLoS ONE 3, e1797. sensory pleasure:anandamide in nucleus accumbens shell enhances “liking” of
Koob, G., 1992. Neural mechanisms of drug reinforcement. Ann. N. Y. Acad. Sci. 654, as weet reward. Neuropsychopharmacology 32, 2267–2278.
171–191. Malcher-Lopes, R., Di, S., Marcheselli, V.S., Weng, F.J., Stuart, C.T., Bazan, N.G.,
Koob, G.F., Le Moal, M., 1997. Science 278, 52–58. Tasker, J.G., 2006. Opposing crosstalk between leptin and glucocorticoids rapidly
Koob, G.F., Le Moal, M., 2005. Plasticity of reward neurocircuitry and the ‘dark side’ modulates synaptic excitation via endocannabinoid release. J. Neurosci. 26,
of drug addiction. Nat. Neurosci. 8, 1442–1444. 6643–6650.
Korte, G., Dreiseitel, A., Schreier, P., Oehme, A., Locher, S., Geiger, S., Heilmann, J., Maldonado, R., Berrendero, F., Ozaita, A., Robledo, P., 2011. Neurochemical basis of
Sand, P.G., 2010. Tea catechins’ affinity for human cannabinoid receptors. Phy- cannabis addiction. Neuroscience 181, 1–17.
tomedicine 17, 19–22. Maldonado, R., Valverde, O., Berrendero, F., 2006. Involvement of the endocannabi-
Korte, G., Dreiseitel, A., Schreier, P., Oehme, A., Locher, S., Hajak, G., Sand, P.G., 2009. noid system in drug addiction. Trends Neurosci. 29, 225–232.
An examination of anthocyanins’ and anthocyanidins’ affinity for cannabinoid Marco, E.M., Romero-Zerbo, S.Y., Viveros, M.P., Bermudez-Silva, F.J., 2012. The role of
receptors. J. Med. Food 12, 1407–1410. the endocannabinoid system in eating disorders: pharmacological implications.
Lam, T.K., Pocai, A., Gutierrez-Juarez, R., Obici, S., Bryan, J., et al., 2005. Hypothalamic Behav. Pharmacol. 23, 526–536.
sensing of circulating fatty acids is required for glucose homeostasis. Nat. Med. Marinelli, S., Di Marzo, V., Berretta, N., Matias, I., Maccarrone, M., Bernardi, G.,
11, 320–327. Mercuri, N.B., 2003. Presynaptic facilitation of glutamatergic synapses to dopa-
Lambert, D.M., Di Marzo, V., 1999. The palmitoylethanolamide and oleamide enig- minergic neurons of the rat substantia nigra by endogenous stimulation of
mas: are these two fatty acid amides cannabimimetic? Curr. Med. Chem. 6, vanilloid receptors. J. Neurosci. 23, 3136–3144.
757–773. Markou, A., Koob, G.F., 1991. Postcocaine anhedonia. An animal model of cocaine
Lambert, D.M., Vandevoorde, S., Diependaele, G., Govaerts, S.J., Robert, A.R., 2001. withdrawal. Neuropsychopharmacology 4, 17–26.
Anticonvulsant activity of N-palmitoylethanolamide, a putative endocannabi- Martin-Fardon, R., Zorrilla, E.P., Ciccocioppo, R., Weiss, F., 2010. Role of innate
noid, in mice. Epilepsia 42, 321–327. and drug-induced dysregulation of brain stress and arousal systems in addic-
Lauritzen, L., Hansen, H.S., Jørgensen, M.H., Michaelsen, K.F., 2001. The essentiality tion: focus on corticotropin-releasing factor, nociceptin/orphanin FQ, and
of long chain n-3 fatty acids in relation to development and function of the brain orexin/hypocretin. Brain Res. 16 (1314), 145–161.
and retina. Prog. Lipid Res. 40, 1–94. Massi, L., Elezgarai, I., Puente, N., Reguero, L., Grandes, P., Manzoni, O.J., Georges,
Le Moal, M., Koob, G.F., 2007. Drug addiction: pathways to the disease and patho- F., 2008. Cannabinoid receptors in the bed nucleus of the stria terminalis con-
physiological perspectives. Eur. Neuropsychopharmacol. 17, 377–393. trol cortical excitation of midbrain dopamine cells in vivo. J. Neurosci. 28,
Lecca, D., Cacciapaglia, F., Valentini, V., Di Chiara, G., 2006. Monitoring extracellular 10496–10508.
dopamine in the rat nucleus accumbens shell and core during acquisition and Massiera, F., Barbry, P., Guesnet, P., Joly, A., Luquet, S., Moreilhon-Brest, C., Mohsen-
maintenance of intravenous WIN55,212-2 self-administration. Psychopharma- Kanson, T., Amri, E.Z., Ailhaud, G., 2010. A western-like fat diet is sufficient to
cology (Berl.) 188, 63–74. induce a gradual enhancement in fat mass over generations. J. Lipid Res. 51,
Lenoir, M., Serre, F., Cantin, L., Ahmed, S., 2007. Intense sweetness surpasses cocaine 2352–2361.
reward. PLoS ONE 2, 698. Massiera, F., Saint-Marc, P., Seydoux, J., Murata, T., Kobayashi, T., Narumiya, S., Gues-
Leung, K., Elmes, M.W., Glaser, S.T., Deutsch, D.G., Kaczocha, M., 2013. Role of FAAH- net, P., Amri, E.Z., Negrel, R., Ailhaud, G., 2003. Arachidonic acid and prostacyclin
like anandamide transporter in anandamide inactivation. PLoS ONE 8, e79355. signaling promote adipose tissue development: a human health concern? J. Lipid
Li, C., Bowe, J.E., Huang, G.C., Amiel, S.A., Jones, P.M., Persaud, S.J., 2011b. Cannabi- Res. 44, 271–279.
noid receptor agonists and antagonists stimulate insulin secretion from isolated Mastinu, A., Pira, M., Pani, L., Pinna, G.A., Lazzari, P., 2012. NESS038C6, a novel selec-
human islets of Langerhans. Diabetes Obes. Metab. 13, 903–910. tive CB1 antagonist agent with anti-obesity activity and improved molecular
Li, C., Jones, P.M., Persaud, S.J., 2011a. Role of the endocannabinoid system in food profile. Behav. Brain Res. 234, 192–204.
intake, energy homeostasis and regulation of the endocrine pancreas. Pharma- Matias, I., Di Marzo, V., 2007. Endocannabinoids and the control of energy balance.
col. Ther. 129, 307–320. Trends Endocrinol. Metab. 18, 27–37.
Lindborg, K.A., Teachey, M.K., Jacob, S., Henriksen, E.J., 2010. Effects of in vitro Matias, I., Gatta-Cherifi, B., Tabarin, A., Clark, S., Leste-Lasserre, T., Marsicano, G.,
antagonism of endocannabinoid-1 receptors on the glucose transport system Piazza, P.V., Cota, D., 2012. Endocannabinoids measurement in human saliva as
in normal and insulin-resistant rat skeletal muscle. Diabetes Obes. Metab. 12, potential biomarker of obesity. PLoS ONE 7, e42399.
722–730. Matias, I., Gonthier, M.P., Orlando, P., Martiadis, V., De Petrocellis, L., Cervino, C.,
Liu, Q.R., Pan, C.H., Hishimoto, A., Li, C.Y., Xi, Z.X., Llorente-Berzal, A., Viveros, Petrosino, S., Hoareau, L., Festy, F., Pasquali, R., Roche, R., Maj, M., Pagotto, U.,
M.P., Ishiguro, H., Arinami, T., Onaivi, E.S., Uhl, G.R., 2009. Species differences Monteleone, P., Di Marzo, V., 2006. Regulation, function, and dysregulation of
in cannabinoid receptor 2 (CNR2 gene): identification of novel human and endocannabinoids in models of adipose and beta-pancreatic cells and in obesity
rodent CB2 isoforms, differential tissue expression and regulation by cannabi- and hyperglycemia. J. Clin. Endocrinol. Metab. 91, 3171–3180.
noid receptor ligands. Genes Brain Behav. 8, 519–530. Matias, I., Gonthier, M.P., Petrosino, S., Docimo, L., Capasso, R., Hoareau, L., Mon-
Lo Verme, J., Gaetani, S., Fu, J., Oveisi, F., Burton, K., Piomelli, D., 2005. Regulation of teleone, P., Roche, R., Izzo, A.A., Di Marzo, V., 2007. Role and regulation of
food intake by oleoylethanolamide. Cell. Mol. Life Sci. 62, 708–716. acylethanolamides in energy balance: focus on adipocytes and beta-cells. Br.
Lockie, S.H., Czyzyk, T.A., Chaudhary, N., Perez-Tilve, D., Woods, S.C., Oldfield, B.J., J. Pharmacol. 152, 676–690.
Statnick, M.A., Tschop, M.H., 2011. CNS opioid signaling separates cannabinoid Matias, I., Leonhardt, M., Lesage, J., De Petrocellis, L., Dupouy, J.P., Vieau,
receptor 1-mediated effects on body weight and mood-related behavior in mice. D., Di Marzo, V., 2003. Effect of maternal under-nutrition on pup body
Endocrinology 152, 3661–3667. weight and hypothalamic endocannabinoid levels. Cell. Mol. Life Sci. 60,
LoVerme, J., Duranti, A., Tontini, A., Spadoni, G., Mor, M., Rivara, S., Stella, N., Xu, C., 382–389.
Tarzia, G., Piomelli, D., 2009. Synthesis and characterization of a peripherally Matias, I., Petrosino, S., Racioppi, A., Capasso, R., Izzo, A.A., Di Marzo, V., 2008a. Dys-
restricted CB1 cannabinoid antagonist, URB447, that reduces feeding and body- regulation of peripheral endocannabinoid levels in hyperglycemia and obesity:
weight gain inmice. Bioorg. Med. Chem. Lett. 19, 639–643. effect of high fat diets. Mol. Cell. Endocrinol. 286, S66–S78.
Lüscher, C., Malenka, R., 2011. Drug-evoked synaptic plasticity in addiction: from Matias, I., Vergoni, A.V., Petrosino, S., Ottani, A., Pocai, A., Bertolini, A., Di Marzo,
molecular changes to circuit remodeling. Neuron 69, 650–663. V., 2008b. Regulation of hypothalamic endocannabinoid levels by neuro-
Maccarrone, M., Dainese, E., Oddi, S., 2010a. Intracellular trafficking of anandamide: peptides and hormones involved in food intake and metabolism: insulin and
new concepts for signaling. Trends Biochem. Sci. 35, 601–608. melanocortins. Neuropharmacology 54, 206–212.
Maccarrone, M., Di Rienzo, M., Finazzi-Agrò, A., Rossi, A., 2003a. Leptin activates Matsuda, L.A., Lolait, S.J., Brownstein, M.J., Young, A.C., Bonner, T.I., 1990. Structure
the anandamide hydrolase promoter in human T lymphocytes through STAT3. of a cannabinoid receptor and functional expression of the cloned cDNA. Nature
J. Biol. Chem. 278, 13318–13324. 346, 561–564.
Maccarrone, M., Gasperi, V., Catani, M.V., Diep, T.A., Dainese, E., Hansen, H.S., Max, M., Shanker, Y.G., Huang, L., Rong, M., Liu, Z., Campagne, F., Weinstein, H.,
Avigliano, L., 2010b. The endocannabinoid system and its relevance for nutrition. Damak, S., Margolskee, R.F., 2001. Tas1r3, encoding a new candidate taste
Annu. Rev. Nutr. 30, 423–440. receptor, is allelic to the sweet responsiveness locus Sac. Nat. Genet. 28,
Maccarrone, M., Bari, M., Battista, N., Di Rienzo, M., Finazzi-Agrò, A., 2001. Endoge- 58–63.
nous cannabinoids in neuronal and immune cells: toxic effects, levels and McAleavey, K.M.A.F.M., 2001. Eating disorders: are they addictions? A dialogue. J.
degradation. Funct. Neurol. 16, 53–60. Soc. Work Pract. Addict. 1, 107–113.
Maccarrone, M., Bari, M., Di Rienzo, M., Finazzi-Agrò, A., Rossi, A., 2003b. Pro- McElroy, S.L., Shapira, N.A., Arnold, L.M., Keck, P.E., Rosenthal, N.R., Wu, S.C., Capece,
gesterone activates fatty acid amide hydrolase (FAAH) promoter in human T J.A., Fazzio, L., Hudson, J.I., 2004. Topiramate in the long-term treatment of binge
lymphocytes through the transcription factor Ikaros. Evidence for a synergistic eating disorder associated with obesity. J. Clin. Psychiatry 65, 1453–1469.
effect of leptin. J. Biol. Chem. 278, 32726–32732. Mechoulam, R., 1970. Marihuana chemistry. Science 168, 1159–1166.
Maccarrone, M., Pauselli, R., Di Rienzo, M., Finazzi-Agrò, A., 2002. Binding, degra- Mechoulam, R., Ben-Shabat, S., Hanus, L., Ligumsky, M., Kaminski, N.E., Schat, A.R.,
dation and apoptotic activity of stearoylethanolamide in rat C6 glioma cells. Gopher, A., Almog, S., Martin, B.R., Compton, D.R., Pertwee, R.G., Griffin, G.,
Biochem. J. 366, 137–144. Bayewitch, M., Barg, J., Zvi Vogel, Z., 1995. Identification of an endogenous
Maccarrone, M., 2004. Sex and drug abuse: a role for retrograde endocannabinoids? 2-monoglyceride, present in canine gut, that binds to cannabinoid receptors.
Trends Pharmacol. Sci. 25, 455–456. Biochem. Pharmacol. 50, 83–90.
Maccioni, P., Pes, D., Carai, M.A., Gessa, G.L., Colombo, G., 2008. Suppression by the Mechoulam, R., Gaoni, Y., 1965. A total synthesis of dl-delta-1 tetrahydrocannabinol,
cannabinoid CB1 receptor antagonist, rimonabant, of the reinforcing and moti- the active constituent of hashish. J. Am. Chem. Soc. 87, 3273–3275.
vational properties of achocolate-flavoured beverageinrats. Behav. Pharmacol. Melis, M., Muntoni, A.L., Pistis, M., 2012. Endocannabinoids and the processing of
19, 197–209. value-related signals. Front. Pharmacol. 3, 7.
222 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

Melis, M., Pistis, M., 2007. Endocannabinoid signaling in midbrain dopamine neu- Oddi, S., Fezza, F., Pasquariello, N., D’Agostino, A., Catanzaro, G., De Simone, C.,
rons: more than physiology? Curr. Neuropharmacol. 5, 268–277. Rapino, C., Finazzi-Agrò, A., Maccarrone, M., 2009. Molecular identification of
Melis, M., Pistis, M., Perra, S., Muntoni, A.L., Pillolla, G., Gessa, G.L., 2004. Endo- albumin and Hsp70 as cytosolic anandamide-binding proteins. Chem. Biol. 16,
cannabinoids mediate presynaptic inhibition of glutamatergic transmission in 624–632.
rat ventral tegmental area dopamine neurons through activation of CB1 recep- Oddi, S., Fezza, F., Pasquariello, N., De Simone, C., Rapino, C., Dainese, E., Finazzi-
tors. J. Neurosci. 24, 53–62. Agrò, A., Maccarrone, M., 2008. Evidence for the intracellular accumulation of
Meule, A., 2011. How prevalent is “food addiction”? Front. Psychiatry 2, 61. anandamide in adiposomes. Cell. Mol. Life Sci. 65, 840–850.
Meyer, F., 2008. Alleviation of both binge eating and sexual dysfunction with nal- Okamoto, Y., Morishita, J., Tsuboi, K., Tonai, T., Ueda, N., 2004. Molecular character-
trexone. J. Clin. Psychopharmacol. 28, 722–723. ization of a phospholipase D generating anandamide and its congeners. J. Biol.
Micioni Di Bonaventura, M.V., Ciccocioppo, R., Romano, A., Bossert, J.M., Rice, K.C., Chem. 279, 5298–5305.
Ubaldi, M.St., Laurent, R., Gaetani, S., Massi, M., Shaham, Y., Cifani, C., 2014. Oleson, E.B., Cachope, R., Fitoussi, A., Tsutsui, K., Wu, S., Gallegos, J.A., Cheer, J.F.,
Role of bed nucleus of the stria terminalis corticotrophin-releasing factor recep- 2014. Cannabinoid receptor activation shifts temporally engendered patterns
tors in frustration stress-induced binge-like palatable food consumption in of dopamine release. Neuropsychopharmacology 39, 1441–1452.
female rats with a history of food restriction. J. Neurosci. 34, 11316–11324, Osei-Hyiaman, D., De Petrillo, M., Pacher, P., Liu, J., Radaeva, S., Bátkai, S.,
http://dx.doi.org/10.1523/JNEUROSCI.1854-14.2014. Harvey-White, J., Mackie, K., Offertáler, L., Wang, L., Kunos, G., 2005.
Micioni Di Bonaventura, M.V., Cifani, C., Lambertucci, C., Volpini, R., Cristalli, G., Endocannabinoid activation at hepatic CB1 receptors stimulates fatty acid
Massi, M., 2012a. A2A adenosine receptor agonists reduce both high-palatability synthesis and contributes to diet-induced obesity. J. Clin. Invest. 115,
and low-palatability food intake in female rats. Behav. Pharmacol. 23, 1298–1305.
567–574. Osei-Hyiaman, D., Liu, J., Zhou, L., Godlewski, G., Harvey-White, J., Jeong, W.I., Bátkai,
Micioni Di Bonaventura, M.V., Vitale, G., Massi, M., Cifani, C., 2012b. Effect of hyper- S., Marsicano, G., Lutz, B., Buettner, C., Kunos, G., 2008. Hepatic CB1 receptor is
icum perforatum extract in an experimental model of binge eating in female required for development of diet-induced steatosis, dyslipidemia, and insulin
rats. J. Obes. 2012, 956137. and leptin resistance in mice. J. Clin. Invest. 118, 3160–3169.
Monteleone, P., Bifulco, M., DiFilippo, C., Gazzerro, P., Canestrelli, B., Monteleone, F., Oswald, K.D., Murdaugh, D.L., King, V.L., Boggiano, M.M., 2011. Motivation for palat-
Proto, M.C., Di Genio, M., Grimaldi, C., Maj, M., 2009. Association of CNR1 and able food despite consequences in an animal model of binge eating. Int. J. Eat.
FAAH endocannabinoid gene polymorphisms with anorexia nervosa and bulimia Disord. 44, 203–211.
nervosa: evidence for synergistic effects. Genes Brain Behav. 8, 728–732. Overton, H.A., Fyfe, M.C., Reynet, C., 2008. GPR119, a novel G protein-coupled recep-
Monteleone, P., Matias, I., Martiadis, V., De Petrocellis, L., Maj, M., Di Marzo, V., tor target for the treatment of type 2 diabetes and obesity. Br. J. Pharmacol. 153,
2005. Blood levels of the endocannabinoid anandamide are increased in anorexia S76–S81.
nervosa and in binge-eating disorder, but not in bulimia nervosa. Neuropsy- Pagano, C., Rossato, M., Vettor, R., 2008. Endocannabinoids, adipose tissue and lipid
chopharmacology 30, 1216–1221. metabolism. J. Neuroendocrinol. 1, 124–129.
Monteleone, P., Maj, M., 2013. Dysfunctions of leptin, ghrelin, BDNF and endo- Pagotto, U., Marsicano, G., Cota, D., Lutz, B., Pasquali, R., 2006. The emerging role
cannabinoids in eating disorders: beyond the homeostatic control of food intake. of the endocannabinoid system in endocrine regulation and energy balance.
Psychoneuroendocrinology 38, 312–330. Endocr. Rev. 27, 73–100.
Monteleone, P., Piscitelli, F., Scognamiglio, P., Monteleone, A.M., Canestrelli, B., Di Parolaro, D., Rubino, T., Viganò, D., Massi, P., Guidali, C., Realini, N., 2010. Cellu-
Marzo, V., Maj, M., 2012. Hedonic eating is associated with increased peripheral lar mechanisms underlying the interaction between cannabinoid and opioid
levels of ghrelin and the endocannabinoid 2-arachidonoyl-glycerol in healthy system. Curr. Drug Targets 11, 393–405.
humans: a pilot study. J. Clin. Endocrinol. Metab. 97, E917–E924. Parylak, S.L., Koob, G.F., Zorrilla, E.P., 2011. The dark side of food addiction. Physiol.
Monteleone, P., Tortorella, A., Martiadis, V., DiFilippo, C., Canestrelli, B., Maj, M., 2008. Behav. 104, 149–156.
The cDNA 385C to A missense polymorphism of the endocannabinoid degrading Pavón, F.J., Serrano, A., Pérez-Valero, V., Jagerovic, N., Hernández-Folgado, L.,
enzymefattyacidamidehydrolase (FAAH) is associated with over-weight/obesity Bermúdez-Silva, F.J., Macías, M., Goya, P., de Fonseca, F.R., 2008. Central versus
but not with binge eating disorder in overweight/obese women. Psychoneu- peripheral antagonism of cannabinoid CB1 receptor in obesity: effects of LH-21,
roendocrinology 33, 546–550. a peripherally acting neutral cannabinoid receptor antagonist, in Zucker rats. J.
Montmayeur, J.P., Liberles, S.D., Matsunami, H., Buck, L.B., 2001. A candidate taste Neuroendocrinol. 20, 116–123.
receptor gene near a sweet taste locus. Nat. Neurosci. 4, 492–498. Pedram, P., Wadden, D., Amini, P., Gulliver, W., Randell, E., Cahill, F., Vasdev, S.,
Morrison, C.D., Berthoud, H.R., 2007. Neurobiology of nutrition and obesity. Nutr. Goodridge, A., Carter, J.C., Zhai, G., Ji, Y., Sun, G., 2013. Food addiction: its preva-
Rev. 65, 517–534. lence and significant association with obesity in the general population. PLoS
Muccioli, G.G., Naslain, D., Bäckhed, F., Reigstad, C.S., Lambert, D.M., Delzenne, N.M., ONE 4 (8(9)), e74832.
Cani, P.D., 2010. The endocannabinoid system links gut microbiota to adipogen- Pelchat, M.L., 2002. Of human bondage: food craving, obsession, compulsion, and
esis. Mol. Syst. Biol. 6, 392. addiction. Physiol. Behav. 76, 389–395.
Muller, T.D., Reichwald, K., Bronner, G., Kirschner, J., Nguyen, T.T., Scherag, A., Herzog, Pertwee, R.G., 2008. Ligands that target cannabinoid receptors in the brain: from
W., Herpertz-Dahlmann, B., Lichtner, P., Meitinger, T., Platzer, M., Schafer, H., THC to anandamide and beyond. Addict. Biol. 13, 147–159.
Hebebrand, J., Hinney, A., 2008. Lack of association of genetic variants in genes Pi-Sunyer, F., Aronne, L.J., Heshmati, H.M., Devin, J., Rosenstock, J., 2006. Effect of
of the endocannabinoid system with anorexia nervosa. Child Adolesc. Psychiatry rimonabant, a Cannabinoid-1 Receptor blocker, on weight and cardiometabolic
Ment. Health 2, 33. risk factors in overweight or obese PatientsRIO-North America: a randomized
Munro, S., Thomas, K.L., Abu-Shaar, M., 1993. Molecular characterization of a periph- controlled trial. J. Am. Med. Assoc. 15, 761–1252.
eral receptor for cannabinoids. Nature 2, 61–65. Piccoli, L., Micioni Di Bonaventura, M.V., Cifani, C., Costantini, V.J., Massagrande,
Naruse, T., Amano, H., Koizumi, Y., 1991. Possible involvement of dopamine D- M., Montanari, D., Martinelli, P., Antolini, M., Ciccocioppo, R., Massi, M., Merlo-
1 and D-2 receptors in diazepam-induced hyperphagia in rats. Fundam. Clin. Pich, E., Di Fabio, R., Corsi, M., 2012. Role of orexin-1 receptor mechanisms on
Pharmacol. 5, 677–693. compulsive food consumption in a model of binge eating in female rats. Neu-
Navarro, M., Chowen, J., Rocio, A., Carrera, M., del Arco, I., Villanua, M.A., Martin, ropsychopharmacology 37, 1999–2011.
Y., Roberts, A.J., Koob, G.F., de Fonseca, F.R., 1998. CB1 cannabinoid receptor Piomelli, D., 2005. The endocannabinoid system: a drug discovery perspective. Curr.
antagonist-induced opiate withdrawal in morphine-dependent rats. Neurore- Opin. Investig. Drugs 6, 672–679.
port 9, 3397–3402. Piomelli, D., 2013. A fatty gut feeling. Trends Endocrinol. Metab. 24, 332–341.
Newberry, E.P., Kennedy, S.M., Xie, Y., Luo, J., Crooke, R.M., Graham, M.J., Fu, J., Porter, A.C., Sauer, J.M., Knierman, M.D., Becker, G.W., Berna, M.J., Bao, J., Nomikos,
Piomelli, D., Davidson, N.O., 2012. Decreased body weight and hepatic steatosis G.G., Carter, P., Bymaster, F.P., Leese, A.B., Felder, C.C., 2002. Characterization
with altered fatty acid ethanolamide metabolism in aged L-Fabp-/- mice. J. Lipid of a novel endocannabinoid, virodhamine, with antagonist activity at the CB1
Res. 53, 744–754. receptor. J. Pharmacol. Exp. Ther. 301, 1020–1024.
Ng Cheong Ton, J.M., Gardner, E.L., 1986. Effects of delta-9-tetrahydrocannabinol on Provensi, G., Coccurello, R., Umehara, H., Munari, L., Giacovazzo, G., Galeotti, N., Nosi,
dopamine release in the brain: intracranial dialysis experiments. Soc. Neurosci. D., Gaetani, S., Romano, A., Moles, A., Blandina, P., Passani, M.B., 2014. Satiety
Abstr. 12, 135. factor oleoylethanolamide recruits the brain histaminergic system to inhibit
Nissen, S.E., Nicholls, S.J., Wolski, K., Rodés-Cabau, J., Cannon, C.P., Deanfield, J.E., food intake. Proc. Natl. Acad. Sci. U. S. A. 111, 11527–11532.
Després, J.P., Kastelein, J.J., Steinhubl, S.R., Kapadia, S., Yasin, M., Ruzyllo, W., Quarta, C., Bellocchio, L., Mancini, G., Mazza, R., Cervino, C., Braulke, L.J., Fekete,
Gaudin, C., Job, B., Hu, B., Bhatt, D.L., Lincoff, A.M., Tuzcu, E.M., 2008. Effect of C., Latorre, R., Nanni, C., Bucci, M., Clemens, L.E., Heldmaier, G., Watanabe, M.,
rimonabant on progression of atherosclerosis in patients with abdominal obe- Leste-Lassere, T., Maitre, M., Tedesco, L., Fanelli, F., Reuss, S., Klaus, S., Srivastava,
sity and coronary artery disease: the STRADIVARIUS randomized controlled trial. R.K., Monory, K., Valerio, A., Grandis, A., De Giorgio, R., Pasquali, R., Nisoli, E.,
J. Am. Med. Assoc. 299, 1547–1560. Cota, D., Lutz, B., Marsicano, G., Pagotto, U., 2010. CB (1) signaling in forebrain
Nong, L., Newton, C., Friedman, H., Klein, T.W., 2001. CB1 and CB2 receptor mRNA and sympathetic neurons is a key determinant of endocannabinoid actions on
expression in human peripheral blood mononuclear cells (PBMC) from various energy balance. Cell Metab. 11, 273–285.
donor types. Adv. Exp. Med. Biol. 493, 229–233. Randolph, T.G., 1956. The descriptive features of food addiction addictive eating and
Nunez, E., Benito, C., Pazos, M.R., Barbachano, A., Fajardo, O., González, S., Tolón, drinking. Q. J. Stud. Alcohol 17, 198–224.
R.M., Romero, J., 2004. Cannabinoid CB2 receptors are expressed by perivascular Riegel, A.C., Lupica, C.R., 2004. Independent presynaptic and postsynaptic mecha-
microglial cells in the human brain: an immunohistochemical study. Synapse 53, nisms regulate endocannabinoid signaling at multiple synapses in the ventral
208–213. tegmentalarea. J. Neurosci. 24, 11070–11078.
O’Keefe, L., Simcocks, A.C., Hryciw, D.H., Mathai, M.L., McAinch, A.J., 2014. The Rigotti, N.A., Gonzales, D., Dale, L.C., Lawrence, D., Chang, Y., 2009. A randomized con-
cannabinoid receptor 1 and its role in influencing peripheral metabolism. Dia- trolled trial of adding the nicotine patch to rimonabant for smoking cessation:
betes Obes. Metab. 16, 294–304. efficacy, safety and weight gain. Addiction 104, 266–276.
C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224 223

Robledo, P., Berrendero, F., Ozaita, A., Maldonado, R., 2008. Advances in the field of Solinas, M., Goldberg, S.R., 2005. Motivational effects of cannabinoids and opioids
cannabinoid-opioid cross-talk. Addict. Biol. 13, 213–224. on food reinforcement depend on simultaneous activation of cannabinoid and
Roche, R., Hoareau, L., Bes-Houtmann, S., Gonthier, M.P., Laborde, C., Baron, J.F., Haf- opioid systems. Neuropsychopharmacology 30, 2035–2045.
faf, Y., Cesari, M., Festy, F., 2006. Presence of the cannabinoid receptors, CB1 and Solinas, M., Justinova, Z., Goldberg, S.R., Tanda, G., 2006. Anandamide administra-
CB2 , in human omental and subcutaneous adipocytes. Histochem. Cell Biol. 126, tion alone and after inhibition of fatty acid amide hydrolase (FAAH) increases
177–187. dopamine levels in the nucleus accumbens shell in rats. J. Neurochem. 98,
Rodriguez de Fonseca, F., Navarro, M., Gomez, R., Escuredo, L., Nava, F., Fu, J., 408–419.
Murillo-Rodriguez, E., Giuffrida, A., LoVerme, J., Gaetani, S., Kathuria, S., Gall, C., Solinas, M., Yasar, S., Goldberg, S.R., 2007. Endocannabinoid system involvement in
Piomelli, D., 2001. An anorexic lipid mediator regulated by feeding. Nature 414, brain reward processes related to drug abuse. Pharmacol. Res. 56, 393–405.
209–212. Soria, E., Matias, I., Di Marzo, Prospero-Garcia, O., 2006. Inhibition of endocannabi-
Romano, A., Cassano, T., Tempesta, B., Cianci, S., Dipasquale, P., Coccurello, R., Cuomo, noid degradation in the nucleus accumbens shell increases food intake and
V., Gaetani, S., 2013a. The satiety signal oleoylethanolamide stimulates oxytocin activates the hypothalamus in rats. In: 36th Neuroscience Annual Meeting,
neurosecretion from rat hypothalamic neurons. Peptides 49C, 21–26. Washington, USA, Poster number 457.17/CC9.
Romano, A., Coccurello, R., Giacovazzo, G., Bedse, G., Moles, A., Gaetani, S., Soria-Gómez, E., Bellocchio, L., Reguero, L., Lepousez, G., Martin, C., Bendahmane, M.,
2014. Oleoylethanolamide: a novel potential pharmacological alternative to Ruehle, S., Remmers, F., Desprez, T., Matias, I., Wiesner, T., Cannich, A., Nissant,
cannabinoid antagonists for the control of appetite. Biomed. Res. Int., 10, A., Wadleigh, A., Pape, H.C., Chiarlone, A.P., Quarta, C., Verrier, D., Vincent, P.,
http://dx.doi.org/10.1155/2014/203425, Article ID 203425. Massa, F., Lutz, B., Guzmán, M., Gurden, H., Ferreira, G., Lledo, P.M., Grandes,
Romano, A., Potes, C.S., Tempesta, B., Cassano, T., Cuomo, V., Lutz, T., Gaetani, P., Marsicano, G., 2014. The endocannabinoid system controls food intake via
S., 2013b. Hindbrain noradrenergic input to the hypothalamic PVN medi- olfactory processes. Nat. Neurosci. 17, 407–415.
ates the activation of oxytocinergic neurons induced by the satiety factor Spoto, B., Fezza, F., Parlongo, G., Battista, N., Sgro’, E., Gasperi, V., Zoccali, C.,
oleoylethanolamide. Am. J. Physiol. Endocrinol. Metab. 305, E1266–E1273. Maccarrone, M., 2006. Human adipose tissue binds and metabolizes the
Rosenstock, J., Hollander, P., Chevalier, S., Iranmanesh, A., 2008. SERENADE: the endocannabinoids anandamide and 2-arachidonoylglycerol. Biochimie 88,
study evaluating rimonabant efficacy in drug-naive diabetic patients: effects of 1889–1897.
monotherapy with rimonabant, the first selective CB1 receptor antagonist, on Spring, B., Schneider, K., Smith, M., Kendzor, D., Appelhans, B., Hedeker, D., Pagoto,
glycemic control, body weight, and lipid profile in drug-naive type 2 diabetes. S., 2008. Abuse potential of carbohydrates for overweight carbohydrate cravers.
Diabetes Care 31, 2169–2176. Psychopharmacology (Berl.) 197, 637–647.
Ross, R.A., 2009. The enigmatic pharmacology of GPR55. Trends Pharmacol. Sci. 30, Starowicz, K.M., Cristino, L., Matias, I., Capasso, R., Racioppi, A., Izzo, A.A., Di Marzo,
156–163. V., 2008. Endocannabinoid dysregulation in the pancreas and adipose tissue of
Rossetti, C., Spena, G., Halfon, O., Boutrel, B., 2013. Evidence for a compulsive-like mice fed with a high-fat diet. Obesity 16, 553–565.
behavior in rats exposed to alternate access to highly preferred palatable food. Steiner, M.A., Wotjak, C.T., 2008. Role of the endocannabinoid system in regula-
Addict. Biol., http://dx.doi.org/10.1111/adb.12065. tion of the hypothalamic-pituitary-adrenocortical axis. Prog. Brain Res. 170,
Rouzer, C.A., Marnett, L.J., 2011. Endocannabinoid oxygenation by cyclooxygenases, 397–432.
lipoxygenases, and cytochromes P450: cross-talk between the eicosanoid and Stice, E., Spoor, S., Bohon, C., Small, D.M., 2008. Relation between obesity and
endocannabinoid signaling pathways. Chem. Rev. 111, 5899–5921. blunted striatal response to food is moderated by TaqIA A1 allele. Science 322,
Salamone, J.D., Correa, M., 2013. Dopamine and food addiction: lexicon badly 449–452.
needed. Biol. Psychiatry 73, e15–e24. Strader, A.D., Woods, S.C., 2005. Gastrointestinal hormones and food intake. Gas-
Salem, N., Pawlosky, R., Wegher, B., Hibbeln, J., 1999. In vivoconversion of linoleic troenterology 128, 175–191.
acid to arachidonic acid in human adults. Prostaglandins Leukot. Essent. Fatty Sugiura, T., Kondo, S., Sukagawa, A., Nakane, S., Shinoda, A., Itoh, K., Yamashita, A.,
Acids 60, 407–410. Waku, K., 1995. 2-Arachidonoylglycerol: a possible endogenous cannabinoid
Sanchis–Segura, C., Cline, B.H., Marsicano, G., Lutz, B., Spanagel, R., 2004. Reduced receptor ligand in brain. Biochem. Biophys. Res. Commun. 215, 89–97.
sensitivity to reward in CB1 knockout mice. Psychopharmacology (Berl.) 176, Suri, A., Szallasi, A., 2008. The emerging role of TRPV1 in diabetes and obesity. Trends
223–232. Pharmacol. Sci. 29, 29–36.
Sawzdargo, M., Nguyen, T., Lee, D.K., Lynch, K.R., Cheng, R., Heng, H.H., George, S.R., Swanson, S.A., Crow, S.J., Le Grange, D., Swendsen, J., Merikangas, K.R., 2011. Preva-
O’Dowd, B.F., 1999. Identification and cloning of three novel human G protein- lence and correlates of eating disorders in adolescents: results from the National
coupled receptor genes GPR52, PsiGPR53 and GPR55: GPR55 is extensively Comorbidity Survey Replication Adolescent Supplement. Arch. Gen. Psychiatry
expressed in human brain. Brain Res. Mol. Brain Res. 64, 193–198. 68, 714–723.
Scavone, J.L., Sterling, R.C., Van Bockstaele, E.J., 2013. Cannabinoid and opioid inter- Szabo, B., Siemes, S., Wallmichrath, I., 2002. Inhibition of GABAergic neurotrans-
actions: implications for opiate dependence and withdrawal. Neuroscience 248, mission in the ventral tegmental area by cannabinoids. Eur. J. Neurosci. 15,
637–654. 2057–2061.
Scheen, A.J., Finer, N., Hollander, P., Jensen, M.D., Van Gaal, L.F., Group, RI-DS, 2006. Tallett, A.J., Blundell, J.E., Rodgers, R.J., 2009. Effects of acute low-dose combined
Efficacy and tolerability of rimonabant in overweight or obese patients with type treatment with naloxone and AM 251 on food intake, feeding behaviour and
2 diabetes: a randomised controlled study. Lancet 368, 1660–1672. weight gain in rats. Pharmacol. Biochem. Behav. 91, 358–366.
Schroeder, M., Eberlein, C., de Zwaan, M., Kornhuber, J., Bleich, S., Frieling, H., 2012. Tam, J., Cinar, R., Liu, J., Godlewski, G., Wesley, D., Jourdan, T., Szanda, G., Mukhopad-
Lower levels of cannabinoid 1 receptor mRNA in female eating disorder patients: hyay, B., Chedester, L., Liow, J.S., Innis, R.B., Cheng, K., Rice, K.C., Deschamps, J.R.,
association with wrist cutting as impulsive self-injurious behavior. Psychoneu- Chorvat, R.J., McElroy, J.F., Kunos, G., 2012. Peripheral cannabinoid-1 receptor
roendocrinology 37, 2032–2036. inverse agonism reduces obesity by reversing leptin resistance. Cell Metab. 16,
Serrano, A., Del Arco, I., Javier Pavón, F., Macías, M., Perez-Valero, V., 167–179.
Rodríguez de Fonseca, F., 2008. The cannabinoid CB1 receptor antagonist Tam, J., Vemuri, V.K., Liu, J., Bátkai, S., Mukhopadhyay, B., Godlewski, G., Osei-
SR141716A (Rimonabant) enhances the metabolic benefits of long-term Hyiaman, D., Ohnuma, S., Ambudkar, S.V., Pickel, J., Makriyannis, A., Kunos, G.,
treatment with oleoylethanolamide in Zucker rats. Neuropharmacology 54, 2010. Peripheral CB1 cannabinoid receptor blockade improves cardiometabolic
226–234. risk in mouse models of obesity. J. Clin. Invest. 120, 2953–2966.
Shoaib, M., 2008. The cannabinoid antagonist AM251 attenuates nicotine self- Tanda, G., Goldberg, S.R., 2003. Cannabinoids: reward, dependence, and underlying
administration and nicotine-seeking behaviour in rats. Neuropharmacology 54, neurochemical mechanisms – a review of recent preclinical data. Psychophar-
438–444. macology (Berl.) 169, 115–134.
Siegfried, Z., Kanyas, K., Latzer, Y., Karni, O., Bloch, M., Lerer, B., Berry, E.M., 2004. Tanda, G., Pontieri, F.E., Di Chiara, G., 1997. Cannabinoid and heroin activation of
Association study of cannabinoid receptor gene (CNR1) alleles and anorexia ner- mesolimbic dopamine transmission by a common ␮1 opioid receptor mecha-
vosa: differences between restricting and binging/purging subtypes. Am. J. Med. nism. Science 276, 2048–2050.
Genet. B: Neuropsychiatr. Genet. 125B, 126–130. Teegarden, S.L., Bale, T.L., 2007. Decreases in dietary preference produce increased
Silvestri, C., Di Marzo, V., 2012. Second generation CB1 receptor blockers and emotionality and risk for dietary relapse. Biol. Psychiatry 61, 1021–1029.
other inhibitors of peripheral endocannabinoid overactivity and the ratio- Tellez, L.A., Medina, S., Han, W., Ferreira, J.G., Licona-Limón, P., Ren, X., Lam, T.T.,
nale of their use against metabolic disorders. Expert Opin. Investig. Drugs 21, Schwartz, G.J., de Araujo, I.E., 2013. A gut lipid messenger links excess dietary
1309–1322. fat to dopamine deficiency. Science 341, 800–802.
Sink, K.S., McLaughlin, P.J., Wood, J.A., Brown, C., Fan, P., Vemuri, V.K., Peng, Y., Terry, G.E., Liow, J.S., Zoghbi, S.S., Hirvonen, J., Farris, A.G., Lerner, A., Tauscher, J.T.,
Olszewska, T., Thakur, G.A., Makriyannis, A., Parker, L.A., Salamone, J.D., 2008. Schaus, J.M., Phebus, L., Felder, C.C., Morse, C.L., Hong, J.S., Pike, V.W., Halldin, C.,
The novel cannabinoid CB1 receptor neutral antagonist AM4113 suppresses food Innis, R.B., 2009. Quantitation of cannabinoid CB1 receptors in healthy human
intake and foodreinforced behavior but does not induce signs of nausea in rats. brain using positron emission tomography and an inverse agonist radioligand.
Neuropsychopharmacology 33, 946–955. Neuroimage 48, 362–370.
Sipe, J.C., Waalen, J., Gerber, A., Beutler, E., 2005. Overweight and obesity associated Thabuis, C., Destaillats, F., Landrier, J.F., Tissot-Favre, D., Martin, J.C., 2010. Analysis of
with a missense polymorphism in fatty acid amide hydrolase (FAAH). Int. J. Obes. gene expression pattern reveals potential targets of dietary oleoylethanolamide
29, 755–759. in reducing body fat gain in C3H mice. J. Nutr. Biochem. 21, 922–928.
Skidmore, P.M., Yarnell, J.W., 2004. The obesity epidemic: prospects for prevention. Thamotharan, S., Lange, K., Zale, E.L., Huffhines, L., Fields, S., 2013. The role of impulsi-
QJM 97, 817–825. vity in pediatric obesity and weight status. A meta-analytic review. Clin. Psychol.
Snider, N.T., Walker, V.J., Hollenberg, P.F., 2010. Oxidation of the endogenous Rev. 33, 253–262.
cannabinoid arachidonoyl ethanolamide by the cytochrome P450 monooxy- Thanos, P.K., Ramalhete, R.C., Michaelides, M., Piyis, Y.K., Wang, G.J., Volkow, N.D.,
genases: physiological and pharmacological implications. Pharmacol. Rev. 62, 2008. Leptin receptor deficiency is associated with upregulation of cannabinoid
136–154. 1 receptors in limbic brain regions. Leptin receptor deficiency is associated with
224 C. D’Addario et al. / Neuroscience and Biobehavioral Reviews 47 (2014) 203–224

upregulation of cannabinoid 1 receptors in limbic brain regions. Synapse 62, Volkow, N.D., Fowler, J.S., Wang, G.J., Baler, R., Telang, F., 2009a. Imaging dopamine’s
637–642. role in drug abuse and addiction. Neuropharmacology 56, 3–8.
Timofeeva, E., Baraboi, E.D., Poulin, A.M., Richard, D., 2009. Palatable high-energy Volkow, N.D., O’Brien, C.P., 2007. Issues for DSM-V: should obesity be included as a
diet decreases the expression of cannabinoid type 1 receptor messenger RNA in brain disorder? Am. J. Psychiatry 164, 708–710.
specific brain regions in the rat. J. Neuroendocrinol. 21, 982–992. Volkow, N.D., Wang, G.J., Telang, F., Fowler, J.S., Logan, J., Childress, A.R., Jayne, M., Ma,
Tomasi, D., Volkow, N.D., 2013. Striatocortical pathway dysfunction in addiction and Y., Wong, C., 2006. Cocaine cues and dopamine in dorsal striatum: mechanism
obesity: differences and similarities. Crit. Rev. Biochem. Mol. Biol. 48, 1–19. of craving in cocaine addiction. J. Neurosci. 26, 6583–6588.
Tsuboi, K., Sun, Y.X., Okamoto, Y., Araki, N., Tonai, T., Ueda, N., 2005. Molecular char- Volkow, N.D., Wise, R.A., 2005. How can drug addiction help us understand obesity?
acterization of N-acylethanolamine-hydrolyzing acid amidase, a novel member Nat. Neurosci. 8, 555–560.
of the choloylglycine hydrolase family with structural and functional similarity Wakley, A.A., Rasmussen, E.B., 2009. Effects of cannabinoid drugs on the reinforcing
to acid ceramidase. J. Biol. Chem. 280, 11082–11092. properties of food in gestationally undernourished rats. Pharmacol. Biochem.
Tucci, S.A., Rogers, E.K., Korbonits, M., Kirkham, T.C., 2004. The cannabinoid CB1 Behav. 94, 30–36.
receptor antagonist SR141716 blocks the orexigenic effects of intrahypothala- Wang, G.J., Geliebter, A., Volkow, N.D., Telang, F.W., Logan, J., Jayne, M.C., Galanti,
mic ghrelin. Br. J. Pharmacol. 143, 520–523. K., Selig, P.A., Han, H., Zhu, W., Wong, C.T., Fowler, J.S., 2011. Enhanced striatal
Turnbaugh, P.J., Ley, R.E., Mahowald, M.A., Magrini, V., Mardis, E.R., Gordon, J.I., dopamine release during food stimulation in binge eating disorder. Obesity 19,
2006. An obesity associated gut microbiome with increased capacity for energy 1601–1608.
harvest. Nature 444, 1027–1031. Ward, S.J., Dykstra, L.A., 2005. The role of CB1 receptors in sweet versus fat reinforce-
Van der Stelt, M., van Kuik, J.A., Bari, M., van Zadelhoff, G., Leeflang, B.R., Veldink, G.A., ment: effect of CB1 receptor deletion, CB1 receptor antagonism (SR141716A)
Finazzi-Agrò, A., Vliegenthart, J.F., Maccarrone, M., 2002. Oxygenated metabo- and CB1 receptor agonism (CP-55940). Behav. Pharmacol. 16, 381–388.
lites of anandamide and 2-arachidonoylglycerol: conformational analysis and Watanabe, S., Doshi, M., Hamazaki, T., 2003. n-3 Polyunsaturated fatty acid
interaction with cannabinoid receptors, membrane transporter, and fatty acid (PUFA) deficiency elevates and n-3 PUFA enrichment reduces brain 2-
amide hydrolase. J. Med. Chem. 45, 3709–3720. arachidonoylglycerol level in mice. Prostaglandins Leukot. Essent. Fatty Acids
Van Gaal, L.F., Rissanen, A.M., Scheen, A.J., Ziegler, O., Rossner, S., for the RIO-Europe 69, 51–59.
Study Group, 2005. Effects of the cannabinoid-1 receptor blocker rimonabant on Wei, B.Q., Mikkelsen, T.S., McKinney, M.K., Lander, E.S., Cravatt, B.F., 2006. A second
weight reduction and cardiovascular risk factors in overweight patients: 1-year fatty acid amide hydrolase with variable distribution among placental mam-
experience from the RIO-Europe study. Lancet 365, 1389–1397. mals. J. Biol. Chem. 281, 36569–36578.
Van Hell, H.H., Jager, G., Bossong, M.G., Brouwer, A., Jansma, J.M., Zuurman, L., Van Wiedmer, P., Nogueiras, R., Broglio, F., D’Alessio, D., Tschöp, M.H., 2007. Ghrelin,
Gerven, J., Kahn, R.S., Ramsey, N.F., 2012. Involvement of the endocannabinoid obesity and diabetes. Nat. Clin. Pract. Endocrinol. Metab. 3, 705–712.
system in reward processing in the human brain. Psychopharmacology (Berl.) Williams, C.M., Kirkham, T.C., 1999. Anandamide induces overeating: mediation by
219, 981–990. central cannabinoid (CB1) receptors. Psychopharmacology (Berl.) 143, 315–317.
Van Sickle, M.D., Duncan, M., Kingsley, P.J., Mouihate, A., Urbani, P., Mackie, K., Williams, C.M., Rogers, P.J., Kirkham, T.C., 1998. Hyperphagia in pre-fed rats follow-
Stella, N., Makriyannis, A., Piomelli, D., Davison, J.S., Marnett, L.J., Di Marzo, ing oral delta 9-THC. Physiol. Behav. 65, 343–346.
V., Pittman, Q.J., Patel, K.D., Sharkey, K.A., 2005. Identification and func- Wise, R.A., 2004. Dopamine, learning and motivation. Nat. Rev. Neurosci. 5, 483–494.
tional characterization of brainstem cannabinoid CB2 receptors. Science 310, Wonderlich, S.A., Gordon, K.H., Mitchell, J.E., Crosby, R.D., Engel, S.G., 2009. Int. J. Eat.
329–332. Disord. 42, 687–705.
Verty, A.N., Evetts, M.J., Crouch, G.J., McGregor, I.S., Stefanidis, A., Oldfield, B.J., 2011. Wood, J.T., Williams, J.S., Pandarinathan, L., Janero, D.R., Lammi-Keefe, C.J., Makriyan-
The cannabinoid receptor agonist THC attenuates weight loss in a rodent model nis, A., 2010. Dietarydocosahexaenoic acid supplementation alters select
of activity-based anorexia. Neuropsychopharmacology 36, 1349–1358. physiological endocannabinoid-system metabolites in brain and plasma. J. Lipid
Verty, A.N., McFarlane, J.R., McGregor, I.S., Mallet, P.E., 2004. Evidence for an interac- Res. 51, 1416–1423.
tion between CB1 cannabinoid and melanocortin MCR-4 receptors in regulating Yao, L., Fan, P., Jiang, Z., Mailliard, W.S., Gordon, A.S., Diamond, I., 2003. Addicting
food intake. Endocrinology 145, 3224–3231. drugs utilize a synergistic molecular mechanism in common requiring adeno-
Viscomi, M.T., Oddi, S., Latini, L., Pasquariello, N., Florenzano, F., Bernardi, G., Moli- sine and Gi-␤␥ dimers. Proc. Natl. Acad. Sci. U.S.A. 100, 14379–14384.
nari, M., Maccarrone, M., 2009. Selective CB2 receptor agonism protects central Yoshida, R., Ohkuri, T., Jyotaki, M., Yasuo, T., Horio, N., Yasumatsu, K., Sanematsu,
neurons from remote axotomy-induced apoptosis through the PI3K/Akt path- K., Shigemura, N., Yamamoto, T., Margolskee, R.F., Ninomiya, Y., 2010. Endo-
way. J. Neurosci. 29, 4564–4570. cannabinoids selectively enhance sweet taste. Proc. Natl. Acad. Sci. U.S.A. 107,
Volkow, N.D., Gillespie, H., Mullani, N., Tancredi, L., Grant, C., Valentine, A., Hollister, 935–939.
L., 1996. Brain glucose metabolism in chronic marijuana users at baseline and Zbucki, R.L., Sawicki, B., Hryniewicz, A., Winnicka, M.M., 2008. Cannabinoids enhance
during marijuana intoxication. Psychiatry Res. 67, 29–38. gastric X/A-like cells activity. Folia Histochem. Cytobiol. 46, 219–224.
Volkow, N.D., Wang, G.J., Telang, F., Fowler, J.S., Thanos, P.K., Logan, J., Alexoff, D., Zhang, Y., Von Deneen, K.M., Tian, J., Gold, M.S., Liu, Y., 2011. Food addiction and
Ding, Y.S., Wong, C., Ma, Y., Pradhan, K., 2008. Low dopamine striatal D2 receptors neuroimaging. Curr. Pharm. Des. 17, 1149–1157.
are associated with prefrontal metabolism in obese subjects: possible contribut- Ziauddeen, H., Farooqi, I.S., Fletcher, P.C., 2012. Obesity and the brain: how convinc-
ing factors. Neuroimage 42, 1537–1543. ing is the addiction model? Nat. Rev. Neurosci. 13, 279–286.

View publication stats

Anda mungkin juga menyukai