Next Article in Journal
Molecular Imaging of Vulnerable Coronary Plaque with Radiolabeled Somatostatin Receptors (SSTR)
Next Article in Special Issue
Managing the Unpredictable: Mechanistic Analysis and Clinical Recommendations for Lamotrigine Treatment after Bariatric Surgery
Previous Article in Journal
Dysautonomia and Implications for Anosmia in Long COVID-19 Disease
Previous Article in Special Issue
Changes in Trimethylamine-N-oxide Levels in Obese Patients following Laparoscopic Roux-en-Y Gastric Bypass or Sleeve Gastrectomy in a Korean Obesity Surgical Treatment Study (KOBESS)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Impact of Bariatric Surgery on Adipose Tissue Biology

1
Centro de Investigación Biomédica en Red de Diabetes y Enfermedades Metabólicas Asociadas (CIBERDEM), Instituto de Salud Carlos III (ISCIII), 28029 Madrid, Spain
2
Translational Research in Diabetes, Lipids and Obesity, Institut d’Investigacions Biomèdiques August Pi I Sunyer (IDIBAPS), 08036 Barcelona, Spain
3
Obesity Unit, Endocrinology and Nutrition Department, Hospital Clínic de Barcelona, 08036 Barcelona, Spain
4
Centro de Investigación Biomédica en Red Fisiopatologia de la Obesidad y Nutrición (CIBEROBN), Instituto de Salud Carlos III (ISCIII), 28029 Madrid, Spain
*
Authors to whom correspondence should be addressed.
J. Clin. Med. 2021, 10(23), 5516; https://doi.org/10.3390/jcm10235516
Submission received: 27 September 2021 / Revised: 20 November 2021 / Accepted: 22 November 2021 / Published: 25 November 2021
(This article belongs to the Special Issue Bariatric Surgery: Latest Advances and Prospects)

Abstract

:
Bariatric surgery (BS) procedures are actually the most effective intervention to help subjects with severe obesity achieve significant and sustained weight loss. White adipose tissue (WAT) is increasingly recognized as the largest endocrine organ. Unhealthy WAT expansion through adipocyte hypertrophy has pleiotropic effects on adipocyte function and promotes obesity-associated metabolic complications. WAT dysfunction in obesity encompasses an altered adipokine secretome, unresolved inflammation, dysregulated autophagy, inappropriate extracellular matrix remodeling and insufficient angiogenic potential. In the last 10 years, accumulating evidence suggests that BS can improve the WAT function beyond reducing the fat depot sizes. The causal relationships between improved WAT function and the health benefits of BS merits further investigation. This review summarizes the current knowledge on the short-, medium- and long-term outcomes of BS on the WAT composition and function.

1. Introduction

White adipose tissue (WAT) has evolved to become the largest endocrine organ. Its plasticity in response to excess or deficit of nutrients is crucial to maintain metabolic health. The remodeling and expansion capacity of adipose tissue implies the orchestrated response of adipocytes, immune cells, endothelial cells, fibroblasts, the extracellular matrix, and its secretome (cytokines, hormones, microRNAs) as mediators of crosstalk between the main organs involved in metabolic health. Dysfunctional expansion of adipose tissue emerges as a key determinant of obesity-related complications. WAT expansion beyond the subcutaneous adipose tissue (SAT) capacity leads to visceral adipose tissue (VAT) expansion and ectopic fat deposition in other tissues, which are major contributors to cardiovascular disease and metabolic risk above body mass index (BMI) [1]. The precise mechanism leading to impaired adipose tissue expandability are not fully understood. Bariatric surgery (BS) currently results in weight loss and better control of comorbid obesity conditions than medical therapy. BS is also associated with a reduced risk of mortality and of some types of cancer [2]. Currently, Roux-en-Y gastric bypass (RYGBP), sleeve gastrectomy (SG), and biliopancreatic diversion (BPD) are the main surgical techniques used worldwide [3].
This review aims to delve into the biology of adipose tissue in the context of obesity and its changes after BS.

2. Obesity-Related White Adipose Tissue Dysfunction

To identify what hypothetical benefits BS has on adipose tissue biology, we need to cite first the most consensed features of obesity-related WAT dysfunction: an altered adipokine secretome [4,5], unresolved inflammation [6,7], inappropriate extracellular matrix (ECM) remodeling, and insufficient angiogenic potential [8]. The causal order in this context is not completely known; however, hypertrophic adipocytes seem more prone to this scenario as they reach the diffusional limit of oxygen, resulting in persistent hypoxia and ultimately leading to unhealthy WAT tissue expansion [8]. Given its role in WAT remodeling, some authors add autophagy dysregulation to this context [9,10]. Among these features, inflammation-related phenomena, i.e., impaired adipokine and cytokine secretion, have been undoubtedly the most exhaustively studied and tracked parameters during the postsurgical follow-up period after BS.
Obese WAT is characterized by macrophage infiltration, a condition considered as both the cause and consequence of its immune response, which leads to chronic inflammation [11,12]. Obesity-related accumulation of adipose tissue macrophages (ATMs) has been clearly demonstrated in multiple studies [13,14,15] and the majority of such ATMs accumulate in omental rather than subcutaneous depots [14,15,16]. Thus, while a small number of macrophages, preferentially localized near blood vessels and dispersed among mature adipocytes are found in lean WAT, subjects with severe obesity show a higher abundance of infiltrating macrophages forming crown-like structures (CLS) around single adipocytes [13]. Such macrophages predominantly present the M1 pro-inflammatory phenotype and promote inflammation by releasing tumor necrosis factor alpha (TNF-α) and interleukin 6 (IL-6), thus contributing to insulin resistance. Alternatively activated, M2-like macrophages play a role in WAT expansion, thermoregulation, antigen presentation, and iron homeostasis [17]. In lean humans, the number of M2 ATMs predominates, secreting anti-inflammatory cytokines and utilizing oxidative metabolism to maintain WAT homeostasis. During obesity development, their proportion compared to M1 ATMs decreases and both populations may adopt a glycolytic metabolism [18]. Once the WAT healthy growth capacity is exceeded, the production of specific adipokines and cytokines by adipocytes and ATMs is compromised and can affect other organ systems.
Although the secretome of many other pro- [19,20,21] and anti-inflammatory [19,22,23] mediators have been found altered in the context of obesity, those described here are the most comprehensively evaluated, and their postsurgical modulation at different follow-up times is summarized below.

2.1. Adipose Tissue-Derived Cytokines

Although the systemic impact of WAT cytokine production in the context of obesity and diabetes has been recently called into question [24], it has been consistently shown that BS-induced weight loss progressively decreases the infiltration of macrophages and WAT secretion of pro-inflammatory molecules [25]. Such cytokines can be both released by WAT-resident immune cells or directly from adipocytes. Below, we describe the most well-studied in the context of obesity.

2.1.1. Pro-Inflammatory

TNF-α is a 17 kDa pro-inflammatory cytokine that can be secreted from mature adipocytes but is predominantly produced within WAT stromovascular fraction, including preadipocytes, endothelial cells, smooth muscle cells, fibroblasts, leukocytes, and macrophages [26,27,28]. The latter is thought to be the major responsible for the elevated expression during obesity [29].
IL-1β is a major 17.5 kDa pro-inflammatory cytokine secreted mostly by macrophages [30], and its release from WAT nonfat cells is augmented during obesity [31]. With other inflammatory mediators, its production is greater in the visceral than in the subcutaneous depot [32]. Increased circulating IL-1β levels have been associated with the risk of developing type 2 diabetes [33], inasmuch as IL-1β contributes to inhibiting β-cell function and destroying β-cell mass [34,35] and impairs adipocyte insulin signaling [36].
Adipocytes, fibroblasts, and endothelial and immune cells secrete IL-6, which induces fever and liver production of the acute phase reactants, and also mediates chronic inflammatory responses [37]. Both adipocytes and macrophages are responsible for its overexpression in WAT during obesity [38,39]. Visceral rather than subcutaneous depot seems to be the main source of circulating IL-6 levels [40].
Different cell types express IL-8, such as monocytes, macrophages, fibroblasts, endothelial cells, and adipocytes [41,42]. IL-8 acts as a chemokine, attracting leukocytes [39]; as a pro-angiogenic factor [43]; and an amplifier of inflammation [44]. Secretion of IL-8 from WAT is increased [45], mainly in the visceral depot [37] during obesity, and is associated with insulin resistance [42].
IL-18 is another pro-inflammatory cytokine, produced by both hematopoietic cells and non-hematopoietic cells, which has been found to be increased in obesity [46,47] and associated with the metabolic syndrome independently of obesity and insulin resistance [48].
Chemoattractant chemokine ligand 2 (CCL2), also referred to as MCP-1, is a chemoattractant cytokine produced by, among others, myeloid cells [49] and adipocytes [50]. The latter enhances MCP-1 secretion during obesity [51], recruiting and activating macrophages through the MCP-1/IL-1β/CXCL12 signaling pathway [52]. Nevertheless, WAT expansion augmentation does not influence circulating MCP-1 levels [51].
Three isoforms of transforming growth factor-beta (TGF-β) have been identified in mammals, which are produced by all-white blood cells lineages and, to a lesser extent, by mature adipocytes [53]. Despite originally being thought to have overlapping functions, isoform-specific knockout mouse models revealed non-redundant phenotypes [54,55,56], TGF-1β being the predominant and most important isoform [57]. TGF-1β release from WAT is enhanced in obesity and in response to insulin and inhibitors of TNF-α and IL-β and correlates BMI and adiposity [53].

2.1.2. Anti-Inflammatory

Secreted by Th2 T-cells, M2 macrophages, and adipocytes [39,58], IL-10 is an anti-inflammatory cytokine that suppress macrophage activation [59], which has been inversely associated with BMI and body fat content [60].
Primarily secreted from mast cells and eosinophils, cytokines IL-4 and IL-13 are closely related, where the former stimulates the production of the latter [61], both sharing similar anti-inflammatory functions and receptor complexes [62]. The presence of these cytokines promotes alternative activation of macrophages into M2 cells and inhibits M1-like classical activation [63]. Both IL-4 [64] and IL-13 [65] serum concentrations are increased in obesity. Moreover, recent research showed a role for IL-4 in promoting adipocyte thermogenic capacity [66] and lipolysis [67] through hormone-sensitive lipase (HSL) modulation [68].

2.2. Adipose Tissue-Derived Hormones

2.2.1. Pro-Inflammatory

Mostly produced by adipocytes, leptin is a highly conserved 167 kDa peptide. It is secreted proportionally to the amount of adiposity [69,70]. Leptin acts to reduce the food intake at the level of the hypothalamus and fat stores at the level of the adipocyte [71], as well as promoting pro-inflammatory cytokine production by immune cells [38,58].
Resistin, traditionally considered a WAT-specific secretory factor, is a 12.5 kDa hormone which acts as a modulator of body cholesterol trafficking, increasing low-density lipoprotein (LDL)-cholesterol and degrading liver LDL receptors, thus contributing to atherosclerosis pathogenesis. Within WAT, resistin promotes pro-inflammatory cytokine production through the resistin receptor and is found to be increased in obesity [72]. Nevertheless, mounting evidence reveals inconsistencies between resistin’s role in rodents and humans, and its relationship with insulin resistance in humans is still controversial [73], with arguments existing both for [74,75,76] and against [77,78,79] this association.
Visfatin is another proinflammatory adipokine that plays a role in insulin sensitivity and whose production is increased in obesity and correlates with visceral adiposity [80].

2.2.2. Anti-Inflammatory

Adiponectin is secreted from WAT as an oligomer of varying sizes in an inversely proportional manner to the degree of visceral adiposity [81]. Adiponectin plays an anti-inflammatory role and promotes insulin sensitivity by increasing fatty acid oxidation, thus regulating lipoprotein metabolism and inhibiting hepatic glucose production. The adiponectin-leptin ratio is considered a biomarker of inflammation in WAT [82,83].
Predominantly expressed in the visceral depot [84], omentin -34 kDa- is an anti-inflammatory adipokine with insulin-sensitizing effects whose levels are decreased in obesity and diabetes [85] and inversely correlated BMI [86]. The role of other molecules such as apelin, vaspin, and RBP4 in inflammation is less clear.

2.3. Extracellular Matrix Remodeling and Fibrosis

WAT is a highly dynamic organ, as it is responsible for storing and releasing energy in response to nutrient excess or shortage. As WAT expands (by adipocyte enlargement—hypertrophy; and preadipocyte recruitment—hyperplasia), ECM is remodeled to accommodate healthy WAT expansion. Like in other organs, sustained WAT inflammation can trigger aberrant ECM deposition leading to WAT fibrosis. Profibrotic mediators such as TGF-β or connective tissue growth factor (CTGF) participate in this pathway [87]. When WAT becomes fibrotic, ECM stiffness impedes healthy remodeling, causing the tissue to be metabolically dysfunctional, displaying, e.g., adipocyte death, decreased lipolysis, and disrupted cell–cell interactions [87,88]. Thus, inflammation can disarrange the tight balance between ECM composition, extracellular metalloproteinases (MMPs), and their inhibitors (TIMPs) [89]. Data from three independent studies carried out by Karine Clément’s group in BS subjects identified the degree of fibrosis in SAT as a predictor for poorer weight loss response after BS [90,91,92]. In this context, HIF1α has been proposed to link the hypoxic milieu to fibrosis and inflammation [93]. Certainly, accumulating evidence demands further research on the relationship between multiple ECM components and adipocyte function in the context of obesity [88,94,95,96,97] and diabetes [98,99,100], and possible associations with BS outcomes should be explored in depth.

2.4. Basal and Stimulated Lipolysis

Obesity is associated with an increase in basal lipolysis and impaired insulin ability to suppress the FFA outflow [101,102]. Antagonistically, plasma catecholamines are important stimulators of lipolysis via adrenergic receptors, particularly through beta-1 (ADRB1) and beta-3 adrenergic receptors (ADRB3) in human WAT [103], and catecholamine-stimulated lipolysis has also been found to be impaired in obesity [104]. Although the classical notion of ‘catecholamine resistance’ in obesity seems to receive little attention today, some authors recommend its revisitation [105].
More than two decades ago, Kaartinen et al. found a good correlation between the fat cell size and response to isoproterenol in isolated SAT adipocytes from subjects with obesity undergoing BS [106]. Interestingly, after substantial BS-induced weight loss, the lipolytic effect of isoproterenol stimulation of adrenergic receptors was higher than lean controls, despite no difference in receptor density between groups. Similar results have been reported after short-term nutrition interventions [107,108]. Fasting FFA circulating levels are other relevant measures of basal lipolysis, though they are not only dependent on WAT lipolysis but also on clearance by muscle and the liver.

2.5. Angiogenesis

Adipogenesis and angiogenesis are tightly related processes during ‘healthy’ WAT expansion since adipocyte differentiation trigger blood vessel formation [109,110], and in turn, WAT endothelial cells promote preadipocyte differentiation [111]. Vascular endothelial growth factor (VEGF)-A, highly expressed in WAT, plays a capital role in angiogenesis, and its expression is raised during adipogenesis [112,113]. Besides the family of VEGF factors, the angiopoietin (ANGPT) family is also involved in vascular remodeling, maturation, and stabilization [114]. ANGPT-2, expressed in WAT endothelial cells, is considered a proangiogenic factor. Although its overexpression in mice improved the metabolic status [115], its role in the angiogenic process has not yet been elucidated. Platelet endothelial cell adhesion molecule-1 (PECAM-1/CD31) plays a role as an adhesion and signaling molecule with several roles in vascular and inflammatory processes, and its levels are increased in young men with severe obesity [116].

2.6. Autophagy

Autophagy, the cellular mechanism that promotes cell survival during nutrient depletion, may also be relevant under basal or nutrient excess conditions. During nutrient depletion, autophagy can provide essential components for energy production and biosynthesis. In circumstances of nutrient excess, autophagy plays an important role in eliminating unfolded proteins and toxic aggregates and facilitating endoplasmic reticulum homeostasis [117]. In this regard, liver autophagy has been the subject of extensive research [118], while WAT autophagy has been receiving growing attention in recent years and is now considered a key regulator of adipogenesis [9] with intricate implications in ECM remodeling and inflammation [10]. Although some authors have reported attenuated WAT autophagy in obesity [119], not all studies could confirm the sense in which obesity and/or metabolic disruption is related to WAT autophagy alterations [10], and most studies point to overactivation of WAT autophagy in obesity [120,121,122] and diabetes [123,124]. However, several considerations should be taken into account [10]. Since WAT autophagy can be regarded as a protective mechanism to avoid WAT maladaptation to nutritional stress, this may explain enhanced autophagy despite the increased inflammation in dysfunctional WAT. In addition, autophagy has different functions depending upon the cell type; thus, WAT cell heterogeneity should be taken into consideration. Finally, the varied technical approaches used to measure autophagy and the different depots analyzed could explain conflicting results among these studies. All of this together calls for much more research into the relationship between autophagy, obesity, and BS outcomes.

3. Bariatric Surgery—Related Changes in White Adipose Tissue Biology

Since there is no standardization and the definition of short-, mid- and long-term terminologies can vary among published reports [125], from here on, the current knowledge on this topic is summarized across five follow-up time points commonly used to report BS outcomes: ≤3 months (3 m), 6 m, 1 year (1 y)—all often considered to be short-term; ≥2 y <5 y—referred to as medium-term; and >5 y—frequently regarded as long-term post-surgery. All bariatric interventions considered in Table 1 consisted of SG, RYGB, or BPD.

3.1. Short Term

During the first year, coinciding with the rapid weight loss phase after BS, both SAT [126,127,128,129,130] and VAT [126,127,128,129,130,134,135] depots progressively reduce their size, and this is accompanied by a reduction in the area of subcutaneous [136,137,138] and visceral [138] adipocytes, respectively. A large adipocyte size was independently associated with a lower incidence of insulin resistance 6 months after RYGBP [136].
In the very short term after BS (≤3 months), Cancello et al. showed a significant decrease of total ATMs (HAM56+ cells) in SAT after RYGB [13]. These results were confirmed in another study from the same group, wherein CD40+ cells (M1-like) were also found to be decreased and CD206+ and CD163+ cells (M2-like) increased 3 months after RYGB [16]. This was accompanied by a reversion to the lean WAT profile, with CLS remission and ATMs again located near blood vessels [13].
Such early changes in the WAT cellular composition seem to alter the production of some cytokines, while others generate conflicting results between studies or do not seem to be modulated in the short term after BS. Thus, among the proinflammatory cytokines, MCP-1 was found to be concomitantly decreased during this period [13,150,165,166], while TGF-β or IL-1β seem to decrease only at 1 year after BS [142,147,156]. Reports on IL-6 production give conflicting results at 3 and 6 months but agree on a consistent decrease 1 year after surgery [139,142,143,144,146,150,152,156,158,159,160,168]. Similarly, reduced circulating levels of IL-18 were found 1 y post-BS [141] and after massive BS-induced weight loss, irrespective of the time elapsed since surgery [195,196].
In contrast, there is less consensus about TNF-α and IL-8, which have been found in different studies to both be increased [140,155,156,197], decreased [141,142,143,144,163,164], or unchanged [146,147,148,149,150,151,152,157] during this period. Similarly, BS-related outcomes on anti-inflammatory cytokine production have yielded highly contradictory results between studies during the short-term follow-up period, as is the case with IL-4 [64,140,167], IL-10 [141,147,162,164,167,168], and IL-13 [144,167]. Interestingly, circulating omentin levels decrease as early as 24 h post-BS, before any fat mass loss, and maintained for 1 y [172].
Inasmuch as surgical weight loss predominantly reduces the body fat content, it is understandable that leptin levels were found to be consistently reduced following BS [140,141,144,148,149,150,156,158,162,163,168,170]. The leptin levels were also reduced after the novel endovascular bariatric procedure [198]. Nevertheless, systemic leptin levels are not directly related to the amount of body weight or fat loss, since early reductions of adiposity more dramatically reduce leptin levels than later periods of weight loss [162,170]. Again, there is a lack of consensus regarding the short-term effect of BS on resistin levels, given several studies have found it to be decreased [141,142,145,158,168] or unchanged [150,171]. In the case of visfatin, Lima et al. showed unaltered levels throughout the first year after BS [150].
Despite some conflicting reports in the very short term [146,147,149,150,156], circulating adiponectin levels appear to be consistently increased 1 year after BS [139,141,142,150,152,156,158,168]. For its part, omentin was found to be increased as early as 24 h after BPD [172], and such a change is maintained for up to 1 year [172,173]. Apelin, a multifaceted biomarker [174], and vaspin, an insulin-sensitizing adipokine [175], are less investigated adipokines that showed a short-term reduction after BS. Regarding RBP-4, most studies reported a decrease in the circulating [177,178] or SAT mRNA [176] levels early after BS.
One study performed by Chabot and collaborators showed no resolution of SAT fibrosis 6 months after BS and suggested a transient association between SAT fibrosis and insulin resistance in humans with obesity [180]. Similarly, Katsogiannos et al. did not find significant differences in either the basal or stimulated lipolysis rate in SAT adipocytes at 1 and 6 months after BS but reported a decrease in isoproterenol-stimulated lipolysis at 6 versus 1 month after BS [181]. Conversely, insulin-suppressed free fatty acid (FFA) release has been found to be enhanced at 4 months [137], 7 months [183], and 1 year after RYGBP [101]. While some authors found increased FFA levels in the early months after BS [101,148,181], others reported no differences in this period [148,184,185,186].
García de la Torre et al. found higher VEGF-A levels in obese women undergoing BS compared to lean controls, and such levels significantly decreased 1 y after surgery, irrespective of the surgical procedure performed [187]. At this same follow-up period, another recent study showed, in addition to VEGF-A, lower levels of several angiogenesis biomarkers such as angiopoietin 2 (ANGPT-2), follistatin, hepatocyte growth factor (HGF), and the platelet endothelial cell adhesion molecule (PECAM-1) in patients who underwent SG or laparoscopic adjustable gastric banding (LAGB) [188].
Finally, Soussi et al. found attenuated WAT autophagy in obesity, and pre- versus post-BS comparisons indicated ameliorated adipocyte autophagic clearance in all patients within 3 to 12 months after the intervention, although at different degrees because of the large time-frame in post-surgery sample collection [119].

3.2. Medium Term

Two years after surgery, both visceral and subcutaneous depots maintain reduced sizes [126,131,132] as does the abdominal subcutaneous fat cell volume [132]. There is much less data available on circulating parameters beyond 1 y after BS. While IL-6 levels are consistently found reduced 2 y [145,153], 3 y [154], and 4 y after BS [169], reports on TNF-a continue to report conflicting data [145,154]. Although reports on IL-10 also seem quite inconsistent, some authors find that, after a temporary rise in the short term, its levels return to baseline values at 2 y [145], or even continue falling at 4 y [169].
BS outcomes on leptin and adiponectin levels seem much more solid. Circulating leptin has been repeatedly found to be reduced at 2 [145,153], 3 [154], and 4 y [169], and such reductions seem to be mainly attributed to early changes in WAT. Conversely, adiponectin levels continue to progressively rise in the medium term [145,153,154]. Only one report seems to oppose this view, a contradiction that could arise from the limited number of subjects and the variety of surgical techniques included in the study [169].
Beyond the short-term inconsistencies mentioned above, a single study showed that circulating resistin, after an early decline, recovered baseline levels 2 y after gastric bypass [145]. Finally, the RBP-4 levels were found still lowered 24 months after BS. Such changes were more pronounced in the subgroup without metabolic syndrome and correlated with reductions in the waist and visceral fat diameter [179].
Despite negative results reported by Katsogiannos et al. in the short term in a mixed-sex cohort [181], Löfgren and collaborators found reduced basal and stimulated lipolysis rates at 2 y after BS exclusively in females [182], where differences in the basal rates remained only significant when lipolysis was expressed per cell surface area. In another study, the glycerol release in women who underwent RYGBP was found to be decreased postsurgically at 2 y and then increased dramatically to similar levels observed before surgery at 5 y [133]. Similarly, Manco et al. found reduced FFA levels in normoglucose-tolerant obese women 3 years after BPD [154]. Finally, insulin-mediated suppression of FFA outflow has been found to be enhanced 3 years after RYGBP [102].

3.3. Long Term

Studies on long-term outcomes after BS are restricted almost exclusively to weight-loss parameters. Thus, a recent meta-analysis at 10 or more years after all bariatric procedures reported weighted means of 56.7% excess weight loss (EWL) after GB, 45.9%EWL after LAGB, 74.1%EWL after BPD and 58.3%EWL after SG [199]. The same study reported a 48.9%EWL and 22.2%TWL 20 y after LAGB. Very similar results were previously reported by the same group at 15 y after LAGB [200]. A lower incidence [201] and greater remission [202] of T2DM have also been reported in the long term; reductions in all-cause, cardiovascular, and T2DM mortality have also been found [203]. Nevertheless, the potential impact of body fat loss on these metabolic outcomes deserves further investigation since some variables appear to be more weight-dependent, while others seem to be more adiposity-dependent from the medium-term [204].
Regarding the outcomes in WAT exclusively, we only have evidence from a single study carried out in women by Hoffstedt and collaborators at the long-term follow-up [133]. The authors reported decreased amounts of estimated SAT and VAT at 2 and 5 y and diminished SAT cell volume and increased adiponectin levels at 5 y post-BS. This study also found augmented basal glycerol release from isolated SAT adipocytes at 5 y, despite not finding changes in fasting plasma levels.

3.4. Summary of BS Outcomes on WAT

In summary, after bariatric surgery, SAT and VAT reduce their size progressively during the weight-loss phases. M1-like decrease and M2-like ATMs increase early after surgery; however, there are no data beyond the short term after BS.
Most pro-inflammatory cytokines begin to decrease early after surgery and continue to decline in the medium- and long-term. However, TGF-B or IL1B decrease only after one year of BS. There are controversial data on short-term TNFα and IL-8 levels after surgery as well as in anti-inflammatory cytokine levels in the short- and medium-term after surgery. Leptin levels drop rapidly soon after BS and then continue to decline during the follow-up; conversely, adiponectin and omentin levels rise after surgery. Resistin and visfatin dynamics show less agreement.
Regarding fibrosis, only one study reported no changes at short-term. Gender differences seem to affect basal and stimulated rates of lipolysis, which have been found decreased only in females at mid-term after BS. For its part, insulin inhibition of lipolysis was found consistently enhanced at medium- and long-term after surgery. Finally, autophagy increases and several angiogenesis-related molecules decrease at short-term, although there is a lack of reports on longer follow-up periods.

4. Other Proposed Novel Mechanisms for WAT Improvement after BS

Mitochondrial function and biogenesis have been found to be impaired in obesity, and T2DM and BS may attenuate mitochondrial damage in adipocytes. Thus, Varela-Rodríguez and colleagues reported an increased mitochondrial density and coverage, together with enhanced mitochondrial function at both the gene and protein level in abdominal SAT after RYGB- or SG-induced weight loss in a reduced cohort of patients [205]. In the very short term after RYGB, an induction of genes involved in mitochondrial biogenesis was found in SAT [206]. Similarly, increased SAT expression of transcripts related to the oxidative phosphorylation (OXPHOS) pathway has been shown 3 m [207] and 1 y after BS [137,208]. More recently, Van der Kolk et al. confirmed these findings in abdominal SAT in the short and medium term after RYGB, while opposite results were found after a low-calorie diet [209], suggesting a BS-specific effect. The authors also showed an induction of the tricarboxylic acid (TCA) cycle and fatty acid oxidation 2 y after surgery.
Beiging is the process through which WAT can change its phenotype to a brown-like adipose tissue known as beige/brite adipose tissue. Accumulating evidence from human and rodent studies in the last years suggest that RYGB predominantly enhances beige thermogenesis, while SG seems to promote brown adipose tissue thermogenesis [210]. A role for bile acids and the gut microbiome has been proposed in these mechanisms. Moreover, such thermogenic effects could depend on the fat content of the postoperative diet.
Beyond the fat mass loss and biological pathways discussed above, other mechanisms could contribute to the improvement of WAT dysfunction after BS. Thus, Frikke-Schmidt and colleagues recently summarized other potential pathways affecting the WAT function that can play a role after BS [105]. Bile acids, whose levels are persistently found to be increased after surgery [211], may improve adipocyte function acting upon the FXR receptor [212]. Another hypothesized mechanism implies gut microbiome composition. It has been found that the composition of the bacteria in the gut change after BS, and accumulating evidence shows how the bacterial composition can modulate the host immune cell population. Thus, there is the possibility that changes in the gut microbiome initiated by BS can significantly impact the WAT metabolic function by modulating immune-resident cells in WAT [213,214]. Finally, as BS has demonstrated effects on the central regulation of metabolism, potential changes in neural enervation to the WAT may mediate physiological changes after BS, as results in mice suggest [215].
Lastly, recent studies have indicated significant alterations in the expression of several putative adipose tissue-derived microRNAs (miRNAs) after BS [216]. Thus, 1 y after BS, Sangiao-Alvarellos et al. reported raised circulating levels of miR-221 and miR-222 [190], and as early as 21 days post-RYGB, Atkin et al. found modulated levels of seven miRNAs (miR-7-5p, let-7f-5p, miR-15b-5p, let-7i-5p, miR-320c, miR-205-5p, and miR-335-5p) [189], mostly related with diabetes and insulin resistance pathways. Regarding SAT expression, Ortega and collaborators identified 12 modulated miRNAs 2 y after BS, some of them previously found raised in mature adipocytes after inflammatory stimulation (such as miR-146b, miR-376c, and again miR-221) [194], while another study from the same group found significant modifications in 15 mature miRNAs, mostly related to cell cycle, metabolism, and inflammation pathways, in women who underwent RYGB [191].

5. Future Perspectives

Better understanding of the fascinating biology of WAT following BS deserves further investigation. Evaluation of the modifications of WAT biology associated not only with time elapsed after surgery but also with the amount of weight loss is a priority. Studies should help us better understand the relationship between shrinkage in WAT volume and improved WAT function with the health benefits of BS. The health burden associated with a particular BMI in subjects with a weight-reduced state following BS appears to be eased as compared to that in subjects with comparable BMI that have not undergone BS. Thus, future studies should help disentangle how BS helps restore the crosstalk between the different components of the WAT as well as the crosstalk between WAT and other organs.

Author Contributions

Conceptualization, Ó.O.-C., J.V. and A.d.H.; writing—original-draft preparation, Ó.O.-C.; writing—review and editing, Ó.O.-C., J.V. and A.d.H.; supervision, J.V. and A.d.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Neeland, I.J.; Ross, R.; Després, J.-P.; Matsuzawa, Y.; Yamashita, S.; Shai, I.; Seidell, J.; Magni, P.; Santos, R.D.; Arsenault, B.; et al. Visceral and ectopic fat, atherosclerosis, and cardiometabolic disease: A position statement. Lancet Diabetes Endocrinol. 2019, 7, 715–725. [Google Scholar] [CrossRef]
  2. Aminian, A.; Zajichek, A.; Arterburn, D.E.; Wolski, K.E.; Brethauer, S.A.; Schauer, P.R.; Kattan, M.W.; Nissen, S.E. Association of Metabolic Surgery with Major Adverse Cardiovascular Outcomes in Patients with Type 2 Diabetes and Obesity. JAMA 2019, 322, 1271–1282. [Google Scholar] [CrossRef] [PubMed]
  3. Welbourn, R.; Hollyman, M.; Kinsman, R.; Dixon, J.; Liem, R.; Ottosson, J.; Ramos, A.; Våge, V.; Al-Sabah, S.; Brown, W.; et al. Bariatric Surgery Worldwide: Baseline Demographic Description and One-Year Outcomes from the Fourth IFSO Global Registry Report 2018. Obes. Surg. 2019, 29, 782–795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Zorena, K.; Jachimowicz-Duda, O.; Ślęzak, D.; Robakowska, M.; Mrugacz, M. Adipokines and Obesity. Potential Link to Metabolic Disorders and Chronic Complications. Int. J. Mol. Sci. 2020, 21, 3570. [Google Scholar] [CrossRef]
  5. Pardo, M.; Roca-Rivada, A.; Seoane, L.M.; Casanueva, F.F. Obesidomics: Contribution of adipose tissue secretome analysis to obesity research. Endocrine 2012, 41, 374–383. [Google Scholar] [CrossRef]
  6. Ellulu, M.S.; Patimah, I.; Khaza, H.; Rahmat, A.; Abed, Y.; Sci, A.M. Obesity and Inflammation: The Linking Mechanism and the Complications. Arch. Med. Sci. 2016, 13, 851–863. [Google Scholar] [CrossRef]
  7. O’Rourke, R.W. Inflammation in obesity-related diseases. Surgery 2009, 145, 255–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Crewe, C.; An, Y.A.; Scherer, P.E. The ominous triad of adipose tissue dysfunction: Inflammation, fibrosis, and impaired angiogenesis. J. Clin. Investig. 2017, 127, 74–82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Ferhat, M.; Funai, K.; Boudina, S. Autophagy in Adipose Tissue Physiology and Pathophysiology. Antioxidants Redox Signal. 2019, 31, 487–501. [Google Scholar] [CrossRef]
  10. Clemente-Postigo, M.; Tinahones, A.; El Bekay, R.; Malagón, M.M.; Tinahones, F.J. The Role of Autophagy in White Adipose Tissue Function: Implications for Metabolic Health. Metabolites 2020, 10, 179. [Google Scholar] [CrossRef]
  11. McNelis, J.C.; Olefsky, J.M. Macrophages, Immunity, and Metabolic Disease. Immunity 2014, 41, 36–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Wellen, K.E.; Hotamisligil, G.S. Obesity-induced inflammatory changes in adipose tissue. J. Clin. Investig. 2003, 112, 1785–1788. [Google Scholar] [CrossRef] [PubMed]
  13. Cancello, R.; Henegar, C.; Viguerie, N.; Taleb, S.; Poitou, C.; Rouault, C.; Coupaye, M.; Pelloux, V.; Hugol, D.; Bouillot, J.-L.; et al. Reduction of Macrophage Infiltration and Chemoattractant Gene Expression Changes in White Adipose Tissue of Morbidly Obese Subjects after Surgery-Induced Weight Loss. Diabetes 2005, 54, 2277–2286. [Google Scholar] [CrossRef] [Green Version]
  14. Cancello, R.; Tordjman, J.; Poitou, C.; Guilhem, G.; Bouillot, J.L.; Hugol, D.; Coussieu, C.; Basdevant, A.; Hen, A.B.; Bedossa, P.; et al. Increased infiltration of macrophages in omental adipose tissue is associated with marked hepatic lesions in morbid human obesity. Diabetes 2006, 55, 1554–1561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Shapiro, H.; Pecht, T.; Shaco-Levy, R.; Harman-Boehm, I.; Kirshtein, B.; Kuperman, Y.; Chen, A.; Blüher, M.; Shai, I.; Rudich, A. Adipose tissue foam cells are present in human obesity. J. Clin. Endocrinol. Metab. 2013, 98, 1173–1181. [Google Scholar] [CrossRef] [Green Version]
  16. Aron-Wisnewsky, J.; Tordjman, J.; Poitou, C.; Darakhshan, F.; Hugol, D.; Basdevant, A.; Aissat, A.; Guerre-Millo, M.; Clément, K. Human adipose tissue macrophages: M1 and M2 cell surface markers in subcutaneous and omental depots and after weight loss. J. Clin. Endocrinol. Metab. 2009, 94, 4619–4623. [Google Scholar] [CrossRef] [Green Version]
  17. Hill, A.A.; Reid Bolus, W.; Hasty, A.H. A decade of progress in adipose tissue macrophage biology. Immunol. Rev. 2014, 262, 134–152. [Google Scholar] [CrossRef] [Green Version]
  18. Castoldi, A.; De Souza, C.N.; Saraiva Câmara, N.O.; Moraes-Vieira, P.M. The macrophage switch in obesity development. Front. Immunol. 2016, 6, 1. [Google Scholar] [CrossRef] [Green Version]
  19. Mancuso, P. The role of adipokines in chronic inflammation. ImmunoTargets Ther. 2016, 5, 47–56. [Google Scholar] [CrossRef] [Green Version]
  20. Osorio-Conles, Ó.; Guitart, M.; Moreno-Navarrete, J.M.; Escoté, X.; Duran, X.; Fernandez-Real, J.M.; Gómez-Foix, A.M.; Fernández-Veledo, S.; Vendrell, J. Adipose tissue and serum CCDC80 in obesity and its association with related metabolic disease. Mol. Med. 2017, 23, 225–234. [Google Scholar] [CrossRef]
  21. Caër, C.; Rouault, C.; Le Roy, T.; Poitou, C.; Aron-Wisnewsky, J.; Torcivia, A.; Bichet, J.C.; Clément, K.; Guerre-Millo, M.; André, S. Immune cell-derived cytokines contribute to obesity-related inflammation, fibrogenesis and metabolic deregulation in human adipose tissue. Sci. Rep. 2017, 7, 3000. [Google Scholar] [CrossRef]
  22. Ohashi, K.; Shibata, R.; Murohara, T.; Ouchi, N. Role of anti-inflammatory adipokines in obesity-related diseases. Trends Endocrinol. Metab. 2014, 25, 348–355. [Google Scholar] [CrossRef]
  23. Osorio-Conles, O.; Guitart, M.; Chacón, M.R.; Maymo-Masip, E.; Moreno-Navarrete, J.M.; Montori-Grau, M.; Näf, S.; Fernandez-Real, J.M.; Vendrell, J.; Gómez-Foix, A.M. Plasma PTX3 protein levels inversely correlate with insulin secretion and obesity, whereas visceral adipose tissue PTX3 gene expression is increased in obesity. Am. J. Physiol. Endocrinol. Metab. 2011, 301, E1254–E1261. [Google Scholar] [CrossRef] [PubMed]
  24. Rakotoarivelo, V.; Lacraz, G.; Mayhue, M.; Brown, C.; Rottembourg, D.; Fradette, J.; Ilangumaran, S.; Menendez, A.; Langlois, M.-F.; Ramanathan, S. Inflammatory Cytokine Profiles in Visceral and Subcutaneous Adipose Tissues of Obese Patients Undergoing Bariatric Surgery Reveal Lack of Correlation With Obesity or Diabetes. EBioMedicine 2018, 30, 237–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Labrecque, J.; Laforest, S.; Michaud, A.; Biertho, L.; Tchernof, A. Impact of Bariatric Surgery on White Adipose Tissue Inflammation. Can. J. Diabetes 2017, 41, 407–417. [Google Scholar] [CrossRef] [PubMed]
  26. Fain, J.N.; Bahouth, S.W.; Madan, A.K. TNFα release by the nonfat cells of human adipose tissue. Int. J. Obes. 2004, 28, 616–622. [Google Scholar] [CrossRef] [Green Version]
  27. Xu, H.; Barnes, G.T.; Yang, Q.; Tan, G.; Yang, D.; Chou, C.J.; Sole, J.; Nichols, A.; Ross, J.S.; Tartaglia, L.A.; et al. Chronic inflammation in fat plays a crucial role in the development of obesity-related insulin resistance. J. Clin. Investig. 2003, 112, 1821–1830. [Google Scholar] [CrossRef]
  28. De Taeye, B.M.; Novitskaya, T.; McGuinness, O.P.; Gleaves, L.; Medda, M.; Covington, J.W.; Vaughan, D.E. Macrophage TNF-α contributes to insulin resistance and hepatic steatosis in diet-induced obesity. Am. J. Physiol. Endocrinol. Metab. 2007, 293, E713–E725. [Google Scholar] [CrossRef]
  29. Weisberg, S.P.; McCann, D.; Desai, M.; Rosenbaum, M.; Leibel, R.L.; Ferrante, A.W. Obesity is associated with macrophage accumulation in adipose tissue. J. Clin. Investig. 2003, 112, 1796–1808. [Google Scholar] [CrossRef]
  30. Sims, J.E.; Smith, D.E. The IL-1 family: Regulators of immunity. Nat. Rev. Immunol. 2010, 10, 89–102. [Google Scholar] [CrossRef]
  31. Fain, J.N. Release of Interleukins and Other Inflammatory Cytokines by Human Adipose Tissue Is Enhanced in Obesity and Primarily due to the Nonfat Cells. Vitam. Horm. 2006, 74, 443–477. [Google Scholar]
  32. Koenen, T.B.; Stienstra, R.; Van Tits, L.J.; Joosten, L.A.B.; Van Velzen, J.F.; Hijmans, A.; Pol, J.A.; Van Der Vliet, J.A.; Netea, M.G.; Tack, C.J.; et al. The inflammasome and caspase-1 activation: A new mechanism underlying increased inflammatory activity in human visceral adipose tissue. Endocrinology 2011, 152, 3769–3778. [Google Scholar] [CrossRef] [Green Version]
  33. Spranger, J.; Kroke, A.; Möhlig, M.; Hoffmann, K.; Bergmann, M.M.; Ristow, M.; Boeing, H.; Pfeiffer, A.F.H. Inflammatory cytokines and the risk to develop type 2 diabetes: Results of the prospective population-based European Prospective Investigation into Cancer and Nutrition (EPIC)-Potsdam study. Diabetes 2003, 52, 812–817. [Google Scholar] [CrossRef] [Green Version]
  34. Dinarello, C.A.; Donath, M.Y.; Mandrup-Poulsen, T. Role of IL-1β in type 2 diabetes. Curr. Opin. Endocrinol. Diabetes Obes. 2010, 17, 314–321. [Google Scholar] [CrossRef] [PubMed]
  35. Donath, M.Y.; Dalmas, É.; Sauter, N.S.; Böni-Schnetzler, M. Inflammation in obesity and diabetes: Islet dysfunction and therapeutic opportunity. Cell Metab. 2013, 17, 860–872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Gao, D.; Madi, M.; Ding, C.; Fok, M.; Steele, T.; Ford, C.; Hunter, L.; Bing, C. Interleukin-1β mediates macrophage-induced impairment of insulin signaling in human primary adipocytes. Am. J. Physiol. Endocrinol. Metab. 2014, 307, E289–E304. [Google Scholar] [CrossRef] [PubMed]
  37. Juge-Aubry, C.E.; Henrichot, E.; Meier, C.A. Adipose tissue: A regulator of inflammation. Best Pract. Res. Clin. Endocrinol. Metab. 2005, 19, 547–566. [Google Scholar] [CrossRef]
  38. Rasouli, N.; Kern, P.A. Adipocytokines and the metabolic complications of obesity. J. Clin. Endocrinol. Metab. 2008, 93, s64–s73. [Google Scholar] [CrossRef] [Green Version]
  39. Balistreri, C.R.; Caruso, C.; Candore, G. The role of adipose tissue and adipokines in obesity-related inflammatory diseases. Mediators Inflamm. 2020, 2010, 802078. [Google Scholar] [CrossRef]
  40. Fontana, L.; Eagon, J.C.; Trujillo, M.E.; Scherer, P.E.; Klein, S. Visceral Fat Adipokine Secretion Is Associated With Systemic Inflammation in Obese Humans. Diabetes 2007, 56, 1010–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Bruun, J.M.; Pedersen, S.B.; Richelsen, B. Regulation of Interleukin 8 Production and Gene Expression in Human Adipose Tissue in Vitro1. J. Clin. Endocrinol. Metab. 2001, 86, 1267–1273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Kobashi, C.; Asamizu, S.; Ishiki, M.; Iwata, M.; Usui, I.; Yamazaki, K.; Tobe, K.; Kobayashi, M.; Urakaze, M. Inhibitory effect of IL-8 on insulin action in human adipocytes via MAP kinase pathway. J. Inflamm. 2009, 6, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Koch, A.E.; Polverini, P.J.; Kunkel, S.L.; Harlow, L.A.; DiPietro, L.A.; Elner, V.M.; Elner, S.G.; Strieter, R.M. Interleukin-8 as a macrophage-derived mediator of angiogenesis. Science 1992, 258, 1798–1801. [Google Scholar] [CrossRef] [PubMed]
  44. Baggiolini, M.; Clark-Lewis, I. Interleukin-8, a chemotactic and inflammatory cytokine. FEBS Lett. 1992, 307, 97–101. [Google Scholar] [CrossRef] [Green Version]
  45. Kim, C.S.; Park, H.S.; Kawada, T.; Kim, J.H.; Lim, D.; Hubbard, N.E.; Kwon, B.S.; Erickson, K.L.; Yu, R. Circulating levels of MCP-1 and IL-8 are elevated in human obese subjects and associated with obesity-related parameters. Int. J. Obes. 2006, 30, 1347–1355. [Google Scholar] [CrossRef] [Green Version]
  46. Esposito, K.; Pontillo, A.; Di Palo, C.; Giugliano, G.; Masella, M.; Marfella, R.; Giugliano, D. Effect of Weight Loss and Lifestyle Changes on Vascular Inflammatory Markers in Obese Women: A Randomized Trial. J. Am. Med. Assoc. 2003, 289, 1799–1804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Bruun, J.M.; Stallknecht, B.; Helge, J.W.; Richelsen, B. Interleukin-18 in plasma and adipose tissue: Effects of obesity, insulin resistance, and weight loss. Eur. J. Endocrinol. 2007, 157, 465–471. [Google Scholar] [CrossRef] [Green Version]
  48. Hung, J.; McQuillan, B.M.; Chapman, C.M.L.; Thompson, P.L.; Beilby, J.P. Elevated interleukin-18 levels are associated with the metabolic syndrome independent of obesity and insulin resistance. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 1268–1273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Gschwandtner, M.; Derler, R.; Midwood, K.S. More Than Just Attractive: How CCL2 Influences Myeloid Cell Behavior Beyond Chemotaxis. Front. Immunol. 2019, 10, 2759. [Google Scholar] [CrossRef] [Green Version]
  50. Fasshauer, M.; Blüher, M. Adipokines in health and disease. Trends Pharmacol. Sci. 2015, 36, 461–470. [Google Scholar] [CrossRef]
  51. Dahlman, I.; Kaaman, M.; Olsson, T.; Tan, G.D.; Bickerton, A.S.T.; Wåhlén, K.; Andersson, J.; Nordström, E.A.; Blomqvist, L.; Sjögren, A.; et al. A unique role of monocyte chemoattractant protein 1 among chemokines in adipose tissue of obese subjects. J. Clin. Endocrinol. Metab. 2005, 90, 5834–5840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Arendt, L.M.; McCready, J.; Keller, P.J.; Baker, D.D.; Naber, S.P.; Seewaldt, V.; Kuperwasser, C. Obesity promotes breast cancer by CCL2-mediated macrophage recruitment and angiogenesis. Cancer Res. 2013, 73, 6080–6093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Fain, J.N.; Tichansky, D.S.; Madan, A.K. Transforming Growth Factor β1 release by human adipose tissue is enhanced in obesity. Metabolism 2005, 54, 1546–1551. [Google Scholar] [CrossRef]
  54. Kulkarni, A.B.; Huh, C.G.; Becker, D.; Geiser, A.; Lyght, M.; Flanders, K.C.; Roberts, A.B.; Sporn, M.B.; Ward, J.M.; Karlsson, S. Transforming growth factor β1 null mutation in mice causes excessive inflammatory response and early death. Proc. Natl. Acad. Sci. USA 1993, 90, 770–774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Proetzel, G.; Pawlowski, S.A.; Wiles, M.V.; Yin, M.; Boivin, G.P.; Howles, P.N.; Ding, J.; Ferguson, M.W.J.; Doetschman, T. Transforming growth factor–β3 is required for secondary palate fusion. Nat. Genet. 1995, 11, 409–414. [Google Scholar] [CrossRef]
  56. Sanford, L.P.; Ormsby, I.; Gittenberger-de Groot, A.C.; Sariola, H.; Friedman, R.; Boivin, G.P.; Lou Cardell, E.; Doetschman, T. TGFβ2 knockout mice have multiple developmental defects that are non-overlapping with other TGFβ knockout phenotypes. Development 1997, 124, 2659–2670. [Google Scholar] [CrossRef]
  57. Li, M.O.; Wan, Y.Y.; Sanjabi, S.; Robertson, A.K.L.; Flavell, R.A. Transforming growth factor-β regulation of immune responses. Annu. Rev. Immunol. 2006, 24, 99–146. [Google Scholar] [CrossRef]
  58. Ouchi, N.; Parker, J.L.; Lugus, J.J.; Walsh, K. Adipokines in inflammation and metabolic disease. Nat. Rev. Immunol. 2011, 11, 85–97. [Google Scholar] [CrossRef]
  59. O’Farrell, A.M.; Liu, Y.; Moore, K.W.; Mui, A.L.F. IL-10 inhibits macrophage activation and proliferation by distinct signaling mechanisms: Evidence for Stat3-dependent and -independent pathways. EMBO J. 1998, 17, 1006–1018. [Google Scholar] [CrossRef]
  60. Charles, B.A.; Doumatey, A.; Huang, H.; Zhou, J.; Chen, G.; Shriner, D.; Adeyemo, A.; Rotimi, C.N. The roles of IL-6, IL-10, and IL-1RA in obesity and insulin resistance in African-Americans. J. Clin. Endocrinol. Metab. 2011, 96, E2018–E2022. [Google Scholar] [CrossRef]
  61. McLeod, J.J.A.; Baker, B.; Ryan, J.J. Mast cell production and response to IL-4 and IL-13. Cytokine 2015, 75, 57–61. [Google Scholar] [CrossRef] [Green Version]
  62. Van Dyken, S.J.; Locksley, R.M. Interleukin-4-and interleukin-13-mediated alternatively activated macrophages: Roles in homeostasis and disease. Annu. Rev. Immunol. 2013, 31, 317–343. [Google Scholar] [CrossRef] [Green Version]
  63. Benoit, M.; Desnues, B.; Mege, J.-L. Macrophage Polarization in Bacterial Infections. J. Immunol. 2008, 181, 3733–3739. [Google Scholar] [CrossRef] [Green Version]
  64. Dandona, P.; Ghanim, H.; Monte, S.V.; Caruana, J.A.; Green, K.; Abuaysheh, S.; Lohano, T.; Schentag, J.; Dhindsa, S.; Chaudhuri, A. Increase in the mediators of asthma in obesity and obesity with type 2 diabetes: Reduction with weight loss. Obesity 2014, 22, 356–362. [Google Scholar] [CrossRef]
  65. Schmidt, F.M.; Weschenfelder, J.; Sander, C.; Minkwitz, J.; Thormann, J.; Chittka, T.; Mergl, R.; Kirkby, K.C.; Faßhauer, M.; Stumvoll, M.; et al. Inflammatory cytokines in general and central obesity and modulating effects of physical Activity. PLoS ONE 2015, 10, e0121971. [Google Scholar] [CrossRef]
  66. Lizcano, F.; Vargas, D.; Gómez, Á.; Torrado, A. Human ADMC-derived adipocyte thermogenic capacity is regulated by IL-4 receptor. Stem Cells Int. 2017, 2017, 2767916. [Google Scholar] [CrossRef] [Green Version]
  67. Tsao, C.H.; Shiau, M.Y.; Chuang, P.H.; Chang, Y.H.; Hwang, J. Interleukin-4 regulates lipid metabolism by inhibiting adipogenesis and promoting lipolysis. J. Lipid Res. 2014, 55, 385–397. [Google Scholar] [CrossRef] [Green Version]
  68. Shiau, M.Y.; Chuang, P.H.; Yang, C.P.; Hsiao, C.W.; Chang, S.W.; Chang, K.Y.; Liu, T.M.; Chen, H.W.; Chuang, C.C.; Yuan, S.Y.; et al. Mechanism of Interleukin-4 Reducing Lipid Deposit by Regulating Hormone-Sensitive Lipase. Sci. Rep. 2019, 9, 11974. [Google Scholar] [CrossRef] [PubMed]
  69. Frederich, R.C.; Hamann, A.; Anderson, S.; Löllmann, B.; Lowell, B.B.; Flier, J.S. Leptin levels reflect body lipid content in mice: Evidence for diet-induced resistance to leptin action. Nat. Med. 1995, 1, 1311–1314. [Google Scholar] [CrossRef] [PubMed]
  70. Maffei, M.; Halaas, J.; Ravussin, E.; Pratley, R.E.; Lee, G.H.; Zhang, Y.; Fei, H.; Kim, S.; Lallone, R.; Ranganathan, S.; et al. Leptin levels in human and rodent: Measurement of plasma leptin and ob RNA in obese and weight-reduced subjects. Nat. Med. 1995, 1, 1155–1161. [Google Scholar] [CrossRef] [PubMed]
  71. Thaler, J.P.; Schwartz, M.W. Minireview: Inflammation and obesity pathogenesis: The hypothalamus heats up. Endocrinology 2010, 151, 4109–4115. [Google Scholar] [CrossRef] [PubMed]
  72. Vendrell, J.; Broch, M.; Vilarrasa, N.; Molina, A.; Gómez, J.M.; Gutiérrez, C.; Simón, I.; Soler, J.; Richart, C. Resistin, adiponectin, ghrelin, leptin, and proinflammatory cytokines: Relationships in obesity. Obes. Res. 2004, 12, 962–971. [Google Scholar] [CrossRef] [PubMed]
  73. Huang, X.; Yang, Z. Resistin’s, obesity and insulin resistance: The continuing disconnect between rodents and humans. J. Endocrinol. Investig. 2016, 39, 607–615. [Google Scholar] [CrossRef] [PubMed]
  74. Silha, J.V.; Krsek, M.; Skrha, J.V.; Sucharda, P.; Nyomba, B.L.G.; Murphy, L.J. Plasma resistin, adiponectin and leptin levels in lean and obese subjects: Correlations with insulin resistence. Eur. J. Endocrinol. 2003, 149, 331–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Rajala, M.W.; Qi, Y.; Patel, H.R.; Takahashi, N.; Banerjee, R.; Pajvani, U.B.; Sinha, M.K.; Gingerich, R.L.; Scherer, P.E.; Ahima, R.S. Regulation of resistin expression and circulating levels in obesity, diabetes, and fasting. Diabetes 2004, 53, 1671–1679. [Google Scholar] [CrossRef] [Green Version]
  76. Smith, S.R.; Bai, F.; Charbonneau, C.; Janderová, L.; Argyropoulos, G. A promoter genotype and oxidative stress potentially link resistin to human insulin resistance. Diabetes 2003, 52, 1611–1618. [Google Scholar] [CrossRef] [Green Version]
  77. Savage, D.B.; Sewter, C.P.; Klenk, E.S.; Segal, D.G.; Vidal-Puig, A.; Considine, R.V.; O’Rahilly, S. Resistin/Fizz3 Expression in Relation to Obesity and Peroxisome Proliferator Activated Receptor-γ Action in Humans. Diabetes 2001, 50, 2199–2202. [Google Scholar] [CrossRef] [Green Version]
  78. Way, J.M.; Görgün, C.Z.; Tong, Q.; Uysal, K.T.; Brown, K.K.; Harrington, W.W.; Oliver, W.R.; Willson, T.M.; Kliewer, S.A.; Hotamisligil, G.S. Adipose Tissue Resistin Expression is Severely Suppressed in Obesity and Stimulated by Peroxisome Proliferator-activated Receptor γ Agonists. J. Biol. Chem. 2001, 276, 25651–25653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Heilbronn, L.K.; Rood, J.; Janderova, L.; Albu, J.B.; Kelley, D.E.; Ravussin, E.; Smith, S.R. Relationship between serum resistin concentrations and insulin resistance in nonobese, obese, and obese diabetic subjects. J. Clin. Endocrinol. Metab. 2004, 89, 1844–1848. [Google Scholar] [CrossRef] [Green Version]
  80. Chang, Y.H.; Chang, D.M.; Lin, K.C.; Shin, S.J.; Lee, Y.J. Visfatin in overweight/obesity, type 2 diabetes mellitus, insulin resistance, metabolic syndrome and cardiovascular diseases: A meta-analysis and systemic review. Diabetes. Metab. Res. Rev. 2011, 27, 515–527. [Google Scholar] [CrossRef]
  81. Fantuzzi, G. Adiponectin and inflammation: Consensus and controversy. J. Allergy Clin. Immunol. 2008, 121, 326–330. [Google Scholar] [CrossRef]
  82. Frühbeck, G.; Catalán, V.; Rodríguez, A.; Gómez-Ambrosi, J. Adiponectin-leptin ratio: A promising index to estimate adipose tissue dysfunction. Relation with obesity-associated cardiometabolic risk. Adipocyte 2018, 7, 57–62. [Google Scholar] [CrossRef]
  83. Frühbeck, G.; Catalán, V.; Rodríguez, A.; Ramírez, B.; Becerril, S.; Salvador, J.; Colina, I.; Gómez-Ambrosi, J. Adiponectin-leptin ratio is a functional biomarker of adipose tissue inflammation. Nutrients 2019, 11, 454. [Google Scholar] [CrossRef] [Green Version]
  84. Yang, R.Z.; Lee, M.J.; Hu, H.; Pray, J.; Bin Wu, H.; Hansen, B.C.; Shuldiner, A.R.; Fried, S.K.; McLenithan, J.C.; Gong, D.W. Identification of omentin as a novel depot-specific adipokine in human adipose tissue: Possible role in modulating insulin action. Am. J. Physiol. Endocrinol. Metab. 2006, 290, E1253–E1261. [Google Scholar] [CrossRef] [PubMed]
  85. Tan, B.K.; Adya, R.; Randeva, H.S. Omentin: A Novel Link between Inflammation, Diabesity, and Cardiovascular Disease. Trends Cardiovasc. Med. 2010, 20, 143–148. [Google Scholar] [CrossRef] [PubMed]
  86. De Souza Batista, C.M.; Yang, R.Z.; Lee, M.J.; Glynn, N.M.; Yu, D.Z.; Pray, J.; Ndubuizu, K.; Patil, S.; Schwartz, A.; Kligman, M.; et al. Omentin plasma levels and gene expression are decreased in obesity. Diabetes 2007, 56, 1655–1661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Buechler, C.; Krautbauer, S.; Eisinger, K. Adipose tissue fibrosis. World J. Diabetes 2015, 6, 548–553. [Google Scholar] [CrossRef]
  88. Sun, K.; Tordjman, J.; Clément, K.; Scherer, P.E. Fibrosis and Adipose Tissue Dysfunction. Cell Metab. 2013, 18, 470–477. [Google Scholar] [CrossRef] [Green Version]
  89. Debari, M.K.; Abbott, R.D. Adipose tissue fibrosis: Mechanisms, models, and importance. Int. J. Mol. Sci. 2020, 21, 6030. [Google Scholar] [CrossRef]
  90. Divoux, A.; Tordjman, J.; Lacasa, D.; Veyrie, N.; Hugol, D.; Aissat, A.; Basdevant, A.; Guerre-Millo, M.; Poitou, C.; Zucker, J.-D.; et al. Fibrosis in human adipose tissue: Composition, distribution, and link with lipid metabolism and fat mass loss. Diabetes 2010, 59, 2817–2825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Abdennour, M.; Reggio, S.; Le Naour, G.; Liu, Y.; Poitou, C.; Aron-Wisnewsky, J.; Charlotte, F.; Bouillot, J.L.; Torcivia, A.; Sasso, M.; et al. Association of adipose tissue and liver fibrosis with tissue stiffness in morbid obesity: Links with diabetes and BMI loss after gastric bypass. J. Clin. Endocrinol. Metab. 2014, 99, 898–907. [Google Scholar] [CrossRef] [Green Version]
  92. Bel Lassen, P.; Charlotte, F.; Liu, Y.; Bedossa, P.; Le Naour, G.; Tordjman, J.; Poitou, C.; Bouillot, J.-L.; Genser, L.; Zucker, J.-D.; et al. The FAT Score, a Fibrosis Score of Adipose Tissue: Predicting Weight-Loss Outcome after Gastric Bypass. J. Clin. Endocrinol. Metab. 2017, 102, 2443–2453. [Google Scholar] [CrossRef]
  93. Marcelin, G.; Silveira, A.L.M.; Martins, L.B.; Ferreira, A.V.M.; Clément, K. Deciphering the cellular interplays underlying obesityinduced adipose tissue fibrosis. J. Clin. Investig. 2019, 129, 4032–4040. [Google Scholar] [CrossRef] [PubMed]
  94. Spencer, M.; Unal, R.; Zhu, B.; Rasouli, N.; McGehee, R.E.; Peterson, C.A.; Kern, P.A. Adipose tissue extracellular matrix and vascular abnormalities in obesity and insulin resistance. J. Clin. Endocrinol. Metab. 2011, 96, E1990–E1998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Lin, D.; Chun, T.-H.; Kang, L. Adipose extracellular matrix remodelling in obesity and insulin resistance. Biochem. Pharmacol. 2016, 119, 8–16. [Google Scholar] [CrossRef] [Green Version]
  96. Lackey, D.E.; Burk, D.H.; Ali, M.R.; Mostaedi, R.; Smith, W.H.; Park, J.; Scherer, P.E.; Seay, S.A.; McCoin, C.S.; Bonaldo, P.; et al. Contributions of adipose tissue architectural and tensile properties toward defining healthy and unhealthy obesity. Am. J. Physiol. Endocrinol. Metab. 2014, 306, E233–E246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. McCulloch, L.J.; Rawling, T.J.; Sjöholm, K.; Franck, N.; Dankel, S.N.; Price, E.J.; Knight, B.; Liversedge, N.H.; Mellgren, G.; Nystrom, F.; et al. COL6A3 is regulated by leptin in human adipose tissue and reduced in obesity. Endocrinology 2015, 156, 134–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Sun, K.; Park, J.; Gupta, O.T.; Holland, W.L.; Auerbach, P.; Zhang, N.; Goncalves Marangoni, R.; Nicoloro, S.M.; Czech, M.P.; Varga, J.; et al. Endotrophin triggers adipose tissue fibrosis and metabolic dysfunction. Nat. Commun. 2014, 5, 3485. [Google Scholar] [CrossRef] [Green Version]
  99. Lawler, H.M.; Underkofler, C.M.; Kern, P.A.; Erickson, C.; Bredbeck, B.; Rasouli, N. Adipose Tissue Hypoxia, Inflammation, and Fibrosis in Obese Insulin-Sensitive and Obese Insulin-Resistant Subjects. J. Clin. Endocrinol. Metab. 2016, 101, 1422–1428. [Google Scholar] [CrossRef] [Green Version]
  100. Baker, N.A.; Muir, L.A.; Washabaugh, A.R.; Neeley, C.K.; Chen, S.Y.P.; Flesher, C.G.; Vorwald, J.; Finks, J.F.; Ghaferi, A.A.; Mulholland, M.W.; et al. Diabetes-specific regulation of adipocyte metabolism by the adipose tissue extracellular matrix. J. Clin. Endocrinol. Metab. 2017, 102, 1032–1043. [Google Scholar] [CrossRef] [Green Version]
  101. Camastra, S.; Gastaldelli, A.; Mari, A.; Bonuccelli, S.; Scartabelli, G.; Frascerra, S.; Baldi, S.; Nannipieri, M.; Rebelos, E.; Anselmino, M.; et al. Early and longer term effects of gastric bypass surgery on tissue-specific insulin sensitivity and beta cell function in morbidly obese patients with and without type 2 diabetes. Diabetologia 2011, 54, 2093–2102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Curry, T.B.; Roberts, S.K.; Basu, R.; Basu, A.; Schroeder, D.; Joyner, M.J.; Miles, J.M. Gastric bypass surgery is associated with near-normal insulin suppression of lipolysis in nondiabetic individuals. Am. J. Physiol. Metab. 2011, 300, E746–E751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Evans, B.A.; Merlin, J.; Bengtsson, T.; Hutchinson, D.S. Adrenoceptors in white, brown, and brite adipocytes. Br. J. Pharmacol. 2019, 176, 2416–2432. [Google Scholar] [CrossRef]
  104. Duncan, R.E.; Ahmadian, M.; Jaworski, K.; Sarkadi-Nagy, E.; Sul, H.S. Regulation of lipolysis in adipocytes. Annu. Rev. Nutr. 2007, 27, 79–101. [Google Scholar] [CrossRef] [Green Version]
  105. Frikke-Schmidt, H.; O’Rourke, R.W.; Lumeng, C.N.; Sandoval, D.A.; Seeley, R.J. Does bariatric surgery improve adipose tissue function? Obes. Rev. 2016, 17, 795–809. [Google Scholar] [CrossRef] [Green Version]
  106. Kaartinen, J.M.; LaNoue, K.F.; Martin, L.F.; Vikman, H.L.; Ohisalo, J.J. β-Adrenergic responsiveness of adenylate cyclase in human adipocyte plasma membranes in obesity and after massive weight reduction. Metabolism 1995, 44, 1288–1292. [Google Scholar] [CrossRef]
  107. Reynisdottir, S.; Langin, D.; Carlstrom, K.; Holm, C.; Rossner, S.; Arner, P. Effects of weight reduction on the regulation of lipolysis in adipocytes of women with upper-body obesity. Clin. Sci. 1995, 89, 421–429. [Google Scholar] [CrossRef]
  108. Mauriège, P.; Imbeault, P.; Langin, D.; Lacaille, M.; Alméras, N.; Tremblay, A.; Després, J.P. Regional and gender variations in adipose tissue lipolysis in response to weight loss. J. Lipid Res. 1999, 40, 1559–1571. [Google Scholar] [CrossRef]
  109. Castellot, J.J.; Karnovsky, M.J.; Spiegelman, B.M. Differentiation-dependent stimulation of neovascularization and endothelial cell chemotaxis by 3T3 adipocytes. Proc. Natl. Acad. Sci. USA 1982, 79, 5597–5601. [Google Scholar] [CrossRef] [Green Version]
  110. Montesano, R.; Mouron, P.; Orci, L. Vascular outgrowths from tissue explants embedded in fibrin or collagen gels: A simple in vitro model of angiogenesis. Cell Biol. Int. Rep. 1985, 9, 869–875. [Google Scholar] [CrossRef]
  111. Varzaneh, F.E.; Shillabeer, G.; Wong, K.L.; Lau, D.C.W. Extracellular matrix components secreted by microvascular endothelial cells stimulate preadipocyte differentiation in vitro. Metabolism 1994, 43, 906–912. [Google Scholar] [CrossRef]
  112. Fukumura, D.; Ushiyama, A.; Duda, D.G.; Xu, L.; Tam, J.; Krishna, V.; Chatterjee, K.; Garkavtsev, I.; Jain, R.K. Paracrine regulation of angiogenesis and adipocyte differentiation during in vivo adipogenesis. Circ. Res. 2003, 93, e88–e97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Zhang, Q.X.; Magovern, C.J.; Mack, C.A.; Budenbender, K.T.; Wilson, K.; Rosengart, T.K. Vascular endothelial growth factor is the major angiogenic factor in omentum: Mechanism of the omentum-mediated angiogenesis. J. Surg. Res. 1997, 67, 147–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Thomas, M.; Augustin, H.G. The role of the angiopoietins in vascular morphogenesis. Angiogenesis 2009, 12, 125–137. [Google Scholar] [CrossRef] [PubMed]
  115. An, Y.A.; Sun, K.; Joffin, N.; Zhang, F.; Deng, Y.; Donzé, O.; Kusminski, C.M.; Scherer, P.E. Angiopoietin-2 in white adipose tissue improves metabolic homeostasis through enhanced angiogenesis. eLife 2017, 6, e24071. [Google Scholar] [CrossRef]
  116. Montoro, J.I.M.; García, J.C.F.; Vega, M.M.; Fuentes, M.D.; Tinahones, F.J. PECAM-1 levels are increased in morbidly obese young men. Atherosclerosis 2020, 315, e59–e60. [Google Scholar] [CrossRef]
  117. Kanasaki, K.; Kawakita, E.; Koya, D. Relevance of autophagy induction by gastrointestinal hormones: Focus on the incretin-based drug target and glucagon. Front. Pharmacol. 2019, 10, 476. [Google Scholar] [CrossRef]
  118. Ueno, T.; Komatsu, M. Autophagy in the liver: Functions in health and disease. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 170–184. [Google Scholar] [CrossRef]
  119. Soussi, H.; Reggio, S.; Alili, R.; Prado, C.; Mutel, S.; Pini, M.; Rouault, C.; Clément, K.; Dugail, I. DAPK2 Downregulation Associates With Attenuated Adipocyte Autophagic Clearance in Human Obesity. Diabetes 2015, 64, 3452–3463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Kovsan, J.; Blüher, M.; Tarnovscki, T.; Klöting, N.; Kirshtein, B.; Madar, L.; Shai, I.; Golan, R.; Harman-Boehm, I.; Schön, M.R.; et al. Altered autophagy in human adipose tissues in obesity. J. Clin. Endocrinol. Metab. 2011, 96, E268–E277. [Google Scholar] [CrossRef]
  121. Jansen, H.J.; Van Essen, P.; Koenen, T.; Joosten, L.A.B.; Netea, M.G.; Tack, C.J.; Stienstra, R. Autophagy activity is up-regulated in adipose tissue of obese individuals and modulates proinflammatory cytokine expression. Endocrinology 2012, 153, 5866–5874. [Google Scholar] [CrossRef] [Green Version]
  122. Nuñez, C.E.; Rodrigues, V.S.; Gomes, F.S.; De Moura, R.F.; Victorio, S.C.; Bombassaro, B.; Chaim, E.A.; Pareja, J.C.; Geloneze, B.; Velloso, L.A.; et al. Defective regulation of adipose tissue autophagy in obesity. Int. J. Obes. 2013, 37, 1473–1480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Öst, A.; Svensson, K.; Ruishalme, I.; Brännmark, C.; Franck, N.; Krook, H.; Sandström, P.; Kjolhede, P.; Strålfors, P. Attenuated mTOR signaling and enhanced autophagy in adipocytes from obese patients with type 2 diabetes. Mol. Med. 2010, 16, 235–246. [Google Scholar] [CrossRef] [PubMed]
  124. Rodríguez, A.; Gómez-Ambrosi, J.; Catalán, V.; Rotellar, F.; Valentí, V.; Silva, C.; Mugueta, C.; Pulido, M.R.; Vázquez, R.; Salvador, J.; et al. The ghrelin O-Acyltransferase-Ghrelin system reduces TNF-α-Induced apoptosis and autophagy in human visceral adipocytes. Diabetologia 2012, 55, 3038–3050. [Google Scholar] [CrossRef] [Green Version]
  125. Mahawar, K.K. Defining Short-term, Medium-term, Long-term, and Very Long-term Follow-up after Bariatric Surgery. Obes. Surg. 2018, 28, 1425–1426. [Google Scholar] [CrossRef]
  126. Meyer-Gerspach, A.C.; Peterli, R.; Moor, M.; Madörin, P.; Schötzau, A.; Nabers, D.; Borgwardt, S.; Beglinger, C.; Bieri, O.; Wölnerhanssen, B.K. Quantification of Liver, Subcutaneous, and Visceral Adipose Tissues by MRI Before and After Bariatric Surgery. Obes. Surg. 2019, 29, 2795–2805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Keidar, A.; Appelbaum, L.; Schweiger, C.; Hershkop, K.; Matot, I.; Constantini, N.; Sosna, J.; Weiss, R. Baseline Abdominal Lipid Partitioning Is Associated with the Metabolic Response to Bariatric Surgery. Obes. Surg. 2014, 24, 1709–1716. [Google Scholar] [CrossRef]
  128. Umemura, A.; Sasaki, A.; Nitta, H.; Otsuka, K.; Suto, T.; Wakabayashi, G. Effects of changes in adipocyte hormones and visceral adipose tissue and the reduction of obesity-related comorbidities after laparoscopic sleeve gastrectomy in Japanese patients with severe obesity. Endocr. J. 2014, 61, 381–391. [Google Scholar] [CrossRef] [Green Version]
  129. Dadson, P.; Ferrannini, E.; Landini, L.; Hannukainen, J.C.; Kalliokoski, K.K.; Vaittinen, M.; Honka, H.; Karlsson, H.K.; Tuulari, J.J.; Soinio, M.; et al. Fatty acid uptake and blood flow in adipose tissue compartments of morbidly obese subjects with or without type 2 diabetes: Effects of bariatric surgery. Am. J. Physiol. Endocrinol. Metab. 2017, 313, E175–E182. [Google Scholar] [CrossRef]
  130. Kim, M.K.; Jang, E.-H.; Hong, O.-K.; Chun, H.-J.; Yoo, S.-J.; Baek, K.-H.; Kim, W.; Kim, E.K.; Song, K.-H.; Kwon, H.-S. Changes in Serum Levels of Bone Morphogenic Protein 4 and Inflammatory Cytokines after Bariatric Surgery in Severely Obese Korean Patients with Type 2 Diabetes. Int. J. Endocrinol. 2013, 2013, 681205. [Google Scholar] [CrossRef] [Green Version]
  131. Kim, M.K.; Kim, W.; Kwon, H.-S.; Baek, K.-H.; Kim, E.K.; Song, K.-H. Effects of bariatric surgery on metabolic and nutritional parameters in severely obese Korean patients with type 2 diabetes: A prospective 2-year follow up. J. Diabetes Investig. 2014, 5, 221–227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Andersson, D.P.; Hogling, D.E.; Thorell, A.; Toft, E.; Qvisth, V.; Näslund, E.; Thörne, A.; Wirén, M.; Löfgren, P.; Hoffstedt, J.; et al. Changes in subcutaneous fat cell volume and insulin sensitivity after weight loss. Diabetes Care 2014, 37, 1831–1836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Hoffstedt, J.; Andersson, D.P.; Hogling, D.E.; Theorell, J.; Näslund, E.; Thorell, A.; Ehrlund, A.; Rydén, M.; Arner, P. Long-term protective changes in adipose tissue after gastric bypass. Diabetes Care 2017, 40, 77–84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Lehmann, S.; Linder, N.; Retschlag, U.; Schaudinn, A.; Stange, R.; Garnov, N.; Dietrich, A.; Oberbach, A.; Kahn, T.; Busse, H. MRI assessment of changes in adipose tissue parameters after bariatric surgery. PLoS ONE 2018, 13, e0206735. [Google Scholar] [CrossRef]
  135. Favre, L.; Marino, L.; Roth, A.; Acierno, J.; Hans, D.; Demartines, N.; Pitteloud, N.; Suter, M.; Collet, T.-H. The Reduction of Visceral Adipose Tissue after Roux-en-Y Gastric Bypass Is more Pronounced in Patients with Impaired Glucose Metabolism. Obes. Surg. 2018, 28, 4006–4013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Cotillard, A.; Poitou, C.; Torcivia, A.; Bouillot, J.-L.; Dietrich, A.; Klöting, N.; Grégoire, C.; Lolmede, K.; Blüher, M.; Clément, K. Adipocyte Size Threshold Matters: Link with Risk of Type 2 Diabetes and Improved Insulin Resistance After Gastric Bypass. J. Clin. Endocrinol. Metab. 2014, 99, E1466–E1470. [Google Scholar] [CrossRef]
  137. Hansen, M.; Lund, M.T.; Gregers, E.; Kraunsøe, R.; Van Hall, G.; Helge, J.W.; Dela, F. Adipose tissue mitochondrial respiration and lipolysis before and after a weight loss by diet and RYGB. Obesity 2015, 23, 2022–2029. [Google Scholar] [CrossRef]
  138. Torriani, M.; Oliveira, A.L.; Azevedo, D.C.; Bredella, M.A.; Yu, E.W. Effects of Roux-en-Y Gastric Bypass Surgery on Visceral and Subcutaneous Fat Density by Computed Tomography. Obes. Surg. 2015, 25, 381–385. [Google Scholar] [CrossRef]
  139. Illán-Gómez, F.; Gonzálvez-Ortega, M.; Orea-Soler, I.; Alcaraz-Tafalla, M.S.; Aragón-Alonso, A.; Pascual-Díaz, M.; Pérez-Paredes, M.; Lozano-Almela, M.L. Obesity and inflammation: Change in adiponectin, C-reactive protein, tumour necrosis factor-alpha and interleukin-6 after bariatric surgery. Obes. Surg. 2012, 22, 950–955. [Google Scholar] [CrossRef]
  140. Lips, M.A.; van Klinken, J.B.; Pijl, H.; Janssen, I.; Willems van Dijk, K.; Koning, F.; van Harmelen, V. Weight loss induced by very low calorie diet is associated with a more beneficial systemic inflammatory profile than by Roux-en-Y gastric bypass. Metabolism 2016, 65, 1614–1620. [Google Scholar] [CrossRef]
  141. Vilarrasa, N.; Vendrell, J.; Sánchez-Santos, R.; Broch, M.; Megia, A.; Masdevall, C.; Gomez, N.; Soler, J.; Pujol, J.; Bettónica, C.; et al. Effect of weight loss induced by gastric bypass on proinflammatory interleukin-18, soluble tumour necrosis factor-α receptors, C-reactive protein and adiponectin in morbidly obese patients. Clin. Endocrinol. 2007, 67, 679–686. [Google Scholar] [CrossRef]
  142. Schmatz, R.; Bitencourt, M.R.; Patias, L.D.; Beck, M.; Alvarez, G.D.; Zanini, D.; Gutierres, J.M.; Diehl, L.N.; Pereira, L.B.; Leal, C.A.; et al. Evaluation of the biochemical, inflammatory and oxidative profile of obese patients given clinical treatment and bariatric surgery. Clin. Chim. Acta 2017, 465, 72–79. [Google Scholar] [CrossRef]
  143. Viana, E.C.; Araujo-Dasilio, K.L.; Miguel, G.P.S.; Bressan, J.; Lemos, E.M.; Moyses, M.R.; De Abreu, G.R.; De Azevedo, J.L.M.C.; Carvalho, P.S.; Passos-Bueno, M.R.S.; et al. Gastric bypass and sleeve gastrectomy: The same impact on IL-6 and TNF-α. Prospective clinical trial. Obes. Surg. 2013, 23, 1252–1261. [Google Scholar] [CrossRef] [PubMed]
  144. Nestvold, T.K.; Nielsen, E.W.; Ludviksen, J.K.; Fure, H.; Landsem, A.; Lappegård, K.T. Lifestyle changes followed by bariatric surgery lower inflammatory markers and the cardiovascular risk factors C3 and C4. Metab. Syndr. Relat. Disord. 2015, 13, 29–35. [Google Scholar] [CrossRef] [PubMed]
  145. Farias, G.; Netto, B.D.M.; Boritza, K.; Bettini, S.C.; Vilela, R.M.; Dâmaso, A.R. Impact of Weight Loss on Inflammation State and Endothelial Markers Among Individuals with Extreme Obesity After Gastric Bypass Surgery: A 2-Year Follow-up Study. Obes. Surg. 2020, 30, 1881–1890. [Google Scholar] [CrossRef]
  146. Sdralis, E.; Argentou, M.; Mead, N.; Kehagias, I.; Alexandridis, T.; Kalfarentzos, F. A prospective randomized study comparing patients with morbid obesity submitted to sleeve gastrectomy with or without omentectomy. Obes. Surg. 2013, 23, 965–971. [Google Scholar] [CrossRef]
  147. Dillard, T.H.; Purnell, J.Q.; Smith, M.D.; Raum, W.; Hong, D.; Laut, J.; Patterson, E.J. Omentectomy added to Roux-en-Y gastric bypass surgery: A randomized, controlled trial. Surg. Obes. Relat. Dis. 2013, 9, 269–275. [Google Scholar] [CrossRef] [Green Version]
  148. Pardina, E.; Ferrer, R.; Baena-Fustegueras, J.A.; Rivero, J.; Lecube, A.; Fort, J.M.; Vargas, V.; Catalán, R.; Peinado-Onsurbe, J. Only C-reactive protein, but not TNF-α or IL6, reflects the improvement in inflammation after bariatric surgery. Obes. Surg. 2012, 22, 131–139. [Google Scholar] [CrossRef] [PubMed]
  149. Miller, G.D.; Nicklas, B.J.; Fernandez, A. Serial changes in inflammatory biomarkers after Roux-en-Y gastric bypass surgery. Surg. Obes. Relat. Dis. 2011, 7, 618–624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Lima, M.M.O.; Pareja, J.C.; Alegre, S.M.; Geloneze, S.R.; Kahn, S.E.; Astiarraga, B.D.; Chaim, É.A.; Baracat, J.; Geloneze, B. Visceral fat resection in humans: Effect on insulin sensitivity, beta-cell function, adipokines, and inflammatory markers. Obesity 2013, 21, E182–E189. [Google Scholar] [CrossRef]
  151. Bachmayer, C.; Lammert, A.; Hasenberg, T.; Hammes, H.P. Healthy obese and post bariatric patients—Metabolic and vascular patterns. Exp. Clin. Endocrinol. Diabetes 2013, 121, 483–487. [Google Scholar] [CrossRef]
  152. Salman, M.A.; Abdallah, A.; Mikhail, H.M.S.; Abdelsalam, A.; Ibrahim, A.H.; Sultan, A.A.E.A.; El-ghobary, M.; Ismail, A.A.M.; Abouelregal, T.E.; Omar, M.G.; et al. Long-term Impact of Mini-Gastric Bypass on Inflammatory Cytokines in Cohort of Morbidly Obese Patients: A Prospective Study. Obes. Surg. 2020, 30, 2338–2344. [Google Scholar] [CrossRef] [PubMed]
  153. Pérez-Romero, N.; Serra, A.; Granada, M.L.; Rull, M.; Alastrué, A.; Navarro-Díaz, M.; Romero, R.; Fernández-Llamazares, J. Effects of two variants of Roux-en-Y gastric bypass on metabolism behaviour: Focus on plasma ghrelin concentrations over a 2-year follow-up. Obes. Surg. 2010, 20, 600–609. [Google Scholar] [CrossRef]
  154. Manco, M.; Fernandez-Real, J.M.; Equitani, F.; Vendrell, J.; Mora, M.E.V.; Nanni, G.; Tondolo, V.; Calvani, M.; Ricart, W.; Castagneto, M.; et al. Effect of massive weight loss on inflammatory adipocytokines and the innate immune system in morbidly obese women. J. Clin. Endocrinol. Metab. 2007, 92, 483–490. [Google Scholar] [CrossRef]
  155. Jürets, A.; Itariu, B.K.; Keindl, M.; Prager, G.; Langer, F.; Grablowitz, V.; Zeyda, M.; Stulnig, T.M. Upregulated TNF Expression 1 Year After Bariatric Surgery Reflects a Cachexia-Like State in Subcutaneous Adipose Tissue. Obes. Surg. 2017, 27, 1514–1523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Lindegaard, K.K.; Jorgensen, N.B.; Just, R.; Heegaard, P.M.; Madsbad, S. Effects of Roux-en-Y gastric bypass on fasting and postprandial inflammation-related parameters in obese subjects with normal glucose tolerance and in obese subjects with type 2 diabetes. Diabetol. Metab. Syndr. 2015, 7, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Sajoux, I.; Lorenzo, P.M.; Gomez-Arbelaez, D.; Zulet, M.A.; Abete, I.; Castro, A.I.; Baltar, J.; Portillo, M.P.; Tinahones, F.J.; Martinez, J.A.; et al. Effect of a very-low-calorie ketogenic diet on circulating myokine levels compared with the effect of bariatric surgery or a low-calorie diet in patients with obesity. Nutrients 2019, 11, 2368. [Google Scholar] [CrossRef] [Green Version]
  158. Marantos, G.; Daskalakis, M.; Karkavitsas, N.; Matalliotakis, I.; Papadakis, J.A.; Melissas, J. Changes in metabolic profile and adipoinsular axis in morbidly obese premenopausal females treated with restrictive bariatric surgery. World J. Surg. 2011, 35, 2022–2023. [Google Scholar] [CrossRef]
  159. Rega-Kaun, G.; Kaun, C.; Ebenbauer, B.; Jaegersberger, G.; Prager, M.; Wojta, J.; Hohensinner, P.J. Bariatric surgery in morbidly obese individuals affects plasma levels of protein C and thrombomodulin. J. Thromb. Thrombolysis 2019, 47, 51–56. [Google Scholar] [CrossRef] [Green Version]
  160. Holdstock, C.; Lind, L.; Eden Engstrom, B.; Ohrvall, M.; Sundbom, M.; Larsson, A.; Karlsson, F.A. CRP reduction following gastric bypass surgery is most pronounced in insulin-sensitive subjects. Int. J. Obes. 2005, 29, 1275–1280. [Google Scholar] [CrossRef] [Green Version]
  161. Li, Y.; Guan, W.; Ma, S.; Lin, S.; Yang, N.; Liu, R.; Liang, H.; Zhou, H. Lipopolysaccharide and inflammatory cytokines levels decreased after sleeve gastrectomy in Chinese adults with obesity. Endocr. J. 2019, 66, 337–347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Mallipedhi, A.; Prior, S.L.; Barry, J.D.; Caplin, S.; Baxter, J.N.; Stephens, J.W. Changes in inflammatory markers after sleeve gastrectomy in patients with impaired glucose homeostasis and type 2 diabetes. Surg. Obes. Relat. Dis. 2014, 10, 1123–1128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Baltieri, L.; Cazzo, E.; de Souza, A.L.; Alegre, S.M.; de Paula Vieira, R.; Antunes, E.; de Mello, G.C.; Claudio Martins, L.; Chaim, E.A. Influence of weight loss on pulmonary function and levels of adipokines among asthmatic individuals with obesity: One-year follow-up. Respir. Med. 2018, 145, 48–56. [Google Scholar] [CrossRef] [PubMed]
  164. Dalmas, E.; Rouault, C.; Abdennour, M.; Rovere, C.; Rizkalla, S.; Bar-Hen, A.; Nahon, J.L.; Bouillot, J.L.; Guerre-Millo, M.; Clément, K.; et al. Variations in circulating inflammatory factors are related to changes in calorie and carbohydrate intakes early in the course of surgery-induced weight reduction1-3. Am. J. Clin. Nutr. 2011, 94, 450–458. [Google Scholar] [CrossRef] [PubMed]
  165. Monte, S.V.; Caruana, J.A.; Ghanim, H.; Sia, C.L.; Korzeniewski, K.; Schentag, J.J.; Dandona, P. Reduction in endotoxemia, oxidative and inflammatory stress, and insulin resistance after Roux-en-Y gastric bypass surgery in patients with morbid obesity and type 2 diabetes mellitus. Surgery 2012, 151, 587–593. [Google Scholar] [CrossRef] [PubMed]
  166. Fenske, W.K.; Dubb, S.; Bueter, M.; Seyfried, F.; Patel, K.; Tam, F.W.K.; Frankel, A.H.; Le Roux, C.W. Effect of bariatric surgery-induced weight loss on renal and systemic inflammation and blood pressure: A 12-month prospective study. Surg. Obes. Relat. Dis. 2013, 9, 559–568. [Google Scholar] [CrossRef]
  167. Wolf, R.M.; Jaffe, A.E.; Steele, K.E.; Schweitzer, M.A.; Magnuson, T.H.; Wolfe, A.; Wong, G.W. Cytokine, Chemokine, and Cytokine Receptor Changes Are Associated with Metabolic Improvements after Bariatric Surgery. J. Clin. Endocrinol. Metab. 2018, 104, 947–956. [Google Scholar] [CrossRef] [Green Version]
  168. Jouan, Y.; Blasco, H.; Bongrani, A.; Couet, C.; Dupont, J.; Maillot, F. Preoperative Chemerin Level Is Predictive of Inflammatory Status 1 Year After Bariatric Surgery. Obes. Surg. 2020, 30, 3852–3861. [Google Scholar] [CrossRef]
  169. Min, T.; Prior, S.L.; Dunseath, G.; Churm, R.; Barry, J.D.; Stephens, J.W. Temporal Effects of Bariatric Surgery on Adipokines, Inflammation and Oxidative Stress in Subjects with Impaired Glucose Homeostasis at 4 Years of Follow-up. Obes. Surg. 2020, 30, 1712–1718. [Google Scholar] [CrossRef] [Green Version]
  170. Shimizu, H.; Hatao, F.; Imamura, K.; Takanishi, K.; Tsujino, M. Early Effects of Sleeve Gastrectomy on Obesity-Related Cytokines and Bile Acid Metabolism in Morbidly Obese Japanese Patients. Obes. Surg. 2017, 27, 3223–3229. [Google Scholar] [CrossRef]
  171. De Luis, D.A.; Terroba, M.C.; Cuellar, L.; Conde, R.; Primo, D.; Aller, R.; Sagrado, M.G.; Izaola, O. Resistin levels in morbid obese patients following the biliopancreatic diversion surgery. Horm. Metab. Res. 2011, 43, 205–208. [Google Scholar] [CrossRef]
  172. Lapointe, M.; Poirier, P.; Martin, J.; Bastien, M.; Auclair, A.; Cianflone, K. Omentin changes following bariatric surgery and predictive links with biomarkers for risk of cardiovascular disease. Cardiovasc. Diabetol. 2014, 13, 16. [Google Scholar] [CrossRef] [Green Version]
  173. De Luis, D.A.; García Calvo, S.; Lopez Gomez, J.J.; Izaola, O.; Primo, D.; Pacheco, D.; Aller, R. Omentin-1 changes following biliopancreatic diversion and relationship with cardiovascular risk factors. Ann. Nutr. Metab. 2018, 73, 106–112. [Google Scholar] [CrossRef] [PubMed]
  174. Soriguer, F.; Garrido-Sanchez, L.; Garcia-Serrano, S.; Garcia-Almeida, J.M.; Garcia-Arnes, J.; Tinahones, F.J.; Garcia-Fuentes, E. Apelin levels are increased in morbidly obese subjects with type 2 diabetes mellitus. Obes. Surg. 2009, 19, 1574–1580. [Google Scholar] [CrossRef] [PubMed]
  175. Handisurya, A.; Riedl, M.; Vila, G.; Maier, C.; Clodi, M.; Prikoszovich, T.; Ludvik, B.; Prager, G.; Luger, A.; Kautzky-Willer, A. Serum vaspin concentrations in relation to insulin sensitivity following RYGB-induced weight loss. Obes. Surg. 2010, 20, 198–203. [Google Scholar] [CrossRef]
  176. Janke, J.; Engeli, S.; Boschmann, M.; Adams, F.; Böhnke, J.; Luft, F.C.; Sharma, A.M.; Jordan, J. Retinol-binding protein 4 in human obesity. Diabetes 2006, 55, 2805–2810. [Google Scholar] [CrossRef] [Green Version]
  177. Haider, D.G.; Schindler, K.; Prager, G.; Bohdjalian, A.; Luger, A.; Wolzt, M.; Ludvik, B. Serum retinol-binding protein 4 is reduced after weight loss in morbidly obese subjects. J. Clin. Endocrinol. Metab. 2007, 92, 1168–1171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Wang, X.; Huang, Y.; Gao, J.; Sun, H.; Jayachandran, M.; Qu, S. Changes of serum retinol-binding protein 4 associated with improved insulin resistance after laparoscopic sleeve gastrectomy in Chinese obese patients. Diabetol. Metab. Syndr. 2020, 12, 7. [Google Scholar] [CrossRef]
  179. Tschoner, A.; Sturm, W.; Engl, J.; Kaser, S.; Laimer, M.; Laimer, E.; Weiss, H.; Patsch, J.R.; Ebenbichler, C.F. Retinol-binding protein 4, visceral fat, and the metabolic syndrome: Effects of weight loss. Obesity 2008, 16, 2439–2444. [Google Scholar] [CrossRef]
  180. Chabot, K.; Gauthier, M.S.; Garneau, P.Y.; Rabasa-Lhoret, R. Evolution of subcutaneous adipose tissue fibrosis after bariatric surgery. Diabetes Metab. 2017, 43, 125–133. [Google Scholar] [CrossRef]
  181. Katsogiannos, P.; Kamble, P.G.; Boersma, G.J.; Karlsson, F.A.; Lundkvist, P.; Sundbom, M.; Pereira, M.J.; Eriksson, J.W. Early Changes in Adipose Tissue Morphology, Gene Expression, and Metabolism after RYGB in Patients with Obesity and T2D. J. Clin. Endocrinol. Metab. 2019, 104, 2601–2613. [Google Scholar] [CrossRef]
  182. Löfgren, P.; Hoffstedt, J.; Rydén, M.; Thörne, A.; Holm, C.; Wahrenberg, H.; Arner, P. Major Gender Differences in the Lipolytic Capacity of Abdominal Subcutaneous Fat Cells in Obesity Observed before and after Long-Term Weight Reduction. J. Clin. Endocrinol. Metab. 2002, 87, 764–771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Soriguer, F.; García-Serrano, S.; García-Almeida, J.M.; Garrido-Sánchez, L.; García-Arnés, J.; Tinahones, F.J.; Cardona, I.; Rivas-Marín, J.; Gallego-Perales, J.L.; García-Fuentes, E. Changes in the serum composition of free-fatty acids during an intravenous glucose tolerance test. Obesity 2009, 17, 10–15. [Google Scholar] [CrossRef]
  184. Carvalho, B.M.; Oliveira, A.G.; Ueno, M.; Araújo, T.G.; Guadagnini, D.; Carvalho-Filho, M.A.; Geloneze, B.; Lima, M.M.O.; Pareja, J.C.; Carvalheira, J.B.C.; et al. Modulation of double-stranded RNA-activated protein kinase in insulin sensitive tissues of obese humans. Obesity 2013, 21, 2452–2457. [Google Scholar] [CrossRef] [PubMed]
  185. Maymó-Masip, E.; Fernández-Veledo, S.; España, A.G.; Vázquez-Carballo, A.; Tinahones, F.J.; García-Fuentes, E.; Garrifo-Sanchez, L.; Del Mar Rodriguez, M.; Vendrell, J.; Chacón, M.R. The rise of soluble TWEAK levels in severely obese subjects after bariatric surgery may affect adipocyte-cytokine production induced by TNFα. J. Clin. Endocrinol. Metab. 2013, 98, E1323–E1333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Dharuri, H.; ’t Hoen, P.A.C.; van Klinken, J.B.; Henneman, P.; Laros, J.F.J.; Lips, M.A.; el Bouazzaoui, F.; van Ommen, G.J.B.; Janssen, I.; van Ramshorst, B.; et al. Downregulation of the acetyl-CoA metabolic network in adipose tissue of obese diabetic individuals and recovery after weight loss. Diabetologia 2014, 57, 2384–2392. [Google Scholar] [CrossRef]
  187. De La Torre, N.G.; Rubio, M.A.; Bordiú, E.; Cabrerizo, L.; Aparicio, E.; Hernández, C.; Sánchez-Pernaute, A.; Díez-Valladares, L.; Torres, A.J.; Puente, M.; et al. Effects of weight loss after bariatric surgery for morbid obesity on vascular endothelial growth factor-A, adipocytokines, and insulin. J. Clin. Endocrinol. Metab. 2008, 93, 4276–4281. [Google Scholar] [CrossRef] [Green Version]
  188. Wiewiora, M.; Mertas, A.; Gluck, M.; Nowowiejska-Wiewiora, A.; Czuba, Z.; Piecuch, J. Effect of Weight Loss Surgery on Biomarkers of Angiogenesis in Obese Patients. Obes. Surg. 2020, 30, 3417–3425. [Google Scholar] [CrossRef]
  189. Atkin, S.L.; Ramachandran, V.; Yousri, N.A.; Benurwar, M.; Simper, S.C.; McKinlay, R.; Adams, T.D.; Najafi-Shoushtari, S.H.; Hunt, S.C. Changes in blood microRNA expression and early metabolic responsiveness 21 days following bariatric surgery. Front. Endocrinol. 2019, 10, 773. [Google Scholar] [CrossRef]
  190. Sangiao-Alvarellos, S.; Theofilatos, K.; Barwari, T.; Gutmann, C.; Takov, K.; Singh, B.; Juiz-Valiña, P.; Varela-Rodríguez, B.M.; Outeiriño-Blanco, E.; Duregotti, E.; et al. Metabolic recovery after weight loss surgery is reflected in serum microRNAs. BMJ Open Diabetes Res. Care 2020, 8, e001441. [Google Scholar] [CrossRef]
  191. Ortega, F.J.; Mercader, J.M.; Moreno-Navarrete, J.M.; Nonell, L.; Puigdecanet, E.; Rodriquez-Hermosa, J.I.; Rovira, O.; Xifra, G.; Guerra, E.; Moreno, M.; et al. Surgery-induced weight loss is associated with the downregulation of genes targeted by MicroRNAs in adipose tissue. J. Clin. Endocrinol. Metab. 2015, 100, E1467–E1476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Macartney-Coxson, D.; Danielson, K.; Clapham, J.; Benton, M.C.; Johnston, A.; Jones, A.; Shaw, O.; Hagan, R.D.; Hoffman, E.P.; Hayes, M.; et al. MicroRNA Profiling in Adipose Before and After Weight Loss Highlights the Role of miR-223-3p and the NLRP3 Inflammasome. Obesity 2020, 28, 570–580. [Google Scholar] [CrossRef] [PubMed]
  193. Mysore, R.; Ortega, F.J.; Latorre, J.; Ahonen, M.; Savolainen-Peltonen, H.; Fischer-Posovszky, P.; Wabitsch, M.; Olkkonen, V.M.; Fernández-Real, J.M.; Haridas, P.A.N. MicroRNA-221-3p regulates angiopoietin-like 8 (ANGPTL8) expression in adipocytes. J. Clin. Endocrinol. Metab. 2017, 102, 4001–4012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Ortega, F.J.; Moreno, M.; Mercader, J.M.; Moreno-Navarrete, J.M.; Fuentes-Batllevell, N.; Sabater, M.; Ricart, W.; Fernández-Real, J.M. Inflammation triggers specific microRNA profiles in human adipocytes and macrophages and in their supernatants. Clin. Epigenetics 2015, 7, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Botella-Carretero, J.I.; Álvarez-Blasco, F.; Ángeles Martinez-García, M.; Luque-Ramírez, M.; San Millán, J.L.; Escobar-Morreale, H.F. The decrease in serum IL-18 levels after bariatric surgery in morbidly obese women is a time-dependent event. Obes. Surg. 2007, 17, 1199–1208. [Google Scholar] [CrossRef]
  196. Schernthaner, G.H.; Kopp, H.P.; Kriwanek, S.; Krzyzanowska, K.; Satler, M.; Koppensteiner, R.; Schernthaner, G. Effect of massive weight loss induced by bariatric surgery on serum levels of interleukin-18 and monocyte-chemoattractant-protein-1 in morbid obesity. Obes. Surg. 2006, 16, 709–715. [Google Scholar] [CrossRef]
  197. Catalán, V.; Gómez-Ambrosi, J.; Rodríguez, A.; Ramírez, B.; Rotellar, F.; Valentí, V.; Silva, C.; Gil, M.J.; Salvador, J.; Frühbeck, G. Increased Tenascin C And Toll-Like Receptor 4 Levels in Visceral Adipose Tissue as a Link between Inflammation and Extracellular Matrix Remodeling in Obesity. J. Clin. Endocrinol. Metab. 2012, 97, E1880–E1889. [Google Scholar] [CrossRef] [Green Version]
  198. Sangiorgi, G.M.; Cereda, A.; Porchetta, N.; Benedetto, D.; Matteucci, A.; Bonanni, M.; Chiricolo, G.; De Lorenzo, A. Endovascular Bariatric Surgery as Novel Minimally Invasive Technique for Weight Management in the Morbidly Obese: Review of the Literature. Nutrients 2021, 13, 2541. [Google Scholar] [CrossRef]
  199. O’Brien, P.E.; Hindle, A.; Brennan, L.; Skinner, S.; Burton, P.; Smith, A.; Crosthwaite, G.; Brown, W. Long-Term Outcomes After Bariatric Surgery: A Systematic Review and Meta-analysis of Weight Loss at 10 or More Years for All Bariatric Procedures and a Single-Centre Review of 20-Year Outcomes After Adjustable Gastric Banding. Obes. Surg. 2019, 29, 3–14. [Google Scholar] [CrossRef] [Green Version]
  200. O’Brien, P.E.; MacDonald, L.; Anderson, M.; Brennan, L.; Brown, W.A. Long-term outcomes after bariatric surgery: Fifteen-year follow-up of adjustable gastric banding and a systematic review of the bariatric surgical literature. Ann. Surg. 2013, 257, 87–94. [Google Scholar] [CrossRef]
  201. Carlsson, L.M.S.; Peltonen, M.; Ahlin, S.; Anveden, Å.; Bouchard, C.; Carlsson, B.; Jacobson, P.; Lönroth, H.; Maglio, C.; Näslund, I.; et al. Bariatric Surgery and Prevention of Type 2 Diabetes in Swedish Obese Subjects. N. Engl. J. Med. 2012, 367, 695–704. [Google Scholar] [CrossRef]
  202. Sjöström, L.; Lindroos, A.; Peltonen, M.; Torgerson, J.; Bouchard, C.; Carlsson, B.; Dahlgren, S.; Larsson, B.; Narbro, K.; Sjöström, C.D.; et al. Lifestyle, Diabetes, and Cardiovascular Risk Factors 10 Years after Bariatric Surgery. N. Engl. J. Med. 2004, 351, 2683–2693. [Google Scholar] [CrossRef] [PubMed]
  203. Adams, T.D.; Gress, R.E.; Smith, S.C.; Halverson, R.C.; Simper, S.C.; Rosamond, W.D.; LaMonte, M.J.; Stroup, A.M.; Hunt, S.C. Long-Term Mortality after Gastric Bypass Surgery. N. Engl. J. Med. 2007, 357, 753–761. [Google Scholar] [CrossRef] [Green Version]
  204. Gómez-Ambrosi, J.; Andrada, P.; Valentí, V.; Rotellar, F.; Silva, C.; Catalán, V.; Rodríguez, A.; Ramírez, B.; Moncada, R.; Escalada, J.; et al. Dissociation of body mass index, excess weight loss and body fat percentage trajectories after 3 years of gastric bypass: Relationship with metabolic outcomes. Int. J. Obes. 2017, 41, 1379–1387. [Google Scholar] [CrossRef] [PubMed]
  205. Varela-Rodríguez, B.M.; Juiz-Valiña, P.; Varela, L.; Outeiriño-Blanco, E.; Bravo, S.B.; García-Brao, M.J.; Mena, E.; Noguera, J.F.; Valero-Gasalla, J.; Cordido, F.; et al. Beneficial Effects of Bariatric Surgery-Induced by Weight Loss on the Proteome of Abdominal Subcutaneous Adipose Tissue. J. Clin. Med. 2020, 9, 213. [Google Scholar] [CrossRef] [Green Version]
  206. Jahansouz, C.; Serrot, F.J.; Frohnert, B.I.; Foncea, R.E.; Dorman, R.B.; Slusarek, B.; Leslie, D.B.; Bernlohr, D.A. Roux-en-Y Gastric Bypass Acutely Decreases Protein Carbonylation and Increases Expression of Mitochondrial Biogenesis Genes in Subcutaneous Adipose Tissue. Obes. Surg. 2015, 25, 2376–2385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Henegar, C.; Tordjman, J.; Achard, V.; Lacasa, D.; Cremer, I.; Guerre-Millo, M.; Poitou, C.; Basdevant, A.; Stich, V.; Viguerie, N.; et al. Adipose tissue transcriptomic signature highlights the pathological relevance of extracellular matrix in human obesity. Genome Biol. 2008, 9, 1–32. [Google Scholar] [CrossRef]
  208. Mardinoglu, A.; Heiker, J.T.; Gärtner, D.; Björnson, E.; Schön, M.R.; Flehmig, G.; Klöting, N.; Krohn, K.; Fasshauer, M.; Stumvoll, M.; et al. Extensive weight loss reveals distinct gene expression changes in human subcutaneous and visceral adipose tissue. Sci. Rep. 2015, 5, 14841. [Google Scholar] [CrossRef]
  209. van der Kolk, B.W.; Muniandy, M.; Kaminska, D.; Alvarez, M.; Ko, A.; Miao, Z.; Valsesia, A.; Langin, D.; Vaittinen, M.; Pääkkönen, M.; et al. Differential Mitochondrial Gene Expression in Adipose Tissue Following Weight Loss Induced by Diet or Bariatric Surgery. J. Clin. Endocrinol. Metab. 2021, 106, 1312–1324. [Google Scholar] [CrossRef]
  210. Hankir, M.K.; Seyfried, F. Do Bariatric Surgeries Enhance Brown/Beige Adipose Tissue Thermogenesis? Front. Endocrinol. 2020, 11, 275. [Google Scholar] [CrossRef]
  211. De Giorgi, S.; Campos, V.; Egli, L.; Toepel, U.; Carrel, G.; Cariou, B.; Rainteau, D.; Schneiter, P.; Tappy, L.; Giusti, V. Long-term effects of Roux-en-Y gastric bypass on postprandial plasma lipid and bile acids kinetics in female non diabetic subjects: A cross-sectional pilot study. Clin. Nutr. 2015, 34, 911–917. [Google Scholar] [CrossRef]
  212. Ryan, K.K.; Tremaroli, V.; Clemmensen, C.; Kovatcheva-Datchary, P.; Myronovych, A.; Karns, R.; Wilson-Pérez, H.E.; Sandoval, D.A.; Kohli, R.; Bäckhed, F.; et al. FXR is a molecular target for the effects of vertical sleeve gastrectomy. Nature 2014, 509, 183–188. [Google Scholar] [CrossRef] [Green Version]
  213. Magalhaes, I.; Kiaf, B.; Lehuen, A. iNKT and MAIT cell alterations in diabetes. Front. Immunol. 2015, 6, 341. [Google Scholar] [CrossRef] [PubMed]
  214. Magalhaes, I.; Pingris, K.; Poitou, C.; Bessoles, S.; Venteclef, N.; Kiaf, B.; Beaudoin, L.; Da Silva, J.; Allatif, O.; Rossjohn, J.; et al. Mucosal-associated invariant T cell alterations in obese and type 2 diabetic patients. J. Clin. Investig. 2015, 125, 1752–1762. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Neinast, M.D.; Frank, A.P.; Zechner, J.F.; Li, Q.; Vishvanath, L.; Palmer, B.F.; Aguirre, V.; Gupta, R.K.; Clegg, D.J. Activation of natriuretic peptides and the sympathetic nervous system following Roux-en-Y gastric bypass is associated with gonadal adipose tissues browning. Mol. Metab. 2015, 4, 427–436. [Google Scholar] [CrossRef]
  216. Langi, G.; Szczerbinski, L.; Kretowski, A. Meta-Analysis of Differential miRNA Expression after Bariatric Surgery. J. Clin. Med. 2019, 8, 1220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Table 1. Short-, medium- and long-term outcomes of bariatric surgery on fat depot parameters, circulating and adipose tissue expression levels of cytokines, adipokines, and microRNAs.
Table 1. Short-, medium- and long-term outcomes of bariatric surgery on fat depot parameters, circulating and adipose tissue expression levels of cytokines, adipokines, and microRNAs.
Short-TermMedium-TermLong-Term
Parameter≤3 m≈6 m1 y≥2 y≥5 y
Depot size
 Subcutaneous[126][126,127,128,129][126,127,130][126,131,132][133]
 Visceral[126,134][126,127,128,129][126,127,130,135][126,131,132][133]
Fat cell area
 Subcutaneous-[136,137][137,138][132][133]
 Visceral--[138]--
Proinflammatory cytokines
 TNF-α[139,140][141][142,143,144][145]-
=[146,147]=[148,149,150]=[146,148,150,151,152]=2 y [153], 3 y [154]-
SAT [155]
 IL-1β=[147]-[142]--
 IL-6=[147,150,156]=[141,149,150,157][139,142,143,144,150,152,156,158,159,160][24]-
[140,156,161][139,160,162], SAT [148]---
 IL-8=[147]=[157]=[152]--
[156][163]
[161,164]
 IL-18--[141]--
 MCP-1SAT [13][164,165][150,164,166]--
 TGF-β=[156]-[156]--
Anti-inflammatory cytokines
 IL-4=[140][167]--
MNC [64]
 IL-10=[147][141,167][168]=2 y [145], 4 y [169]-
=[162][164]-
[164]
 IL-13-[167][144]--
Proinflammatory adipokines
 Leptin[140,149,162,170][141,148,149,150,162,170][144,148,150,156,158,163,168]2 y [145,153],
3 y [154], 4 y [169]
-
 Resistin=[150][141][142,158,168][145]-
[150]=[150,171]
 Visfatin=[150]=[150]=[150]--
Anti-inflammatory adipokines
 Adiponectin[147,149,156][139,141][139,141,142,150,156,158,168]2 y [145,153],
3 y [154]
[133]
=[146,150,156]=[150,162]=4 y [169]
 Omentin[172][172][172,173]--
Other adipokines
 Apelin-[174]---
 Vaspin--[175]--
 RBP-4SAT [176][177,178][151][179]-
=[176]
Fibrosis
 Subcutaneous-=[180]---
Lipolysis
 Basal=[181]=Isolated SAT adipocytes [181]-=Male [182]=[133]
Female [133,182]SAT release [133]
 Stimulated=Isolated SAT adipocytes [181]Isolated SAT adipocytes vs. 1 m [181]=Male [182]-
Female [182]
 Insulin-supressed-[137,183][102]-
 FFA[148,181]=[148,184,185][101]=[102]-
=[186]=[148]
Angiogenesis
 VEGF-A--[187,188]--
 ANGPT-2,  follistatin,
 HGF, PECAM-1
--[188]--
Autophagy
 Subcutaneous-3–12 m post-BS [119]---
microRNAs

7 Circulating miRNAs [189]-=Circulating miR-99b [190]
15 SAT miRNAs [191]-
1 VAT and 13 SAT miRNAs [192] *Circulating miR-221, miR-222 [190]SAT miR-221-3p [193]
12 SAT miRNAs [194]
For cytokines, adipokines, lipolysis, and angiogenesis markers, data refer to circulating levels, unless otherwise stated. ↑, Increased; ↓, decreased; =, equal or inconclusive; -, no data; m, months; y, years; TNF-α, tumor necrosis factor-α; IL, interleukin; MCP-1, monocyte chemoattractant protein-1; TGF-β, transforming growth factor β; RBP-4, retinol binding protein 4; FFA, free fatty acids; VEGF-A, vascular endothelial growth factor A; ANGPT-2, angiopoietin-2; HGF, hepatocyte growth factor; PECAM-1, platelet endothelial cell adhesion molecule-1; SAT, subcutaneous adipose tissue; VAT, visceral adipose tissue; MNC, mononuclear cells. * after significant weight loss, collection time not reported.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Osorio-Conles, Ó.; Vidal, J.; de Hollanda, A. Impact of Bariatric Surgery on Adipose Tissue Biology. J. Clin. Med. 2021, 10, 5516. https://doi.org/10.3390/jcm10235516

AMA Style

Osorio-Conles Ó, Vidal J, de Hollanda A. Impact of Bariatric Surgery on Adipose Tissue Biology. Journal of Clinical Medicine. 2021; 10(23):5516. https://doi.org/10.3390/jcm10235516

Chicago/Turabian Style

Osorio-Conles, Óscar, Josep Vidal, and Ana de Hollanda. 2021. "Impact of Bariatric Surgery on Adipose Tissue Biology" Journal of Clinical Medicine 10, no. 23: 5516. https://doi.org/10.3390/jcm10235516

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop