Abstract

Nonalcoholic fatty liver disease (NAFLD) is a spectrum of disease, extending from simple steatosis to inflammation and fibrosis with a significant risk for the development of cirrhosis. It is highly prevalent and is associated with significant adverse outcomes both through liver-specific morbidity and mortality but, perhaps more important, through adverse cardiovascular and metabolic outcomes. It is closely associated with type 2 diabetes and obesity, and both of these conditions drive progressive disease toward the more advanced stages. The mechanisms that govern hepatic lipid accumulation and the predisposition to inflammation and fibrosis are still not fully understood but reflect a complex interplay between metabolic target tissues including adipose and skeletal muscle, and immune and inflammatory cells. The ability to make an accurate assessment of disease stage (that relates to clinical outcome) can also be challenging. While liver biopsy is still regarded as the gold-standard investigative tool, there is an extensive literature on the search for novel noninvasive biomarkers and imaging modalities that aim to accurately reflect the stage of underlying disease. Finally, although no therapies are currently licensed for the treatment of NAFLD, there are interventions that appear to have proven efficacy in randomized controlled trials as well as an extensive emerging therapeutic landscape of new agents that target many of the fundamental pathophysiological processes that drive NAFLD. It is highly likely that over the next few years, new treatments with a specific license for the treatment of NAFLD will become available.

ESSENTIAL POINTS
  • Nonalcoholic fatty liver disease (NAFLD) is the most prevalent liver condition worldwide and is tightly associated with obesity and type 2 diabetes.

  • Patients with NAFLD have both increased liver-related and cardiovascular morbidity and mortality; the majority of excess deaths occur through cardiovascular disease.

  • Accurate staging of liver disease is important for risk stratification, and this is currently achieved through liver biopsy although there are many established and developing platforms aiming to provide noninvasive and accurate assessments.

  • Lifestyle modification and bariatric surgery (where appropriate), alongside aggressive cardiovascular risk reduction form the mainstays of treatment, and this is best delivered through a multidisciplinary approach that combine dedicated hepatology and metabolic input.

  • The therapeutic landscape is developing rapidly, and although there are currently no licensed therapies, several potentially disease-modifying drugs are now in phase 3 clinical studies.

Nonalcoholic fatty liver disease (NAFLD) is recognized as the hepatic manifestation of metabolic syndrome frequently co-existing with obesity, dyslipidemia, and insulin resistance. It is characterized by hepatic steatosis in the absence of secondary causes such as significant alcohol consumption, chronic viral hepatitis, medications that induce steatosis, or other chronic liver diseases such as autoimmune hepatitis, hemochromatosis, or Wilson’s disease. NAFLD can be subcategorized into nonalcoholic fatty liver (NAFL; formerly simple steatosis) and nonalcoholic steatohepatitis (NASH). NAFL represents steatosis in the absence of histological evidence of hepatocyte injury or inflammation, while NASH is characterized by the presence of steatosis, ballooning degeneration, and lobular inflammation, with or without perisinusoidal fibrosis on liver histology, and has a greater propensity to progress to cirrhosis and hepatocellular carcinoma (HCC).

There are currently many challenges in the diagnosis and management of patients with NAFLD; the adverse outcomes associated with the condition seem clear, yet accurate staging of disease without the use of liver biopsy is not established. In the absence of licensed pharmacotherapy, lifestyle modification remains the mainstay of treatment, but this is difficult to implement and maintain. Despite these challenges, the field is moving rapidly, and in this review, we will summarize the literature detailing current thoughts and concepts in pathogenesis, diagnosis, staging, and clinical management. NAFLD is a condition with increasing prevalence in children; however, the scope of this review will focus largely on data from adult patients.

Epidemiology of NAFLD

Prevalence of NAFLD

Meta-analysis of studies using imaging modalities for diagnosis has confirmed NAFLD as the commonest liver disease worldwide affecting approximately 25% of the global population (1). The prevalence of NAFLD is highest in the Middle East (32%) and South America (30%), lowest in Africa (13%), and intermediate in Europe (24%), Asia (27%) and North America (24%). The global burden of NAFLD has rapidly increased over time with prevalence increasing from 15% to 25% between 2005 and 2010 (1). In the United States, prevalence increased 2.7-fold between 2003 and 2011 (2). However, estimates of NAFLD prevalence vary according to the modality used for diagnosis. Using liver biochemical tests (or liver function tests), abnormalities in the absence of other causes of liver disease as the primary method of diagnosis significantly underestimates the true prevalence of NAFLD by up to 10% (1). A summary of the major factors driving the increasing prevalence of NAFLD is presented schematically in Fig. 1.

Multiple mechanisms and risk factors contribute to the development of hepatic steatosis and NASH. Obesity, insulin resistance and type 2 diabetes remain the key drivers to disease progression. Abbreviations: BMI, body mass index; CVD, cardiovascular disease; DAMPs, damage-associated molecular patterns; DNL, de novo lipogenesis; FFA, free fatty acids, GC, glucocorticoids; HSD17B13, 17β-hydroxysteroid dehydrogenase type 13; IR, insulin resistance; MBOAT7, membrane bound O-acyltransferase domain containing 7; NASH, nonalcoholic steatohepatitis; PCOS, polycystic ovary syndrome; PNPLA3, phospholipase domain-containing protein 3; PAMPs, pathogen associated molecular patterns; ROS, reactive oxygen species; ↓T4, hypothyroidism; TM6SF2, transmembrane 6 superfamily 2.
Figure 1.

Multiple mechanisms and risk factors contribute to the development of hepatic steatosis and NASH. Obesity, insulin resistance and type 2 diabetes remain the key drivers to disease progression. Abbreviations: BMI, body mass index; CVD, cardiovascular disease; DAMPs, damage-associated molecular patterns; DNL, de novo lipogenesis; FFA, free fatty acids, GC, glucocorticoids; HSD17B13, 17β-hydroxysteroid dehydrogenase type 13; IR, insulin resistance; MBOAT7, membrane bound O-acyltransferase domain containing 7; NASH, nonalcoholic steatohepatitis; PCOS, polycystic ovary syndrome; PNPLA3, phospholipase domain-containing protein 3; PAMPs, pathogen associated molecular patterns; ROS, reactive oxygen species; ↓T4, hypothyroidism; TM6SF2, transmembrane 6 superfamily 2.

Incidence of NAFLD

There are fewer published data regarding the incidence of NAFLD. Given that the rates of obesity have increased 2- to 3-fold across the Americas, Europe, and Asia over the past 3 decades, it is generally assumed that NAFLD incidence is rising proportionally. Pooled incidence of NAFLD in Asia and Israel was found to be 52 and 28 per 1000 person-years, respectively (1). A recent community study in North America using coding data for case identification demonstrated NAFLD incidence increased 5-fold between 1997 and 2014 from 62 to 329 per 100 000 person-years (3). The increase was found to be disproportionally higher among adults aged 18 to 39 years in whom the incidence of NAFLD increased 7-fold from 20 to 140 per 100 000 person years. This upward trend was also demonstrated in a second North American study in which overall incidence increased over a 10-year period particularly in the young (<45 years) where incident cases increased by 7.5% each year (2). The rising incidence and shift in the epidemiology toward younger populations serves as a stark reminder that the global burden of NAFLD and its complications including HCC, liver failure, and cardiovascular events is only set to increase.

Epidemiology of NASH, advanced fibrosism, and HCC

Six to 30% of patients with ultrasound (US)-proven NAFLD will have biopsy-proven NASH, corresponding to overall population prevalence between 1.5% and 6.45% (1). Approximately 40% of NASH patients develop fibrosis progression, with an average annual fibrosis stage progression rate of 0.09. The annual HCC incident rate in NASH patients was 5.29 per 1000 person-years. Although this HCC risk is less than that observed in chronic viral hepatitis (4), given the high prevalence of NAFLD, the global burden of NASH-related HCC will continue to rise, and, indeed, HCC is now the fourth leading etiology of cancer mortality worldwide (5). A UK study found that NAFLD-associated HCC had increased 10-fold over a decade (6), highlighting the clear shift in the attributable proportion of HCC from viral hepatitis to NASH.

Ethnicity and NAFLD

The contribution of ethnicity to the prevalence and severity of NAFLD is complex and remains controversial. Studies have demonstrated that Hispanic patients have a disproportionately higher, and African-Americans, a lower prevalence of NAFLD compared with white populations, and while these observations have been confirmed in recent meta-analysis of epidemiological studies from the United States, the interpretation of detailed aspects of the data needs to be considered. Importantly, the ethnicity-related differences were smaller in high-risk groups (obesity and type 2 diabetes mellitus [T2D]) compared with population-based cohorts (7); the prevalence of NAFLD (and NASH) in high-risk Hispanic patients was not different to that of whites. Further studies have suggested that when Hispanic and white patients are matched for obesity, the severity of NASH, and advanced fibrosis are not different between the 2 groups (8). The data with regards to African Americans are a little clearer in that they do have a lower prevalence of NAFLD; however, NASH can develop as frequently as is seen in white patients and, there is no evidence to suggest any difference in the rates of advanced fibrosis with differing ethnicity (9). Overall, these data would suggest that ethnicity might have a relatively greater influence on NAFLD prevalence rather than its severity.

The causes for ethnicity related differences in NAFLD prevalence and severity are complex, with contributions from genetic and environmental factors, socioeconomic status, and differential access to health care. Although obesity and T2D are well established risk factors for NAFLD these factors cannot entirely explain for the observed racial differences considering that metabolic syndrome is highly prevalent in black compared with white populations (10), yet the latter group has far higher rates of NAFLD. Gene variants may have a greater influence over susceptibility to NAFLD prevalence; for example, the prevalence the polymorphism in patatin-like phospholipase domain-containing protein 3 (PNPLA3) in different racial groups was found in 49% of Hispanics, 23% in White persons and 17% in black persons (11).

The epidemiology of NAFLD in South Asian populations has been extensively reviewed. A recent systematic review and meta-analysis (that included data from 237 studies and >13 million participants), concluded that in Asia the prevalence of NAFLD was 29.6% and that the annual incidence rate was 50 cases per 1000 person-years (12). Furthermore, in Asian patients with biopsy-proven NAFLD, NASH was identified in approximately 63.5% of cases (1). Within south Asian populations living in western countries, NAFLD was most prevalent in those patients originating from Bangladesh (13). While the absolute prevalence and incidence rates of NAFLD and NASH are not markedly different compared with other populations (in fact, in absolute terms may actually be slightly lower) (14) for a given body mass index (BMI), the prevalence and severity of NAFLD are likely to be higher mirroring observations with respect to T2D. The reasons for this are not fully understood, but genetic variability, alterations in adipose tissue biology, and lifestyle differences (diet and physical activity) are likely to be important (15).

NAFLD, obesity, and T2D

The prevalence of NAFLD increases in high-risk groups particularly in those with obesity and T2D. In an unselected Italian population sample 91% of patients who were obese (BMI ≥ 30 kg/m2) and 67% of those who were overweight (BMI 25–30 kg/m2) had US evidence of NAFLD (16). Similarly, the pooled prevalence of biopsy-proven NAFLD in morbidly obese patients undergoing bariatric surgery was found to be 91% (17). There has been a global increase in obesity in the last 30 years, which has largely been accounted for by new cases arising in Asia as a result of urbanization, lifestyle changes, Westernized diet and overnutrition. The number of obese people in China was below 0.1 million in 1975 and rose to 43.2 million in 2014, accounting for 16.3% of global obesity (18). The number of obese people in India rose from 0.4 to 9.8 million in the same timeframe (19). This has correlated with year-on-year increases in NAFLD prevalence (20) even in traditionally rural areas (21, 22). In patients with T2D, 2 large studies from Italy reported NAFLD prevalence as 60% to 70% (23, 24), and data from the United Kingdom suggested a prevalence of 42.6% (25). Recent meta-analysis of studies in T2D demonstrated a pooled NAFLD prevalence estimate of 60%, although there was a high degree of heterogeneity among eligible studies with proportions ranging from 29.6% to 87.1% (26). In a global meta-analysis of studies in unselected general populations, obesity and T2D were present in 51% and 22.5% of those with NAFLD, respectively (27). Taken together, these data indicate the close association of obesity and T2D with NAFLD, and in these high-risk groups, NAFLD prevalence is roughly double that found in the general population (7).

NAFLD in patients with BMI <25 kg/m2

Although NAFLD is usually associated with obesity, it can occur in those with a BMI < 25 kg/m2, and this has been extensively and recently reviewed (28, 29). The prevalence of NAFLD ranges from 7% in the United States (30) to 19% in Asia (14), although ethnicity-specific cut-offs for BMI need to be considered when interpreting published epidemiological which can make direct comparisons challenging. Despite patients with lean NAFLD having a more favorable cardiac and metabolic profile than obese patients with NAFLD, they have worse metabolic parameters (dyslipidemia, hypertension, insulin resistance) relative to lean controls without NAFLD (31). This cardio-metabolic risk is in part attributable to alterations in body fat distribution particularly increased visceral adiposity (32). Polymorphisms in PNPLA3 and transmembrane 6 superfamily member 2 (TM6SF2) may also have a relatively greater contribution in the pathogenesis of nonobese versus obesity-associated NAFLD (see Genetic predisposition to NAFLD and disease progression section) (33–35). Data are conflicting regarding the histological correlates of lean NAFLD with studies suggesting both a reduced (35–37) and equivalent (38, 39) prevalence of NASH and/or fibrosis compared with NAFLD in overweight patients. Those with lean NAFLD may however have accelerated progression of fibrosis with 1 study showing a higher rate of developing cirrhosis or HCC over 20 years compared with overweight patients with NAFLD (39); however, this remains controversial. This suggests that once an individual develops NASH, obesity is unlikely to be the primary driver in fibrosis progression. The effect of lean NAFLD on overall prognosis is similarly controversial. While a retrospective study of 646 NAFLD patients (19% with lean NAFLD) found no increase in overall mortality (39), a prospective study of >1090 biopsy-proven NAFLD patients (11.5% with lean NAFLD), published in abstract form only, did demonstrate a higher overall mortality compared with overweight patients with NAFLD despite presenting with a healthier metabolic profile and less advanced fibrosis (40). Further studies are clearly required to better establish the pathogenesis, progression, and prognosis in this condition. Disentangling obesity from insulin resistance is important. Nonobese individuals who have NAFLD are more insulin resistant than those without NAFLD and indeed may be as insulin resistant as some obese individuals, and this may be the most important driver to disease progression (rather than obesity per se) (41). However, consensus remains that NAFLD in nonobese patients is not benign and represents an important adverse risk factor for cardiovascular and liver health.

Associations with other endocrine disease

NAFLD is also associated with other endocrine conditions, and this has been reviewed previously (42). Hypogonadism in men and androgen excess in women are both associated with hepatic steatosis (43). Testosterone treatment has been shown to reduce liver fat in men (44), and we have shown that limiting androgen availability using 5α-reductase inhibition can increase hepatic steatosis (45). In patients with polycystic ovary syndrome (PCOS), there is continued debate as to the relative contributions of androgen excess and systemic insulin resistance as the main driver of the increased prevalence of NAFLD (46). There are currently no published data that have examined the impact of limiting androgen action in patients with PCOS as a strategy to improve NAFLD. Glucocorticoid excess is associated with NAFLD (47), and furthermore limiting endogenous glucocorticoid availability through the prereceptor inhibition of the cortisol-regenerating enzyme, 11β-hydroxysteroid dehydrogenase type 1, deceases liver fat content (48). Hypothyroidism is also associated with NAFLD (49), and it is interesting to note that liver-specific thyroid receptor agonism is being developed as a possible therapeutic strategy (see Thyroid hormone receptor-β agonists section) (50).

NALFD has a sexually dimorphic prevalence (see Gender and age in NAFLD, NASH and HCC section), and estrogens are likely to play a crucial role. Estradiol is believed to be protective for the development of NAFLD (51), and its prevalence increases in postmenopausal women. There is also evidence to suggest that postmenopausal hormone replacement therapy can protect against NAFLD (52, 53). Growth hormone (GH) deficiency is also linked to the development of NAFLD, with some evidence showing that GH replacement is associated with improvements in liver histology (54).

Gender and age in NAFLD, NASH, and HCC

NAFLD is significantly more prevalent in men than women in the general population (55) although NAFLD in people with a BMI < 25 kg/m2 may be more prevalent in women (30). In the Dallas Heart Study, white men had an approximately 2-fold increased prevalence of NAFLD compared with white women (56). Several mechanisms may contribute to this observation including body fat distribution, hepatic fatty acid partitioning, lifestyle, and sex hormone metabolism (55, 57, 58). There also appears to be gender-specific differences in relation to age, with NAFLD prevalence differing relatively little across all age groups in men (59) but increasing significantly after the age of 50 years in women (59, 60). Menopausal status is implicated in this age–sex interaction with up to double the rate of NAFLD found in postmenopausal compared with premenopausal women (59, 61–63). The protective role of estrogens is further supported by the observation that the risk of NAFLD increases in young women having oophorectomy (64) and in those on tamoxifen (65), and that there is a risk-reduction in those receiving hormone replacement therapy (61). In those patients with NAFLD, the risk of NASH may be higher in women and data from the NASH Clinical Research Network (CRN) showed that patients with biopsy-proven NASH were more likely to be female than male in a roughly 2:1 ratio. While this may reflect a higher disease burden in women, it may be accounted for by sex differences among those pursuing and receiving healthcare. Data regarding the impact of gender on fibrosis progression are conflicting. While one systematic review of risk factors for fibrosis progression in NASH found associations with age but not with gender (66), several cross-sectional biopsy studies have shown an increased risk of advanced fibrosis in females compared with males, independent of metabolic risk factors (67, 68). Furthermore, duration of estrogen deficiency in the postmenopausal state does confer fibrosis risk among women with NAFLD (69, 70). Irrespective of etiology of liver disease, HCC has a strong male preponderance, with a male to female ratio estimated to be 2:1 to 2.5:1. However the male–female ratio may be lower in NASH-associated HCC (1.6:1) compared with hepatitis B and C virus (HCV) and alcohol-related cirrhosis. Male patients with NASH also appear to develop HCC at a less advanced stage of liver fibrosis than female patients (71), and their overall survival is worse compared with women (72, 73). Overall prognosis in NAFLD also seems to be effected by gender with National Health and Nutrition Examination Survey (NHANES III) data from North America showing that male gender was significantly associated with overall mortality (74).

Sleep disturbance and NAFLD

Obstructive sleep apnea (OSA) is a highly prevalent disorder affecting 9% to 38% (75) of the general population and is associated with obesity in up to 60% of cases (76). The disease has become increasingly clinically relevant due to its association with cardiovascular morbidity and mortality (77–79). A recent meta-analysis found OSA to be independently associated with the presence of NAFLD and with degree of hepatic steatosis, lobular inflammation, ballooning degeneration, and fibrosis (80). OSA is characterized by the repetitive occurrence of partial or complete pharyngeal collapse during sleep leading over time to chronic intermittent hypoxia, which has been implicated in the pathogenesis and progression of NAFLD (81) through mitochondrial dysfunction (82–84) and increased hepatic de novo lipogenesis (DNL) (85, 86), reactive oxygen species (84, 85, 87), gut permeability (88), and liver inflammation (83, 89). Nocturnal continuous positive airway pressure (CPAP) is the gold-standard treatment for OSA and improves cognitive function, daytime somnolence, and quality of life. Despite observational data suggesting CPAP may improve liver function tests in those with OSA and suspected NAFLD (90) and improve cardiovascular disease (CVD) (91), randomized control trials in adults have failed to demonstrate a significant benefit on metabolic syndrome (92) and noninvasive markers of steatosis, hepatic inflammation, and fibrosis (93, 94).

Diet and NAFLD

Westernized versus Mediterranean diet in NAFLD

NAFLD prevalence has mirrored the global epidemic of obesity and T2D and is associated with consumption of a Westernized diet (95–102) characterized by high intake of fast foods, confections, refined grains, red and processed meats, full-fat dairy products, and soft drinks. NAFLD patients have a higher median daily total energy intake compared with age- and sex-matched healthy controls (100, 103). By contrast, the Mediterranean diet, characterized by low consumption of saturated fat and cholesterol, high consumption of monounsaturated fatty acids, a balanced omega-6 to omega-3 fatty acid ratio, and high content of complex carbohydrates and fiber, has been associated with lower rates of NAFLD (98, 104–106), NASH (106), and fibrosis (107) as well as cardiovascular events and cancer (108–110). Randomized trials have shown the Mediterranean diet to reduce plasma alanine aminotransferase (ALT) levels in obese patients with T2D (111) and improve insulin sensitivity and hepatic steatosis measured by magnetic resonance spectroscopy (MRS) in patients with NAFLD (112, 113), independent of weight loss. The biological mechanisms involved in these improvements are likely to include the anti-inflammatory and lipid lowering properties of the Mediterranean diet and its impact on gut microbiota composition (114). The Mediterranean dietary pattern is therefore the current recommended macronutrient composition in NAFLD and is recommended in the joint European Association for the Study of the Liver (EASL)–European Association for the Study of Diabetes–European Association for the Study of Obesity clinical practice guidelines (115).

Fructose in NAFLD

Fructose is a major component of the 2 most widely used sweeteners: sucrose and high-fructose corn syrup. Its consumption, predominantly in the form of soft drinks, has increased significantly over the last 100 years and now makes up around 15% of the energy consumed as part of a Westernized diet (116). Patients with NAFLD consume nearly 3 times more fructose compared with age-, sex-, and BMI-matched controls (117), and daily intakes of sucrose sweetened drinks significantly increases hepatic and visceral fat accumulation compared with milk, diet cola, and water (118). Fructose consumption was associated with higher fibrosis stage in 427 adults enrolled in the NASH CRN (119). Furthermore, low fructose intake may protect against future development of NAFLD in obese individuals (120). Mechanistically, fructose appears to have a major role in inducing hepatic steatosis both by stimulating hepatic DNL and attenuating β-fatty acid oxidation (121–125). In contrast to glucose, fructose is rapidly phosphorylated by fructokinase leading to a reduction in adenosine triphosphate (ATP). This fall in ATP induces a cascade of impaired protein synthesis (126), oxidative stress, and mitochondrial dysfunction (127, 128), features well established in the pathogenesis of NASH. The drop in intracellular phosphate also drives rapid purine nucleotide turnover culminating in the formation of uric acid (124, 129). Fructose is the only common carbohydrate to generate uric acid during its metabolism, which has also been implicated in the pathogenesis of NAFLD through increased oxidative stress and impaired β-oxidation (127, 128). Epidemiological studies have demonstrated an association between hyperuricemia and NAFLD (130), with meta-analysis showing a dose-dependent rise in incidence of NAFLD by 3% for every 1 mg/dL increase in serum uric acid (131).

Coffee, caffeine, and NAFLD

Coffee is one of the most commonly consumed beverages worldwide with 85% of the US population drinking caffeine daily, 98% of which is accounted for by coffee (132). An association between coffee consumption and a decreased risk of liver disease has been documented in a range of etiologies including NAFLD (133). Using dietary intake data from four continuous cycles of the National Health and Nutrition Examination Surveys, Bierdinc et al. demonstrated caffeine intake, mostly as coffee, to be independently associated with a lower risk of NAFLD (odds ratio [OR] 0.931, 95% confidence interval [CI] 0.900–0.964) (134). Similarly, 2 case-control studies have shown a beneficial effect of coffee on the risk of NAFLD; Catalano et al. and Gutiérrez-Grobe et al. both found that coffee drinking was associated with less severe steatosis on US in Italian and Mexican patients with NAFLD, respectively (135, 136). In contrast, a recent large Italian population study including 2800 participants found that coffee drinking was not associated with decreased odds for hepatic steatosis detected on US (137). Coffee has also been linked with a significant reduction in the risk of fibrosis among NASH patients (138) and those with morbid obesity, but only with filter coffee and not espresso coffee (139). This may be due to the confounding fibrogenic effect of sucrose (glucose + fructose), which may be added in greater quantity to espresso coffee as compared to filter coffee. Recent meta-analysis of observational studies found that regular coffee consumption was significantly associated with reduced hepatic fibrosis in patients with NAFLD, but not with NAFLD prevalence. This antifibrotic effect seems to be related to coffee and not to caffeine specifically, and further work is needed to identify which of the >100 compounds found within coffee confers hepatoprotection. More broadly, among coffee drinkers, a lower mortality and HCC rate has been well documented in patients with cirrhosis across a range of chronic liver disease etiologies (133). As a result, coffee consumption is often recommended for patients with chronic liver disease while recognizing the need for randomized controlled trials and further investigation into the threshold levels of coffee consumption required for benefit.

Pathogenesis of NAFLD

Development of hepatic steatosis—a multiorgan disease

Intrahepatic fat accumulation

Intrahepatic free fatty acids (FFA) can be esterified with glycerol and stored in the form triglyceride (TAG), the predominant lipid accumulating in patients with NAFLD. Alternatively, FFA can undergo β-oxidation, or TAG can be exported from the liver in very low-density lipoprotein (VLDL). Steatosis in NAFLD represents an inherent imbalance in these processes of lipid influx and synthesis versus disposal. There are 3 sources of FFA that contribute to liver TAG in NAFLD: 59% comes from circulating FFA; 26%, from the generation of FFA from nonlipid precursors (including glucose and fructose) through DNL; and 14%, from the diet (140).

Adipose tissue dysfunction

Under normal conditions, adipose tissue is extremely insulin responsive, storing lipid and inhibiting TAG lipolysis. Conversely, with developing insulin resistance, increased circulating FFA are made available for hepatic uptake and storage, at least in part through the impaired ability of insulin to suppress lipolysis (141). Adipose dysfunction has a major role in modulating severity of liver injury and cardiometabolic risk in NASH (142–144). Targeting adipose insulin resistance with weight loss (145) or thiazolidinediones (146) is likely therefore to be of therapeutic benefit (see Peroxisome proliferator-activated receptor gamma agonists section). In turn, NASH itself drives hepatic and peripheral insulin resistance, thereby setting up a vicious cycle of insulin resistance, hepatic FFA flux, steatosis, and inflammation (141, 147) (Fig. 1).

While obesity is a major determinant of the high global prevalence of NAFLD, the precise connection between adiposity and NAFLD remains unclear. Expansion of the peripheral adipose depot may provide a buffering capacity, protecting the liver from excessive FFA flux. This is exemplified in those patients with lipodystrophy, a group of syndromes characterized by partial or complete absence of adipose tissue but severe insulin resistance, ectopic fat accumulation, NAFL, and NASH (148). However, even metabolically healthy obese patients, who may be expected to be relatively protected from steatosis, are still at significant increased risk of NAFLD (149), suggesting that obesity per se is a risk factor for NAFLD independent of insulin sensitivity. Dysfunctional adipose tissue also leads to reduced secretion of adiponectin, an insulin-sensitizing adipokine paradoxically reduced in obesity. Adiponectin has wide-ranging hepatoprotective effects increasing FFA oxidation and decreasing FFA influx, gluconeogenesis, and DNL (150–152). It also has hepatic anti-inflammatory and antifibrotic properties through the suppression of proinflammatory cytokines (eg, tumor necrosis factor alpha and interleukin [IL]-6) (153, 154) and reduced activation and proliferation of hepatic stellate cells (HSC), respectively (155–157). Levels of circulating adiponectin are significantly reduced in patients with NAFLD and NASH and correlate with the degree of hepatic steatosis, necroinflammation, and fibrosis (158–160). Furthermore, plasma levels have been found to be the best indicator of histological improvement of NASH with pioglitazone (161). Adiponectin treatment has been shown to ameliorate NASH in rodent models (162), but human data are currently lacking.

Intrahepatic de novo lipogenesis

The second major source of intrahepatic FFA is synthesized from nonlipid precursors, predominantly glucose and fructose, through DNL. Upregulation of DNL is a hallmark of NAFLD with stable isotope infusion studies showing up to a 3.5-fold increase in patients with NAFLD compared with healthy controls (163, 164). Furthermore, while adipose-derived FFA accounts for the majority of liver TAG, the contribution from this source is less upregulated in NAFLD compared with marked increase in DNL that is seen in NAFLD (163). DNL is a multistep, highly regulated process during which acetyl-coenzyme A (CoA), derived from carbohydrate glycolysis, is converted to malonyl-CoA and, ultimately, palmitic acid under the control of several key enzymes and transcription factors (including sterol regulatory element-binding protein-1c, fatty acid synthase, and acetyl-CoA carboxylase 1), all of which are overexpressed in NAFLD (165, 166). Insulin promotes lipogenesis through the transcription and activation of sterol regulatory element-binding protein-1c, a master regulator of lipogenesis (167), and even in insulin-resistant conditions such as T2D, obesity, and NAFLD, insulin continues to selectively support DNL while failing to reduce hepatic gluconeogenesis (168, 169). Deciphering the precise mechanisms for this differential regulation of hepatic lipid and glucose metabolism has proven difficult with studies implicating a range of insulin-dependent (170–172) and independent molecular pathways (173). Dietary sugars form a major source of substrate for hepatic DNL (169, 174) and even short-term hypercaloric diet, overfeeding with simple carbohydrate markedly increases liver fat through DNL (174). The postprandial surge of intrahepatic carbohydrate via the portal circulation contrasts with dietary lipid, which is absorbed as chylomicrons via lymphatics into the systemic circulation where it is distributed to all metabolically active tissues. This explains the relatively small direct contribution of dietary lipid to intrahepatic TAG found in patients with NAFLD (140). Dietary fructose is particularly lipogenic and a better substrate for DNL than glucose and its increasing consumption has been implicated in the rising global burden of NAFLD (see Fructose in NAFLD section). Furthermore, fructose metabolism is highly energy dependent, which can compound liver injury through a cascade of impaired protein synthesis (126), oxidative stress, and mitochondrial dysfunction (127, 128). Skeletal muscle insulin resistance may also contribute to steatosis by causing a redistribution of postprandial glucose away from peripheral glycogen storage and toward hepatic DNL (175, 176) (Fig. 1). A large population study has also demonstrated skeletal muscle mass to be inversely correlated with NAFLD incidence and positively associated with the resolution of baseline NAFLD (177).

Triglyceride as an epiphenomenon alongside liver injury

Although TAG is the most conspicuous intrahepatic lipid in NAFLD, there are widespread changes in liver lipid composition including diacylglycerol, ceramides, TAG/diacylglycerol ratio, free cholesterol, and phospholipids, which are likely to be of greater pathogenic significance, and these have been reviewed previously (178–180). Indeed, TAG per se is not hepatotoxic, despite its accumulation traditionally being the basis for the clinical grading of NAFLD “severity.” For example, inhibition of TAG incorporation into VLDL leads to increased TAG accumulation but without liver injury and inhibition of diacylglycerol acyltransferase 2 (DGAT2), an enzyme involved in TAG formation, results in a reduction of intrahepatic TAG and worsening of steatohepatitis in mouse models (181). TAG accumulation can therefore be seen as a protective response to the increased burden of hepatotoxic FFA and an epiphenomenon occurring alongside liver injury (181).

The transition to NASH and the development of fibrosis

Hepatic inflammation, immune cells, and inflammasome activation

In NAFLD, the convergence of multiple toxic insults on the liver including FFAs, insulin resistance, gut derived endotoxins and adipose tissue dysfunction culminates in a profound proinflammatory state ultimately propagating NASH and fibrosis. As an example, patients with NASH having increased hepatic and adipose tissue concentrations of tumor necrosis factor-alpha (TNF-α) compared with obese controls, which positively correlates with fibrosis severity (182). Furthermore, the persistent activation or overexpression of transcription factor, nuclear factor kappa light chain enhancer of activated B cells, a crucial regulator of the acute inflammatory response has been found in animal models of NAFLD and insulin resistance (183) as well as patients with NASH (184) (Fig. 1).

The immune response plays a crucial role in triggering and amplifying hepatic inflammation (185). With hepatocyte injury, host biomolecules called damage-associated molecular patterns (DAMPs) are released. These are able to drive inflammation by activating resident macrophages, Kupffer cells (KC) acting via pattern recognition receptors, the best characterized of which are the toll-like receptor (TLR) family. Activated KC then drive the production of proinflammatory cytokines including TNF-α, IL-1b, IL-6, and C–C motif ligand 2 and 5, which exacerbate hepatocyte injury and cell death, predominantly through apoptosis (184). KC also secrete tumor growth factor beta and platelet-derived growth factor, providing additional stimulation to HSC, which promote fibrosis through up-regulation of smooth muscle actin, desmin and type I collagen. KC and HSC can also be activated directly by gut-derived bacterial products, including lipopolysaccharide (LPS), collectively known as pathogen-associated molecular patterns (PAMPs). Suppression of TLR-4, the receptor for LPS, and TLR-9, which binds bacterial deoxyribonucleic acid, have been shown to reduce liver inflammation in experimental rodent models of NASH (186, 187).

DAMPs, PAMPs, and KC also have an important role in the activation of inflammasomes, which are multiprotein intracellular complexes assembled in response to cell damage or pathogens to produce proinflammatory cytokines IL-1β and IL-18. Inflammasomes have a pivotal role in fine tuning the host inflammatory response and have been implicated in a range of acute and chronic liver diseases including NASH (188). Expression of inflammasome-associated proteins are significantly elevated in patients with NASH compared to those with simple steatosis (189), and FFA were found to directly upregulate inflammasome activation and sensitize hepatocytes to the effects of LPS in rodent models of NASH (190). Conversely, inflammasome deficient mice are relatively protected from diet-induced NASH and fibrosis (189).

Neutrophil infiltration is often seen in the livers of patients with NASH and contributes to macrophage recruitment and cell damage through the release of reactive oxygen species (ROS) and inflammatory mediators such as myeloperoxidase and elastase (191). Deficiency of these enzymes in knockout mice attenuates the development of NASH (192) and insulin resistance (193) and diminishes adipose tissue inflammation in response to a high fat diet (192). Neutrophil activation may also contribute to NASH progression with secreted extracellular hepatic neutrophil peptides enhancing liver fibrosis via HSC activation in steatotic mice (194). Beyond KC and neutrophils, however, the role of other innate immune cells is less clear. Natural killer cells have a complex role in cell damage detection having the ability to both enhance and limit the magnitude of inflammation (195). As such, disentangling their involvement in NASH has proven difficult with discrepant data showing both pathogenic and protective influences (196, 197). The role of dendritic cells (DC) is similarly controversial; despite these cells assuming an activated immune phenotype in NASH with increased expression of TNF-α and IL-6, ablation of DC populations is paradoxically associated with worse hepatic inflammation and fibrosis (198). The adaptive immune response is also mobilized in parallel with the innate response with lymphocytes, along with macrophages representing the most abundant inflammatory cells in the lobular and periportal infiltrates of NASH. Experiments in mice demonstrate that activated CD4+ and CD8+ T-lymphocytes are recruited within the liver through DC maturation (198) and upregulated vascular adhesion protein-1 (199) and their prevalence parallels worsening parenchymal injury and lobular inflammation (200). This is analogous to findings in obesity where adipose tissue CD4+/CD8+ T lymphocytes have an important role in supporting macrophages and generating proinflammatory mediators in the cascade toward insulin resistance (201). Data regarding the role of B-lymphocytes is relatively lacking. While studies have shown elevated titers of IgG to antigens originating from oxidative stress in methionine-choline deficient mice (200), further investigation is needed to better clarify the role of humoral immunity in NASH.

Mitochondrial dysfunction

Mitochondria are responsible for the intrahepatic β-oxidation of FFA as well as coordinating the tricarboxylic acid cycle, ATP synthesis through oxidative phosphorylation, and the generation of ROS (202). In patients with NAFLD, mitochondrial biogenesis, mitochondrial mass, and maximal respiratory rates are all upregulated to cope with the increased lipid burden compared with healthy controls (203). Over time, this “mitochondrial flexibility” can become exhausted, leading to uncoupling, enhanced ROS formation, and oxidative stress culminating in NASH and impaired hepatic insulin resistance (203–205). Mitochondrial-derived ROS have a central role in propagating hepatocyte injury by generating both lipid peroxidation products (206) and TNF-α, both of which induce further mitochondrial damage, permeability, and uncoupling (203). Enhanced hepatocellular sensitivity to TNF-α conferred by mitochondrial accumulation of cholesterol and subsequent glutathione depletion may then help the precipitation of NASH (207). Impaired mitophagy, the selective autophagic removal of damaged mitochondria, may also have a pathogenic role in NASH. Under physiological conditions mitophagy helps prevent cell death by reducing oxidative stress and preserving mitochondrial bioenergetics whereas this function is diminished by most features of the metabolic syndrome including obesity, insulin resistance, and dyslipidemia (208, 209). Autophagy has been shown to be significantly inhibited in patients with NASH as compared with steatosis alone and correlates with severity of liver disease and markers of oxidative stress (210).

Bile acids and the pathogenesis of NAFLD

There is an increasing recognition that bile acids are potent signaling molecules with important cell functions over and above their role to facilitate lipid absorption. Primary bile acids (cholic acid and chenodeoxycholic acid) are synthesized within the liver, secreted into the bile and delivered to the intestine where the actions of the gut microbiota generate secondary bile acids that are reabsorbed (>90%), largely in the small intestine into the portal circulation. A smaller proportion enters the systemic circulation.

Bile acids act via a variety of receptors including the farnesoid X receptor (FXR), takeda G-protein-coupled, pregnane X receptor, and the constitutive androstane receptor. Activation of these receptors is dependent upon the individual bile acid, and therefore the impact of alteration of the total bile acid pool depends upon both on changes in quantity and composition.

There is an extensive literature that has examined the impact of genetic manipulation of bile acid receptors (most commonly FXR using rodent models) in the development of NAFLD. Global FXR deletion appears to worsen metabolic phenotype (211), however, alterations in tissue-specific expression are important. Intestinal FXR activation is able to promote NAFLD (212), and activation of these receptors leads to fibroblast factor 19 secretion (FGF19) secretion, which, in addition to being a potent negative regulator of bile acid synthesis, has also been linked to the development of an adverse metabolic phenotype. FGF19 analogs are currently being trialed in the context of NASH (see Fibroblast growth factor 19 analogs section).

Data are conflicting with regards to total fasting and postprandial bile acid levels in obese patients and those with T2D; studies have shown unchanged, decreased, or elevated levels (recently reviewed in (213)). Many small studies have also specifically looked at bile acid levels in the context of NAFLD (using serum, plasma, fecal, and liver samples) (213). In the majority of studies, bile acid levels are elevated in NAFLD and there are some specific associations of individual bile acids with hepatic inflammation (eg, high levels of circulating glychololic acid were associated with lobular inflammation on liver biopsy) (214).

The mechanisms driving the changes in the bile acid pool are not fully determined. Changes in hepatic expression of key enzymes involved in bile acid synthesis have been described (214–216), and, importantly, the changes observed in the gut microbiome will have implications for the generation of secondary bile acids (see The gut microbiome section).

The relevance of bile acids and their receptor activation in the pathogenesis of NALFD is reflected in the fact that several pharmaceutical agents currently under investigation are targeting these signaling pathways, and it is likely that the first therapies licensed for NALFD will be specific FXR agonists (see section Farnesoid X receptor agonists section).

The gut microbiome

The gut microbiome is increasingly recognized in the pathogenesis and progression of a range of diseases including NAFLD/NASH, through the so-called gut–liver axis. Seminal findings initially demonstrated the existence of an “obese microbiota” with an increased capacity to harvest energy from the diet (217). Furthermore, this trait was transmissible, with germfree mice transplanted with feces from an obese donor accumulating more total body fat as compared to germfree mice receiving feces from a lean donor (217). A similar study of microbiota transplantation from human lean donors increased insulin sensitivity in obese individuals with metabolic syndrome (218). This ability of the microbiota to transmit a particular phenotype has also been demonstrated in rodent models of NASH with hepatic steatosis and inflammation being exacerbated in wild-type mice co-housed with mice with inflammasome-mediated liver injury (219). Developing gut dysbiosis (an unhealthy gut microbiome) is complex and highly variable between individuals with multiple environmental influences including diet, lifestyle, antibiotic exposure, and perinatal conditions (220). Due to difficulties in controlling for these variables, alongside small sample sizes and variation in laboratory techniques, microbiota composition studies in NAFLD/NASH are sparse and inconclusive (221). Nevertheless, several studies support an inverse correlation between Bacteroidetes and presence of NAFLD, NASH, and obesity (216, 217, 222). Bacteroidetes are relatively inefficient at dietary energy extraction and their reduction may allow more efficient species to dominate (eg, Firmiculates). Indeed, a 20% increase in Firmicutes and corresponding decrease in Bacteroidetes were associated with an approximate 150 kcal increase in energy nutrient absorption (223). Specific patterns of microbiome composition have also been implicated in fibrosis progression (224, 225) and the pathogenesis of HCC (226, 227). Not only do patients with NAFLD/NASH have compositional differences in microbiota compared to controls, but they also have a greater microbiota volume with the prevalence of small intestinal bacterial overgrowth (SIBO) ranging from 50% to 77% (228, 229). SIBO is associated with several of the pathogenic mechanisms thought to implicate the microbiota in NASH including increased gut permeability through loss of tight-junction integrity (230), enhanced TLR expression, and a greater burden of circulating proinflammatory cytokines (eg, IL-8 and TNF-α) (228, 229). Additional pathways by which dysbiosis and SIBO may adversely impact the liver in NASH include endogenous ethanol production (231) and disruption of both choline (232) and bile acid metabolism (233). The microbiome has also become a potential therapeutic target in the pharmacological treatment of NAFLD. Only data from small studies have been published, and studies are currently actively recruiting (see Modulation of the gut microbiome section).

Genetic predisposition to NAFLD and disease progression

Despite shared environmental risk factors, there remains considerable interindividual variation across the spectrum of NAFLD. For example, normal liver histology can be found even in those with morbid obesity (234), and only a minority of those with NAFLD will progress to NASH, fibrosis, and HCC. These phenotypic inconsistencies suggest a genetic contribution to the disease. Compared with the general population, first-degree relatives of patients with NAFLD have up to 3 times the risk of developing the disease themselves (235). Prospective studies have also demonstrated strong NAFLD concordance between monozygotic versus dizygotic twins and that heritability (the degree of phenotypic variation accounted for by genetics) of steatosis and fibrosis are both ~50% after controlling for age, sex, and ethnicity (236). Several common gene variants have emerged out of genome-wide association studies, which show an association with the development and progression of NAFLD. The most significant and reproducible of which is the PNPLA3 isoleucine to methionine substitution at position 148 (rs738409 C>G encoding for PNPLA3 I148M). Through screening ethnically diverse adults from the Dallas Heart Study, Romeo and colleagues were the first of many to describe a significant association between PNPLA3 I148M and hepatic fat without impacting on components of the metabolic syndrome (11). Since then the association has extended to histological severity, including NASH (odds ratio [OR] 3.24, 95% CI 2.79–3.76) (237), fibrosis (OR 3.11, 95% CI 2.66–3.65) (238), cirrhosis (OR 1.86, 95% CI 1.64–2.12) (239) and HCC (OR 2.32, 95% CI 1.76–3.06) (240). The mechanistic basis of PNPLA3’s relationship with NAFLD is yet to be fully characterized. While wild-type PNPLA3 seems to mediate TAG turnover, deletion of PNPLA3 in knockout mice confers no obvious hepatic phenotype (241). The loss-of-function 148M variant protein, however, manages to escape degradation and its accumulation appears to result in entrapment of TAG in lipid droplets (242) and increased fibrosis through HSC activation (243). Furthermore, reduced expression of PNPLA3 148M due to the co-presence of the 434K variant ameliorates liver damage (244) highlighting 148M downregulation as a potential treatment target in NAFLD (245). The second most important gene variant is the TM6SF2 guanine to adenine substitution at position 167 (rs58542926 E>K encoding for TM6SF2 E167K), which promotes steatosis by interrupting the enrichment of TAG in secreted VLDL (246) while simultaneously conferring protection against circulating dyslipidemia and CVD (247). Other variants with moderate effect size include loss-of-function polymorphisms in membrane bound O-acyltransferase domain containing 7 (MBOAT7) rs641738 and glucokinase regulator (GCKR) rs1260326. Downregulated MBOAT7 appears to mediate steatosis (248, 249), inflammation (249), fibrosis (248, 249), and carcinogenesis (250) through the toxic accumulation of polyunsaturated fatty acids following impaired incorporation into hepatocyte phospholipids while reduced GCKR causes hepatic fat accumulation through constitutive activation of hepatic glucose uptake and increased DNL (251, 252). Finally, it is important to note that NAFLD is a complex condition with a high degree of gene–environment interaction. For example, adiposity has been shown to significantly amplify the genetic risk conferred by common gene variants across the full spectrum of NAFLD (253). There is also evidence that dietary factors, particularly carbohydrate and fructose intake augment the impact of PNPLA3 I148M (254, 255).

More recently, polymorphisms within 17β-hydroxysteroid dehydrogenase type 13 (HSD17B13) have been identified that are associated with NAFLD (although not when adjusted for BMI), as well as with NASH, fibrosis, and cirrhosis (256, 257). HSD17B13 is a lipid associated droplet protein that appears to possess retinol dehydrogenase enzyme activity (258), although its true functional role is yet to be determined. Most recently, a loss of function variant has been shown to protect from the development of HCC in patients with alcohol-related liver disease (259).

Further work is needed to interrogate the precise molecular mechanisms underpinning the association between gene variants and progressive liver disease to identify novel and personalized therapeutic targets.

Clinical determinants of NAFLD progression and prognosis

There are many putative factors that may drive NAFLD progression to the more advanced stages of disease, many of which (eg, genotype, microbiome, mitochondrial function, immune response) are not easily or routinely assessed in clinical practice. As a result, we must look to natural history studies to help provide clinical, biochemical, and histological variables that can be used to help predict which patients will develop severe disease with poorer outcomes. Regarding clinical features, in a paired biopsy study McPherson et al. highlight the impact of insulin resistance on NAFL with 80% of those with fibrosis progression developing T2D at the time of follow-up biopsy compared with 25% of nonprogressors (260). Angulo et al. also showed that diabetes at baseline was associated with overall mortality in NAFLD alongside smoking, age, and absence of statin use (261). Data for biochemical predictors are relatively lacking although in patients with biopsy-proven NAFLD and compensated cirrhosis, lower levels of serum cholesterol, ALT, and platelets were shown to be independently associated with liver-related complications (development of varices, ascites, encephalopathy, and HCC) and higher aspartate aminotransferase (AST)/ALT ratio with overall mortality (262). A serial biopsy study of 320 NAFLD patients presented in abstract form showed that Fib-4 score (comprising ALT, AST, platelet count and age) was the only baseline nonhistological factor associated with fibrosis progression but had inadequate discriminative precision for use in clinical practice with an area under the receiver operator curve (AUROC) of 0.62 (263). Baseline histology provides good prognostic value in NAFLD. Systematic review and meta-analysis of paired biopsy studies has shown that a third of NAFLD patients will have fibrosis progression with overall annual fibrosis progression rate of 0.07 stages for NAFL and 0.14 stages for NASH, corresponding to one stage of fibrosis progression over a median of 14.3 and 7.1 years, respectively (264). Furthermore, several epidemiological studies have now confirmed the presence and severity of fibrosis as the most robust marker for future mortality and liver specific morbidity and have de-emphasized the presence of NASH, which adds very little prognostic value (261, 265, 266).

Clinical Consequences and Outcomes

Cardiovascular risk and mortality in NAFLD

Cardiovascular disease and NAFLD have shared risk factors including insulin resistance, T2D, obesity, hypertension, and dyslipidemia (267). Even after adjusting for conventional risk factors, NAFLD remains an independent risk factor for the development of CVD (27, 268, 269). In their global epidemiology study of NAFLD, Younossi et al. estimated the association between NAFLD and co-existing CVD risk factors including obesity (51%), T2D (23%), hyperlipidemia (69%), hypertension (39%), and metabolic syndrome (43%) (1).

Once CVD is established, there is also strong evidence suggesting that those patients with NAFLD have an increased severity of CVD burden (270). This may be related to the observation that those patients with NASH have additional risk and have a greater degree of endothelial dysfunction, impaired left and right ventricular function, and carotid artery disease increasing the risk of cerebrovascular disease and stroke (271). NASH, through its proinflammatory state, has systemic effects on multiple organ systems and appears to exacerbate insulin resistance and promote dyslipidemia, thus exacerbating the development of CVD (272).

Body fat distribution (over and above obesity and measurements of BMI) is also an important consideration; those patients with greater visceral adipose tissue, which has higher metabolic activity, are more likely to have NAFLD, and have a higher risk of developing cardio-metabolic complications (268).

A recent systematic review and meta-analysis has demonstrated that patients with NAFLD have a greater prevalence of coronary artery disease 7.5% versus 1.4% and stroke 0.9% versus 0.2% when compared to those without NAFLD. The risk of cardiovascular morbidity was 77% higher compared to controls (defined as those patients without NAFLD) and cardiovascular mortality was 46% higher (273). Similarly, a further meta-analysis has shown that the presence of NAFLD is associated with a 64% increase in the risk of both fatal and nonfatal cardiovascular events compared to patients without NAFLD (272). These figures are similar to a recent retrospective study of patients with both NAFLD and T2DM in which the risk of cardiovascular events was 70% higher, and all-cause mortality was 60% higher when compared to those without liver disease (274). In addition to these associations, patients with more advanced NAFLD have worse 10-year cardiovascular risk profiles (269) with poorer cardiovascular outcomes and are approximately 4 times more likely to suffer a fatal CVD event and twice as likely to have a nonfatal CVD event (265, 272).

NAFLD has a bidirectional relationship with T2D; it is a risk factor for the development of T2D with studies showing that those with NAFLD have a 1.5- to 2-fold increased risk of developing T2D. Similarly, those patients with T2D are more likely to develop NAFLD (267, 275). Indeed, the strongest predictor of fibrosis progression in those with NAFLD is the development of incident T2D (260). The presence of NASH is also an important factor with these individuals having an additional up to 3-fold risk of developing T2D (276). In turn, the presence of T2D worsens the course of liver fibrosis, is also more likely to lead to the development of NASH, is a predictor of the development of HCC, and is associated with increased all-cause and cardiovascular mortality (270). Finally, the combination of both T2D and NASH together leads to even worse outcomes, both with respect to CVD and liver-related outcomes.

Patients with NAFLD are also at increased risk of chronic kidney disease (CKD) where the prevalence has been shown to be as high as 55% in some studies even after adjustment for other causal factors including T2D and hypertension. Patients with NASH and advanced fibrosis both have a higher prevalence and incidence of CKD (267, 277, 278). Understanding the precise mechanisms that link CKD and NAFLD remains challenging, but several mechanisms have been postulated including the role of inflamed adipose tissue, circulating proinflammatory cytokines, and FFA that may comprise renal endothelial cell function and induce oxidative stress leading to end-organ damage (279).

All-cause morbidity and mortality in NAFLD

After CVD, cancer is the second clinical entity through which NAFLD exerts its morbidity and mortality, and this is manifested via both nonhepatic and hepatic malignancies. Cancers of the gastrointestinal tract are most closely associated with NAFLD; the strongest association being increased colorectal cancer. Studies have shown a 2- to 3-fold increase in the occurrence of adenomatous polyps or adenocarcinoma of the colon (280), and the risk of colorectal cancer is also influenced by the presence of NASH with this group having both an increased absolute risk as well as relatively higher grade of dysplasia. As such, mathematical modelling has suggested that targeted colorectal cancer screening in NAFLD patients may be cost-effective although further studies regarding patient acceptability and feasibility are required (281).

Though the evidence is less well established, other gastrointestinal cancers are associated with NAFLD including esophageal, stomach, and pancreatic cancer. For nongastrointestinal cancers, those associated with NAFLD include kidney and breast cancer as well as melanoma (280). Of additional note, the development of some cancers may also be gender specific, for example, a preponderance of breast cancer in females and of colorectal cancer in males (280, 282).

Liver-specific morbidity and mortality

Liver failure, including end-stage liver disease is the third clinical scenario through which NAFLD exerts its morbidity and mortality. Consequentially, accurate staging of liver disease is crucially important, allowing for both risk stratification and appropriate monitoring for disease progression and complication development. In this regard, liver biopsy remains the gold standard diagnostic tool allowing for both assessment of liver inflammation and liver fibrosis. Scoring systems used to grade liver biopsy tissue include the NAFLD Activity Score (NAS) to quantify histological features associated with a diagnosis of NASH (283) and the NASH CRN system used to stage fibrosis and to determine early fibrosis (F1–F2) from advanced (clinically significant) fibrosis (F3–F4), with F4 fibrosis being cirrhosis (284) (see Role of liver histology in diagnosis and Risk stratification sections).

Fibrosis progression in NAFLD has been estimated in 30% to 40% of patients with NAFLD (1, 264). Traditionally, it has been assumed that NAFL was an indolent disease with relatively slow progression compared to NASH, which progresses faster to advanced fibrosis. Indeed, it is NASH, which is the predominant form of NAFLD, that is thought to progress to advanced fibrotic disease with estimates suggesting that 10% to 15% of those with NASH progress on to cirrhosis with higher rates of decompensated liver disease and HCC development (27, 285). Furthermore, the presence of T2D in patients with NASH is an independent predictor of liver-related mortality, and, similarly, advanced fibrosis in those with NASH is an independent risk factor for predicting mortality (27).

Nevertheless, recent studies have shown that a significant proportion of those with simple NAFL do progress to advanced fibrosis. McPherson et al. reported that 44% of patients progressed to NASH, 37% progressed to fibrosis, and 22% to advanced fibrosis (260, 264, 286). Although this means that the majority of those with NAFLD have stable disease and do not progress, Singh et al. in their meta-analysis showed rate of progression in those with NAFL at 1 stage of fibrosis progression approximately every 14 years and double this rate for those with NASH at 1 fibrosis stage every 7 years (264).

Previously, it had been thought that only those with advanced liver fibrosis were at risk of increased all-cause mortality (265, 287). However, there is now convincing data indicating that even those patients with early (F1) fibrosis have an increased all-cause mortality compared to reference populations without fibrosis. Furthermore, liver-specific mortality rises exponentially with each increase of fibrosis stage (261, 288). Although still a keenly debated issue, these more recent data also suggest that fibrosis stage, irrespective of the presence of NASH, appears to be the most important determinant of outcomes in patients (265, 266, 289) and that the presence of other risk factors, including but not limited to age, T2D, and smoking appear to accelerate fibrosis progression and the potential development of liver failure. As expected, patients with NASH cirrhosis have worse liver-related and cardiovascular outcomes compared to those patients without cirrhosis; liver-related mortality is estimated at 24% in F3, compared with 59% in those with cirrhosis (F4) (288).

NAFLD and HCC

Between 2004 and 2009, there was an approximate 10% increase in NAFLD associated HCC incidence in the United States (1). NAFLD HCC patients were older, had a reduced life expectancy with more cardiovascular complications, and were more likely to die from their liver malignancy compared with other HCC etiologies. Similarly, in Europe, Dyson et al. (6) reported a 10-fold increase in HCC associated with NAFLD between 2000 and 2010 with patients again being older on average and experiencing greater cardio-metabolic complications, although survival in this cohort was similar to other etiologies of HCC (see Gender and age in NAFLD, NASH and HCC section).

The incidence of HCC is significantly higher in those with NASH compared to those without NASH (1). This is in agreement with data demonstrating that the incidence of HCC in patients with NAFLD cirrhosis is higher than those without cirrhosis. In patients with NAFLD cirrhosis, the incidence of HCC is estimated at 6.7% to 15% over a 5- to 10-year period (290). Studies estimate the incidence of HCC in patients without cirrhosis to be much lower; one large study showing a rate of 2.7% at 10 years, with the highest risk in older patients aged >55 years (291). Cardio-metabolic disease appears to be important in driving the development of HCC in patients both with and without cirrhosis (292, 293). A recent study has shown that patients with NAFLD and T2D were >19 times more likely to develop HCC and >6 times more likely to die from HCC compared to people without liver disease (274).

Current clinical consensus remains that screening for HCC in patients with NAFLD is only advocated for patients with established liver cirrhosis (290).

NAFLD and liver transplantation

NAFLD principally leads to the need for liver transplantation through hepatic decompensation as a consequence of cirrhosis, which affects 15% to 25% of those with NASH, and or the development of HCC in those with established cirrhosis (1).

The first liver transplant worldwide was performed in 1963 at a time when NAFLD/NASH was not a fully appreciated or recognized clinical entity. Indeed, NAFLD/NASH remained an insignificant etiology for liver transplantation until the mid-2000s. Since that time, its prominence and prevalence have risen to mirror the rise of the global obesity and T2D epidemic. NAFLD/NASH is currently the fastest growing etiology for liver failure. It is currently the second leading cause of liver transplantation worldwide and is predicted to be the commonest cause for liver transplantation within the next few years (294, 295). Indeed, the waiting list of entrants with NAFLD on the US transplant register increased by 170% between 2004 and 2013 (296). Importantly, the prevalence of “cryptogenic cirrhosis” was historically up to 30% in those with “all-cause cirrhosis,” of which a substantial number would have undergone liver transplantation. A large number of studies have subsequently demonstrated strong associations between the prevalence of metabolic conditions such as obesity and T2D in those with cryptogenic cirrhosis, which makes it likely that NAFLD accounted for a significant proportion, if not the majority etiology of these cases (297). As diagnostic advances in hepatology have improved, the proportion of cryptogenic cirrhosis diagnoses made has reduced to around 5% (296, 298).

The age of those undergoing liver transplantation for NASH has fallen significantly over the last 15 years. Analysis of transplant registrants in the United Network for Organ Sharing (1995–2015), which has profiled over 180 000 individuals, has shown a marked upward trend in NASH with and without HCC since 2005. Those aged between 35 and 55 years have shown the greatest increase in registration rates for liver transplantation (299). The authors termed this the “adipose wave effect” as it is line with data that show obesity and metabolic diseases continue to affect people at increasingly younger ages. Data from both the United States and Europe demonstrate the prevalence of childhood obesity to be approximately 20% or higher (300) with this trend continuing to increase. Children who are obese tend to remain obese into adulthood and this prolonged “exposure to obesity” is likely to be an important factor in the development of liver failure and HCC that require transplantation.

As the need and rate of transplantation for patients with NAFLD rises, the need for clear guidance as to how to manage and select suitable candidates has grown. The United Kingdom was one of the first countries to publish detailed guidance on NAFLD and liver transplantation in 2012, a significant proportion of which was based on consensus statements (301). As a large number of questions concerning transplantation in NAFLD and NASH continue to remain unanswered, such as natural history and management of co-morbidities in those undergoing transplantation, the International Liver Transplantation Society published the first International Consensus Statement on NAFLD and NASH in the setting of liver transplantation in early 2019 (302). This reviewed the available evidence with the aim of answering both important practical considerations and questions regarding NASH outcomes and natural history. Important aspects in this regard include risk assessment for patients listed for transplant, cardiovascular risk, and management of medical co-morbidities including cardio-metabolic diseases and obesity. Other considerations include the impact on the donor pool (296) and whether there is a need for a specific approach to donor steatosis in those donating organs.

Analysis of outcomes for patients with NASH undergoing liver transplantation is important as numbers of procedures performed continue to rise. Recently reported data from Europe has shown that survival for patients with NASH is similar to those transplanted for other liver diseases (302, 303). These results are consistent with other studies including data from the United States, some of which have suggested a survival advantage for those transplanted for NAFLD compared, for example, to HCV (295, 297, 304, 305). As may be expected, infection and sepsis were the commonest causes of death (~24%) in those with NASH without HCC, followed by cardio-metabolic complications (~5%), with higher age, severity of liver disease, and extremes of BMI (low or high) being predictors of death (303).

Data from a meta-analysis by Wang et al. confirmed that patients with NASH had greater risk of death from sepsis and cardiovascular complications (305), with additional evidence suggesting a lower risk of graft failure posttransplant compared to other liver failure etiologies (302). Additional studies have also demonstrated that patients with extremes of BMI are more at risk from postoperative complications (306), as well as increased cardio-metabolic complications pre- and posttransplantation compared with other non-NAFLD indications for transplantation.

There is evidence demonstrating that complications of cardio-metabolic diseases (eg, mild renal dysfunction around the time of transplant) are linked with adverse all-cause and cardiovascular outcomes (307). Immunosuppressive therapy is used in patients undergoing transplantation. Their role to prevent organ rejection is essential; however, it is well recognized that these agents also have adverse cardio-metabolic effects and potentially could contribute to the de novo development NAFLD or cardio-metabolic disease.

NAFLD recurrence posttransplantation is common and affects at least one-third of patients, although the true number is likely to be higher (296). De novo NAFLD posttransplantation is thought to affect around 18% to 33% of patients (308). This condition remains poorly understood, and longer-term studies are required to fully understand and characterize differences from recurrent NAFLD (eg, whether it is more aggressive in nature). Nevertheless, regular surveillance and aggressive management of NAFLD recurrence, de novo NAFLD, and cardio-metabolic disease is required posttransplantation. In this regard, work is ongoing to assess how patients with NAFLD in need of liver transplantation may be better managed in the future with additional targeted interventions (eg, bariatric surgery) where work is underway to establish the optimal timing and nature of such procedures (302).

Clinical Evaluation and Approach to Management

Establishing the diagnosis

Clinical

A diagnosis of NAFLD may be suspected in patients with mildly elevated liver biochemistry and/or echogenic liver on US scanning in the presence of one or more cardio-metabolic risk factors, such as (abdominal) obesity, T2D or impaired glucose tolerance, and hypertension and dyslipidemia and in the absence of other causes of hepatic steatosis and chronic liver diseases (115, 285). Typically, patients with NAFLD are asymptomatic, although fatigue and dull ache or discomfort in the right upper quadrant of the abdomen are not uncommon. More usually, abnormalities in imaging or liver biochemistry are discovered incidentally, such as the appearance of fatty liver on US scanning or computed tomography imaging performed because of abdominal symptoms or the finding of elevated liver biochemistry in patients undergoing blood test monitoring for other conditions (309).

A clinical history is required to exclude excessive alcohol consumption, which is defined by the EASL as ≥20 g/day for a woman or ≥30 g/day for a man (115) and medications that may cause hepatic steatosis, such as corticosteroids, amiodarone, and tamoxifen. Physical examination is required to determine the presence of abdominal obesity (waist circumference ≥94 cm for Europoid men and ≥80 cm for Europoid women), hypertension (≥130/85 mmHg or treated for hypertension), and any signs of chronic liver disease that might indicate advanced disease or portal hypertension.

Blood testing should include a full/complete blood count, liver biochemistry (including ALT), and AST; renal function; fasting blood glucose and/or glycosylated hemoglobin (HbA1c); blood lipid profile; serum ferritin and transferrin saturation; thyroid function; viral hepatitis B and C serology; liver autoantibodies (antinuclear antibody, antimitochondrial antibody, antismooth muscle antibody, antiliver kidney microsomal antibody); immunoglobulins (A, G, M); and α-1 antitrypsin levels. Other tests that may be required include serum ceruloplasmin and celiac serology in selected cases. Normal liver enzymes and normal liver US do not exclude NAFLD; indeed, of 103 patients with T2D and normal ALT, 51 had NAFLD as defined by proton MRS (1H MRS) (310).

Noninvasive assessments for diagnosis of hepatic steatosis

Imaging, particularly B-mode US, is often used for the diagnosis of NAFLD to look for evidence of hepatic steatosis (115). Typically this is a subjective assessment of increased hepatic echogenicity compared with the parenchyma of the adjacent right kidney. In a meta-analysis of studies including 4720 patients comparing US to liver histology, US was sensitive for the detection of moderate to severe steatosis (approximating to steatosis in >33% of hepatocytes) with a sensitivity of 84.8% (95% CI 79.5–88.9) and specificity of 93.6% (95% CI 87.2–97.0) (252, 311) but did not reliably detect steatosis of <20% (312). Ultrasound is unable to quantify hepatic steatosis, although the severity may be assessed and scored subjectively. Ultrasound is also used to exclude structural hepato-biliary co-morbidity such as gallstones and to look for evidence of macro-nodular cirrhosis and signs of portal hypertension including splenomegaly and the presence of ascites.

No single blood test can diagnose NAFLD, but there are a number of indices based on simple clinical assessments and blood tests that can predict the presence of significant hepatic steatosis. One such test is the Fatty Liver Index, which incorporates BMI with waist circumference, gamma-glutamyl transferase, and plasma TAG levels (313). Is was validated against hepatic steatosis as defined by ultrasonography, itself a surrogate reference marker and may be used in patients with metabolic risk factors to assess for the presence of NAFLD in the absence of US scanning (115). While other simple indices to diagnose hepatic steatosis have been evaluated against reference standards, including histology, 1H-MRS (314) and US, the diagnostic accuracy of three indices, the Fatty Liver Index, NAFLD Liver Fat Score, and Hepatic Steatosis Index were validated retrospectively with a cohort of patients with histologically defined NAFLD and the AUROC varied from 80% to 83% (315). The utility of such markers is unclear given the widespread availability of US, the high prevalence of steatosis in the general population (1), and that hepatic steatosis is frequently detected through other means. Such indices may be helpful in identifying cohorts of patients at high risk of NAFLD in whom further investigation would be indicated or in epidemiological studies.

Role of liver histology in diagnosis

Liver histology obtained by liver biopsy can be helpful diagnostically where the cause of abnormal liver biochemistry is unclear despite having performed a noninvasive liver disease screen. In such cases, in the absence of excessive alcohol intake or steatogenic drugs, the presence of intrahepatocellular lipid droplets in >5% of hepatocytes would be indicative of NAFL. NASH is diagnosed by global histological assessment. Characteristic features include the presence of lobular inflammation and ballooning of hepatocytes with or without fibrosis. Scoring systems for NASH including the NASH CRN score and the SAF (steatosis, activity, fibrosis) score have been derived and are widely used (284, 316). These scoring systems have been developed to standardize assessment of histological lesions seen in NASH and are not intended to be diagnostic, rather to provide a semiquantitative scoring system to improve agreement between histopathologists and assessment of change over time, predominantly in the context of clinical trials. Other histological features that are frequently seen in NASH include portal inflammatory infiltrate, Mallory–Denk bodies (Mallory’s hyaline), and megamitochondria (115). These features may also be seen in alcohol-related liver disease and alcoholic hepatitis so an accurate history of alcohol consumption is crucial to the interpretation of the histology.

Risk stratification

Severity of hepatic fibrosis is the strongest predictor of overall and liver-related outcome in patients with NAFLD (see Clinical determinants of NAFLD progression and prognosis section) (261, 265, 288). Establishing the severity of fibrosis in patients with NAFLD is therefore central to risk stratification. Diagnosis of NASH per se is not important for risk stratification (see Liver–specific morbidity and mortality section) but remains an endpoint in phase 3 clinical trials, given the current understanding of the pathogenesis of the condition (317, 318). Histological fibrosis staging remains the reference standard for gauging the severity of fibrosis in clinical practice and as a clinical trial endpoint. The NASH CRN system for fibrosis staging is widely used and employs a scale of 0 to 4 (284, 319). These stages correspond descriptively to mild fibrosis (stage 1a, 1b, 1c), significant fibrosis (presence of pericellular fibrosis; stage 2), advanced fibrosis (presence of bridging fibrosis; stage 3) and cirrhosis (bridging fibrosis with nodule formation; stage 4). The stages are descriptive of the pattern of fibrosis and the extent of fibrosis can vary substantially within each stage. Quantitation of fibrosis in histological liver samples has been developed by assessment of the collagen proportionate area (CPA) by digital image analysis of liver sections stained with Sirius red for collagen. In a cohort of 437 patients with NAFLD with retrospective follow-up over a median of 103 months, the CPA was an independent predictor of hepatic decompensation and, being a continuous variable, may be a useful surrogate marker for disease severity, particularly in the context of clinical trials (320).

Clinical, routine imaging, and laboratory measurements

Routine liver biochemical tests include measurement of ALT and/or AST, alkaline phosphatase, serum bilirubin, and gamma glutamyl transferase values. Elevations in ALT values are a common finding in patients with NAFLD, but they do not indicate the presence or absence of advanced disease. Importantly, a normal ALT value does not always indicate the absence of NAFLD. A study of 222 patients with NAFLD of whom 23% had elevated ALT and 77% had T2D found that the proportions of those with advanced fibrosis did not differ according to whether or not they had elevated ALT (321). There was a significantly lower proportion of people with NASH in the group with normal ALT compared to the group with elevated ALT making it more likely that NASH is present in the context of elevated ALT.

Noninvasive biomarkers of liver disease severity

Histological assessment through liver biopsy is associated with established risks to the patient (322), sampling variability, as well as inter- and intraobserver variation (323). Considerable progress has been made in the development of noninvasive markers of liver disease severity applied to NAFLD in the last 20 years with a number that are now routinely used in clinical practice. Markers can broadly be divided into “biological” markers (typically blood, but also other biological material) and “physical” markers (typically imaging-based technologies) (324). The performance of these modalities to accurately stage liver disease and therefore provide accurate risk stratification has been reviewed in detail (325) and a summary of their key features is provided next.

Biological markers.

Blood markers can be categorized into simple, usually inexpensive, blood tests, markers or indices based on routine biochemical and hematological tests or more complex, usually proprietary, panels of tests including intermediates of fibrogenesis. These have been reviewed recently and extensively (325–327).

Simple blood marker indices of disease severity are based on routine blood tests, and many have been derived from cohorts of patients with other chronic liver diseases, notably HCV. Central to these is the ratio of AST to ALT to which other components are appended. Examples include the AST/ALT ratio (328); the Fib-4 score, which also includes age and platelet count (329); the NAFLD Fibrosis Score (NFS), which includes platelet count; albumin concentration; BMI and the presence or absence of T2D or impaired fasting glucose (330); and the BARD score, which includes BMI and presence or absence of T2D or impaired fasting glucose (331). A number of these indices employ 2 thresholds or cut-off values, a lower value below which there is a high negative predictive value for the exclusion of advanced fibrosis and a higher value above which the specificity and positive predictive value is maximized (332). The choice of the thresholds to be used depends on how the test is to be employed in clinical practice. These indices were independently validated in a cohorts of patients with biopsy-proven NAFLD and found to have similar performance characteristics, notably for their high negative predictive value excluding advanced fibrosis/cirrhosis in 92% to 95% of cases (332). However, any clinical study validated against liver biopsy is prone to selection bias with a higher proportion of more severe disease stages than is represented in most primary care populations, which would be predicted to improve the negative predictive value of the test at the risk of including many false positive results (lower positive predictive value).

More complex proprietary panel markers have been developed using a number of biochemical tests in combination. The Enhanced Liver Fibrosis (ELF) score is designed to model matrix (including collagen) turnover, combining assays for hyaluronic acid, the amino terminal of procollagenase 3, and tissue inhibitor of metalloproteinase 1, and the diagnostic accuracy (AUROC) for advanced fibrosis in NAFLD was 93% (333). Other proprietary panel tests include FibroTest, which includes serum α2-macroglobulin, apolipoprotein A1, haptoglobin, total bilirubin, and gamma-glutamyl transpeptidase, adjusted for age and gender (334), which performed similarly to, but not better than, the NFS compared to histological fibrosis staging for the diagnostic accuracy across all fibrosis stages in an analysis of 574 patients (335).

Physical markers.

Physical markers are based on assessment of the physical properties of the liver parenchyma and include imaging-based technologies. Elastography is based on the measurement of vibration-induced shear waves through the liver tissue and include a means of generating a pulse or vibration and a means of detecting the velocity of the shear wave generated by the pulse. Broadly these are divided into US-based and magnetic resonance elastography (MRE). The most widely validated and used of these techniques is vibration-controlled transient elastography (TE), marketed as FibroScan®, which employs a probe placed against the skin between the ribs overlying the right lobe of the liver and generates a liver stiffness measurement (LSM) (336). As previously reported (324), there is published evidence in 25 studies and approximately 4000 patients with NAFLD using TE and a review of 10 recent studies comparing TE to histology, using either medium or large probes indicated AUCs for prediction of advanced fibrosis or cirrhosis (F0–F2 vs. F3–F4) of between 0.83 and 0.93. The sensitivity and specificity vary according to the cut-off values used, which may be set according to the disease state being sought, the tolerance of the local service for false positive and false negative results, and setting in which the technique is used. LSM by TE has a significant failure rate, either through an inability to obtain a reading or a technically inadequate reading. It is contraindicated in patients with cardiac pacemakers and implantable defibrillators and in pregnancy. A review of recent studies of TE in patients with NAFLD described inadequate readings (failure to obtain or unreliable) in between 5% and 23% of cases (324). However, a prospective, multicenter study of TE in more than 400 patients with NAFLD found that, when an appropriate probe size was used for each patient, only 3% of readings were invalid (337).

Other US techniques that allow elastography measurement include acoustic radiation force imaging and 2-dimensional shear wave elastography, which are applications integrated on US scanning machines. These techniques have increasingly been validated with similar performance characteristics, but on fewer patients than TE (324).

Magnetic resonance (MR) technologies have been applied in the context of NAFLD. While 1H-MRS (338) and the proton density fat fraction (PDFF) (339, 340) can sensitively quantify hepatic steatosis; however, their use in risk stratification of disease severity is limited, although a recent study suggested that increased liver fat as assessed by PDFF was associated with increased risk of disease progression (341). Magnetic resonance elastography applies the same concept of liver stiffness or elasticity measurement as US-based techniques and has been validated in patients with NAFLD (342) with high diagnostic accuracy for advanced fibrosis; pooled sensitivity was 86% and specificity was 91% in a recent comprehensive meta-analysis (343). MRE requires dedicated hardware and standardization of protocols between centers. A multi-parametric MR imaging (MRI) technique has been developed that incorporates an iron-corrected T1 (longitudinal relaxation time) sequence, which correlates inversely with disease severity in NAFLD (344). As it is a software analysis of predefined sequences, it has the benefits of not requiring specific hardware; is available across different scanner makes, models, and field strengths; and provides reproducible and standardized assessments. Magnetic resonance scanning applications are contraindicated in patients whose bodies contain ferrous material (penetrating eye injuries, shrapnel injuries, and certain implants) and cardiac devices such as pacemakers and implantable defibrillators and also may not be acceptable in patients with anxiety and claustrophobia.

The place of MR techniques in practice is, as yet, unclear. They are unlikely to form standard of care as first-line assessments in resource-constrained environments, but the detail and reproducibility of the data obtained lend them to use in clinical trials and longitudinal observational studies.

Putting risk stratification into practice

With the plethora of options now available for noninvasive risk stratification, attention has turned to development of pragmatic strategies for their use in practice. The low cost and performance characteristics of simple nonproprietary indices such as the Fib-4 and NFS make them an attractive first line for community-based risk stratification to enrich the population for whom more involved second-line tests can be employed.

A study of a UK primary care population of 1118 patients in Birmingham, UK, with elevated liver biochemistry looked at the prevalence of chronic liver diseases, and, after exclusion of other chronic liver diseases, a diagnosis of NAFLD was made in 295 (26.4%) (345). The NFS was used retrospectively to risk stratify these patients: 57% fell into the low risk category for advanced fibrosis, nearly 7.6% in the high-risk category, while the remaining 35% were classified as indeterminate, in whom additional risk stratification by other means was indicated. This study is notable in being one of the first to risk stratify patients in the community by noninvasive means and to establish the prevalence of advanced NAFLD in the population. However, the cause of abnormal liver tests was unclear in 45% despite use of a chronic liver disease screen of blood tests, and US scanning and risk stratification was not performed in these patients. Of the patients, 30.5% were obese, 19% had diabetes, and 41% were hypertensive, so a proportion are likely to have had NAFLD given that US scanning is not sensitive for mild hepatic steatosis, when the patient is obese or in advanced fibrosis where the degree of steatosis can decrease. Conversely, Koehler et al. conducted a population study of older people (mean age, 76 years) in the Netherlands. 3205 people underwent abdominal US scanning, of whom 2811 did not have risk factors for secondary steatosis (346). Of these, 35.1% had evidence of NAFLD, but, importantly, 88% had normal ALT value. This suggests that the prevalence of NAFLD and NAFLD with advanced fibrosis in the Birmingham study may have been underestimated.

In Nottingham, UK, TE was used in primary care for patients at risk of chronic liver disease (347). Risk factors for chronic liver disease were examined in the primary electronic medical record of over 20 000 patients. Those with hazardous alcohol intake or T2D were invited for TE and those without exclusion criteria (n = 2022) were invited for TE, of whom 919 underwent the test. Of these patients, 669 had normal liver stiffness (<8 kPa), and follow-up was arranged in primary care; 230 patients had elevated liver stiffness (≥8 kPa) and were seen in the hepatology clinic, of whom, 27 were diagnosed with cirrhosis (and of whom 16 had NAFLD as the underlying etiology, 3 had alcohol-related liver disease, and 8 had a combination of risk factors). A total of 23 patients with cirrhosis had been identified prior to the use of TE for those at increased risk, indicating a doubling of the pick-up of cirrhosis using this strategy. In another study, among 705 patients with T2D (348), the estimated prevalence of significant fibrosis, advanced fibrosis, and cirrhosis was estimated at 12.7%, 7.3%, and 2.1%, respectively, using predefined liver stiffness thresholds.

Despite extensive validation of simple markers in secondary care, a primary care based study in Edinburgh, UK, examined a cohort of 831 patients with T2D using US scanning followed by a number of simple blood markers (349). The correlation between different classes of biomarker was poor for the prediction of advanced fibrosis, including among those with NAFLD. However, in keeping with the good negative predictive value for advanced fibrosis seen in other studies, there was good agreement for the absence of advanced fibrosis among those patients with values in the lower 95 percentiles for each test. This study illustrates the potential pitfalls of rolling out risk stratification in primary care by extrapolating biomarker data from secondary care, particularly as the “true” disease state of those identified as “low-risk” in a primary care cohort cannot be established easily as liver biopsy cannot be justified in such patients. Nevertheless, on a population basis, these tests serve to enrich the population for those at risk of advanced liver disease and the negative predictive value should improve as the population prevalence of advanced disease decreases from secondary care to primary care populations.

The Birmingham study indicated that when using NFS, 35% of patients with NAFLD in the primary care cohort were classed neither in the high- or low-risk groups and so required further testing (345). This concept underlines the British Society of Gastroenterology’s guidance that, in patients with presumed NAFLD, a simple panel test (eg, Fib-4 or NFS), should in those with indeterminate values be followed by a second line test such as the ELF score or LSM (350). The lack of correlation between ELF or LSM and Fib-4 as determined by Morling et al. may thus be harnessed in sequential testing (349). This was tested retrospectively by Petta et al. in a cohort of 741 patients with biopsy-defined NAFLD (351). Simple biological panel markers including NFS, Fib-4, and LSM with FibroScan were tested against histology using published cut-off values. The markers were evaluated singly and in combination. Combination of Fib-4 or NFS with LSM in all cases led to very low false positive results (1.5%–2.1%) and false negative rates (3.8%–5.5%) for inclusion or exclusion of advanced fibrosis. There was, however, a large range of uncertainty (indeterminate) results comprising more than half of cases, leading to low accuracy in the region of 39% to 43% overall. If tests were employed serially (Fib-4 or NFS first, and then LSM only in those with indeterminate scores) the number of cases in the “uncertainty area” reduced to 6.4% to 8% and the overall accuracy increased to 76.0% to 77.8%, but at the expense of a higher number of false positive and false negative results. Care should be taken when extrapolating these results to a primary care setting as it was a retrospective study against liver biopsy adding selection bias and making it likely that the study population was enriched for advanced fibrosis and with less mild disease. Such a strategy in primary care where the prevalence of advanced disease is lower might be expected to have fewer false negative results, but more false positive results.

A sequential testing strategy was used by in London, UK, by the Camden and Islington care commissioning group and the Royal Free hospital and evaluated from March 2014 to May 2016 (352). The pathway consisted firstly of Fib-4 testing by general practitioners, then those considered low-risk (Fib-4 < 1.3) continued to be managed in primary care. Those with a high-risk Fib-4 test (Fib-4 > 3.25) were referred to the hepatology clinic, while those with an indeterminate result (Fib-4 > 1.3 < 3.25) were advised to undergo second line ELF testing. Of 1452 patients entering this pathway, 1022 fell into the low-risk category and 43 into the high-risk category, while 387 at indeterminate risk underwent ELF testing, of whom 155 had ELF <9.5 and were considered low risk and remained in primary care, while 232 had ELF >9.5 and referral to the liver clinic was recommended. The authors compared referrals before and after the initiative and from practices that had adopted the pathway with those that had not and reported a 5-fold increase in detection of advanced fibrosis, a 3-fold increase in detection of cirrhosis, and an 81% reduction in unnecessary referrals.

Assessment of cardio-metabolic risk

While the focus of this section has been on the risk assessment of the liver disease, CVD remains the chief cause of death among patients with NAFLD. We advocate the use of cardiovascular risk scores in the clinic setting for objective assessment of cardiovascular mortality risk so as to guide optimization of cardio-metabolic risk factors. Examples of cardiovascular risk scores include QRisk3 (353) with an online calculator (https://qrisk.org/three), the closely related JBS3 calculator (http://www.jbs3risk.com) and the Framingham risk score (354). QRisk2 and Framingham risk score were compared retrospectively in a cohort of patients with NAFLD and a further score was derived that included the mean platelet volume and was validated prospectively (355). AUROCs for the prediction of major acute cardiovascular events in the subsequent year were 0.83 for the NAFLD CV score, 0.73 for QRisk2, and 0.64 for Framingham.

Practical considerations in the assessment of NAFLD

Any strategy for risk stratification of NAFLD needs to include consideration of the practicalities of delivering the service beyond the absolute performance characteristics of the tests.

Confounding factors

Any noninvasive test or biomarker is likely to be affected by clinical confounding factors and awareness of these factors is related to how widely studied and validated a given test is. For example, well-known factors that increase liver stiffness and can lead to false positive results include the presence of severe hepatocellular inflammation, high right-sided heart pressures, and recent meals (356). However, when controlling for the degree of fibrosis in an analysis of patients with NAFLD, the degree of steatosis or inflammation was not found to influence the liver stiffness significantly (337). The components of the ELF test can be increased in extra-hepatic fibro-inflammatory conditions, while simple panels such as Fib-4 and NFS may be affected by any condition or concomitant medication that affects the platelet count (such as chronic inflammation, iron deficiency) or the ALT to AST ratio (alcohol use, medications, muscular damage, or inflammation). Future prospective studies should investigate the factors associated with false positive and false negative results, particularly in real-world and primary care cohorts where there are more likely to be co-morbidities present than in the well-characterized secondary care derivation and validation cohorts reported in the initial publications.

Resources, experience, acceptability

The diagnostic performance of risk stratification pathways should be considered in the context of the patient pathway, whereby, typically, patients are detected in primary care, undergo risk stratification and then those considered to be at higher risk undergo further evaluation and treatment in secondary or tertiary care (115). Simple blood markers (Fib-4, NFS) are cheap, available routinely, and can exclude advanced fibrosis with an excellent negative predictive value in the majority of patients. Using a cut-off value for Fib-4 of 1.3 in a UK primary care population, 70% of NAFLD patients were placed in the category at low risk for advanced fibrosis (352), while using a cut-off value for NFS of <1.455 in a UK primary care population, 57% of patients with NAFLD were placed in the category at low risk (345). The availability of resources in secondary care will determine the proportion of patients that can be assessed and treated in the clinic setting, the tolerance of false positive and false negative results and thus the thresholds used for risk stratification. As awareness of NAFLD increases, referrals to secondary care increase, which increases the need for effective risk stratification tools (352). In the absence of effective, discrete interventions in NAFLD, the cost-effectiveness of such pathways is unclear, although an analysis based on TE in primary care and numerous assumptions about disease progression and outcomes suggested that such a risk stratification pathway was likely to be cost-effective (357). The future availability of cost-effective treatments for NAFLD will also influence the need for risk stratification, and the thresholds employed as risk stratification will be required to determine eligibility for therapies.

Given the good performance characteristics of many tests for the detection or exclusion of advanced fibrosis, the decision to employ a given strategy will depend on local resources and the set-up of local services. For example, simple blood markers (Fib-4, NFS) are easily used in primary care. Second line risk stratification could be by complex panel markers (ELF, FibroTest), US-based elastography (TE, shear wave elastography) or MR techniques (MRE, multiparametric MRI). Complex blood markers can be sent from primary care, while TE is available in some primary care settings, but more often in secondary care, so geographical and travel considerations are also relevant to the choice of test. Typically, MR techniques for risk stratification are still confined to tertiary centers.

As most patients with NAFLD are asymptomatic, and there is widespread lack of awareness among healthcare professionals treating patients with diabetes (358), so the number of tests and visits required for risk stratification should be minimized to increase uptake. Of those assessed as high-risk by the Camden and Islington pathway, 45% were not seen in secondary care (352), but a sequential risk stratification pathway in Oxford, UK, used automated reflex testing of ELF in those with an indeterminate Fib-4, and resulted in 80% of those patients with high risk scores being seen in secondary care (J Cobbold, personal communication, June 2019). It may be postulated that comprehensive roll out and monitoring of a robust and acceptable system is more important than the minor differences in diagnostic accuracy of the particular technique(s) used, and recent UK guidance acknowledges the different preferences and experience of risk stratification markers among different clinicians by presenting alternative tools in the sequential testing strategy (350). Publication of further experience in clinical practice of different strategies is awaited.

Towards screening in high-risk groups?

NAFLD is usually asymptomatic and the prevalence of NAFLD is very high in certain groups (see Epidemiology of NAFLD section). Screening of high-risk groups is advocated (115), but the practicalities of such initiatives have not been established. Given that the majority of patients with NAFLD have normal ALT, ALT cannot be used for screening. If the prevalence of NAFLD in high-risk groups (including those with obesity and T2D) may exceed 90%, is US scanning or use of a steatosis-detection algorithm such as the FLI necessary? Alternatively, risk stratification for advanced fibrosis could be employed directly in high-risk groups. Examples include TE in patients with obesity, T2D (or hazardous alcohol intake) (347), and Fib-4 and TE in patients with T2D in secondary care (358, 359). Whether a systematic approach is warranted will rely on the benefit and cost-effectiveness of interventions in NAFLD and related conditions. Novel therapies and management approaches will need to be evaluated to establish their impact (360). Similarly, the performance characteristics of tests will vary by the prevalence of advanced disease in the populations concerned. Both of these issues need to be factored in before widespread screening can be advocated in resource-constrained environments.

A multidisciplinary approach to management

Given that NAFLD is typically diagnosed in patients with abnormal liver biochemistry and/or an echogenic liver on US scanning, gastroenterologists and hepatologists have been the specialists who have managed most incident cases in secondary and tertiary care. Yet, natural history studies demonstrate a diverse range of outcomes for patients diagnosed with NAFLD.

As discussed in the Clinical consequences and outcomes section, it is important to note that CVD is the principal cause of death affecting approximately 40% of patients, while malignant and nonmalignant liver diseases account for the deaths in less than 10% of patients (261, 265). Furthermore, more than 50% of patients attending secondary/tertiary care NAFLD clinics in the United Kingdom have T2D (360, 361). The role of the clinician caring for patients with NAFLD is to reduce the morbidity and mortality associated with the condition, so a “hepato-centric” management approach is insufficient and a multidisciplinary, holistic approach is required. While the personnel required in the multidisciplinary team may vary, the skill sets required are consistent (361):

  • Diagnosis and risk stratification of liver disease

  • Assessment and optimization of cardio-metabolic risk factors

  • Therapeutic optimization of T2D

  • Lifestyle and dietary assessment

  • Lifestyle education and intervention (including diet, exercise and smoking cessation)

  • Application of novel liver-directed therapies in NAFLD and recruitment into clinical trials

A suggested model for a multidisciplinary approach to management of NAFLD is shown in Fig. 2. Management of NAFLD in a multidisciplinary setting is widely advocated, (115, 362, 363), but there are very few data to demonstrate the utility of such an approach in routine clinical practice.

Schematic representation of the structure of a metabolic hepatology clinic (a). Adopting a multidisciplinary approach to nonalcoholic fatty liver disease has the potential to improve cardiovascular health and reduce liver stiffness (b), improve glycemic control in those patients with type 2 diabetes (c) and promote significant weight loss (d). From Moolla A, Motohashi K, Marjot T, Shard A, Ainsworth M, Gray A, Holman R, Pavlides M, Ryan JD, Tomlinson JW, Cobbold JF. A multidisciplinary approach to the management of NAFLD is associated with improvement in markers of liver and cardio-metabolic health. Frontline Gastroenterol. 2019;10:337–346. doi:10.1136/flgastro-2018-101155.
Figure 2.

Schematic representation of the structure of a metabolic hepatology clinic (a). Adopting a multidisciplinary approach to nonalcoholic fatty liver disease has the potential to improve cardiovascular health and reduce liver stiffness (b), improve glycemic control in those patients with type 2 diabetes (c) and promote significant weight loss (d). From Moolla A, Motohashi K, Marjot T, Shard A, Ainsworth M, Gray A, Holman R, Pavlides M, Ryan JD, Tomlinson JW, Cobbold JF. A multidisciplinary approach to the management of NAFLD is associated with improvement in markers of liver and cardio-metabolic health. Frontline Gastroenterol. 2019;10:337–346. doi:10.1136/flgastro-2018-101155.

Cobbold et al. analyzed an ethnically diverse cohort of 180 patients with NAFLD in London, UK, of whom 92 had T2D and who had been seen in a multidisciplinary NAFLD clinic and followed up for a median of 19.5 months (361). From baseline to latest visit, there were significant improvements in median ALT (18%), weight (3.5%), and total cholesterol (2.5%). Twenty-seven percent of patients achieved ≥5% weight loss, and among the patients with diabetes, there was a significant 1.9% reduction in HbA1c. Another study from Birmingham, UK, analyzed data from 65 patients with NAFLD seen in a multidisciplinary NAFLD clinic, of whom 32 had T2D. Over a short median follow-up of 98 days, significant improvements in weight, ALT, and total cholesterol were also seen, and 22% achieved ≥5% weight loss (359). Most recently, Moolla et al. reported experience of a multidisciplinary NAFLD clinic in Oxford, UK, where 165 patients were analyzed of whom 97 had T2D over a median period of 13 months (Fig. 2) (360). As in previous studies, there were significant improvements in ALT, weight, total cholesterol, and HbA1c, but also in this study significant improvements were demonstrated in liver stiffness by TE as a validated marker of fibrosis severity in NAFLD (difference in median values of 1.3 kPa, or 14%) and also the QRisk3 relative risk (difference in median values of 0.1, or 5%). This study highlighted the benefits, particularly in the management of patients with T2D (difference in median HbA1c of 4 mmol/mol, or 7%), with increased use of hypoglycemic agents associated with weight loss such as glucagon-like peptide 1 (GLP-1) agonists and reduced use of agents associated with weight gain such as insulins and sulfonylureas. There was evidence of cost-effectiveness of this approach, particularly in patients with poorly controlled T2D, but the analysis included a number of assumptions. A further UK center has reported in conference proceedings the change in cardio-metabolic endpoints associated with attendance in a multidisciplinary NAFLD clinic (364). Of the 120 patients included, 26 had diabetes at baseline while another 13 cases of diabetes were detected and treated. They reported a median reduction in QRisk-3 score of 5.2% over the follow-up period, similar to that reported by Moolla et al.

None of these studies included a control intervention group, and accordingly they were susceptible to bias. Moreover, the benefit of such approaches are not clear on an “intention to treat” basis. However, these studies serve as a benchmark as to what can be achieved in the clinic and to which novel interventions and methods of care delivery may be compared.

In conclusion, given the high prevalence of NAFLD, noninvasive risk stratification should be performed to determine those at increased risk of morbidity and mortality. The presence of advanced fibrosis in patients with NAFLD confers increased risk of all-cause mortality and numerous strategies are available to include or exclude this group. We advocate sequential testing strategies starting with a low-cost simple blood marker such as Fib-4 or NFS, followed by a second-line test such as TE or ELF to risk stratify those in the “gray zone” from the first test. Those at higher risk should be referred to a secondary care liver clinic. Consideration should be paid to the implementation of any risk stratification pathway to ensure engagement of stakeholders so as to maximize uptake and minimize dropout. A multidisciplinary NAFLD clinic to address liver-related and cardio-metabolic aspects of the condition, particularly diabetes, is recommended and data demonstrating the utility of such clinic models are emerging.

Lifestyle and Surgical Approaches to the Management of NAFLD

Commonly, the first-line recommendation in the management of patients with NAFLD is 7% to 10% weight loss, achieved through lifestyle changes (ie, changes in diet and exercise). While this may be a desirable goal, in practice, implementing and maintaining such a change can be challenging.

Exercise interventions

Physical activity is defined as any body movement generated by the contraction of skeletal muscles that raises energy expenditure above metabolic rate. It is characterized by modality, frequency, intensity, duration, and context of practice (365). Exercise is a subcategory of physical activity that is planned, structured, repetitive, and favors physical fitness maintenance or development (365). Thus, physical inactivity represents the nonachievement of physical activity guidelines; sedentary behaviors are waking behaviors characterized by an energy expenditure ≤1.5 metabolic equivalent while in sitting, reclining, or lying position, with screen time and sitting time typically being the 2 main indicators used to quantify sedentary behaviors (365). Physical activity and sedentary behaviors are not opposites, rather individuals are considered physically active when they achieve the guidelines but that does not preclude them from having a significant proportion of their day in sedentary behaviors (365). The effects of sedentary behaviors have been focused on recently with increasing evidence linking excessive sedentary behaviors and adverse health outcomes (365). Observational studies have suggested a strong positive association between sedentary behavior/decreased physical activity and prevalence of NAFLD (366). In support of these observations, Palve et al. reported that obese individuals classified as fit (based on their measured cardiorespiratory fitness through a peak oxygen uptake test) were at a significantly lower risk of NAFLD than participants classified as obese and unfit (367). As physical activity/exercise has the potential to lower NAFLD risk, a number of interventional studies have been undertaken using various physical activity/exercise regimens. Typically, studies have measured change in liver fat content assessed by imaging techniques such as MRI, 1H-MRS, or US. Other studies have used the less specific and sensitive change in blood liver enzymes as their outcome measure.

The effect of aerobic exercise on liver fat content was first highlight by Johnson et al. (368). They reported that 4-weeks of aerobic exercise reduced liver fat content by 21% (as measured by 1H-MRS) in the absence of significant weight loss in obese individuals. A number of other studies then followed including work by Sullivan et al. (369) who evaluated the weight loss-independent effects following the physical activity guidelines recommended by the United States Department of Health and Human Services on liver fat content and VLDL kinetics in sedentary, obese individuals with NAFLD (n = 18). Study participants were randomized to either 16 weeks of exercise training (45%–55% VO2 peak, 30–60 mins, 5 days/week; n = 12) or to continue usual behavior (control, n = 6). Overall they found exercising training significantly decreased liver fat content (on average, by 10.3%), did not change body weight or body fat, and did not alter hepatic VLDL-TG or VLDL-ApoB100 secretion rates. These findings have been corroborated by a systematic review and meta-analysis investigating the effects of aerobic exercise on liver fat content. They concluded that when comparing interventions that combined exercise with diet to diet alone (omitting studies that achieved substantial weight loss), 30 to 60 minutes of moderate to high-intensity exercise performed on 2 to 5 days/week for 1 to 10 months resulted in a small, but significant reduction in liver fat content despite minimal or no weight loss (370). Notably, this benefit was achieved at exercise levels below the current recommendations for obesity management.

Subsequently, a number of studies have been undertaken investigating the effect of different types of structured exercise on liver fat content. The efficacy of 8 weeks of commonly prescribed aerobic exercise on liver fat content in a group of inactive and overweight/obese adults, independent of dietary intervention or weight loss has also been investigated (371). Participants were randomized to 1 of 4 groups: (i) low- to moderate-intensity, high-volume aerobic exercise 60 minutes, 4 days/week (LO:HI, n = 12); (ii) high-intensity, low-volume aerobic exercise 45 minutes, 3 days/week (HI:LO, n = 12); (iii) low- to moderate-intensity, low-volume aerobic exercise 45 minutes, 3 days/week (LO:LO, n = 12); or (iv) placebo (n = 12). At the end of 8 weeks there was a significant change in group × time interaction (P = .04) in liver fat content which decreased by 2.38 ± 0.73% (mean ± standard error of the mean), 2.62 ± 1.0%, and 0.84 ± 0.47% in the HI:LO, LO:HI, and LO:LO groups respectively, and was independent of weight loss. These findings suggest that aerobic exercise reduces liver fat, and this can occur with either an emphasis on intensity over volume or volume over intensity, and even minimal engagement with exercise (LO:LO group) can lower liver fat in some, but not all, individuals (371). Therefore, it would seem that both the duration of exercise and its intensity are important, allowing individuals to have a more personalized regimen (potentially aiding long-term compliance) to achieve similar results. A recent systematic review and meta-analysis by Smart et al. found that participants in interventions consisting of exercise alone typically had a liver fat content ~3.5% lower than controls subjects (372). Taken together, the evidence suggests that aerobic exercise has a beneficial effect on liver fat content, even when weight loss is not achieved.

Resistance training

Although many studies have investigated the effects of aerobic exercise on liver fat content, population-based studies have also suggested that individuals who engage in resistance training have a lower liver fat content than individuals who don’t. Interventional studies support this suggestion; for example, Hallsworth et al. (373) was one of the first to report that when sedentary adults (n = 11) with NAFLD underwent 8 weeks of resistance exercise (performed 3 times/week consisting of 8 exercises with each session lasting 45–60 minutes) despite no change in body weight or fat mass, there was a relative reduction in liver fat content of 13%, alongside improvements in lipid oxidation, glucose control, and insulin resistance. Subsequently, a number of other studies investigating the effects of resistance training on liver fat content have been reported. For example, Zelber-Sagi et al. (374) randomized patients with NAFLD to either resistance training (3 times/week; n = 31) or the control group of stretching (n = 31) for 3 months. They found the hepato-renal US index to be significantly reduced in the resistance training group, compared to the control group: total, trunk and android fat significantly decreased while lean mass body mass significantly increased in the resistance training group; there was no change in the control group. Although the current data are limited, it would appear that as with aerobic exercise, resistance exercise has a beneficial effect on liver fat content even when weight loss is not achieved.

Comparing aerobic and resistance exercise

Recently, a systematic review compared the effects of aerobic versus resistance exercise on liver fat content and found that of the 18 studies identified as using aerobic exercise, liver fat content decreased in 17 of them while in 7 studies that were identified as resistance training, liver fat content was reported to decrease in 6 (375). In the study where liver fat did not change with aerobic training, a plausible explanation is that despite the intensity and duration being higher than other studies, the length of the study intervention was 7 consecutive days, which was notably shorter than other studies (375). Why liver fat content did not change in the resistance training study remains unclear and cannot be explained by differences in study participants, intervention intensity, duration or length (375). Taken together either aerobic or resistance training, when performed regularly over a periods of longer than 7 days, appear to be beneficial in decreasing liver fat content.

High-intensity interval training

In recent years, high-intensity interval training (HIIT) has become popular. HIIT involves high-intensity exercise divided into bouts and recover periods and has been suggested to provide comparable or greater benefits to cardiorespiratory fitness than continuous moderate-intensity training of longer duration (376). By using a modified HIIT program, that was considered realistic and safe for participants with low baseline fitness, Hallsworth et al. (377) demonstrated that modified HIIT training 3 times/week for 12 weeks resulted in a significant decrease (relative change ~26%) in liver fat content, along with significant decreases in body mass and fat mass, with no change in the control group. Twelve weeks of HIIT training has also been reported to significantly decrease liver fat content (39% relative reduction) in individuals with T2D (metformin and diet controlled) compared to those randomized to standard care (378).

Mechanisms linking exercise interventions and improvements in NAFLD

The mechanisms by which exercise reduces liver fat content appear to have received little attention. It has been demonstrated that a single bout of exercise stimulation increases adipose tissue blood flow and fat mobilization, resulting in the delivery of fatty acids to other organs, such as skeletal muscle (379). It remains unclear if exercise leads to fat mobilization with in the liver, but it could be speculated that fatty acids are liberated from stored TAG during exercise and then these fatty acids, along with those liberated from adipose tissue that enter the liver, are utilized in oxidation, rather than the esterification pathways (379). Indeed, the study by Hallsworth et al. (373) supports this concept as they found an increase in fat oxidation during a submaximal exercise test in individuals who did 8 weeks of resistance training compared to no change in the control group; there was no change in fasting fat oxidation in either group. By investigating VLDL kinetics before and after 16 weeks of moderate-intensity aerobic exercise, Shojaee-Moradie et al. (380) found that although exercise had no effect on VLDL production rates, the clearance rate of VLDL was significantly increased, and this may have contributed to the significant decrease in liver fat that was observed. More work investigating changes in hepatic fatty acid uptake, synthesis, and disposal is required to understand the effect exercise may have on these pathways.

Overall, there is clear evidence demonstrating that exercise, be it aerobic, resistance, or modified HITT training, has the ability to lower liver fat content to a similar degree, even when in the absence of weight loss, which suggests that exercise can be personalized to achieve the best results. For example, in individuals who are unable to undertake aerobic exercise (due to contraindications) then a program of resistance exercise would be of benefit. Equally combining exercise with calorie restriction to achieve weight loss may have an additive effect.

Calorie restriction

A consistent finding in the literature is that a significant reduction in body weight is associated with a significant reduction in liver fat content, highlighting the effectiveness of weight loss as prevention or treatment strategy for NAFLD. Therefore, weight reduction is recommended by all the scientific societies worldwide (381) and is suggested to be achieved through a calorie deficit of between 500 and 1600 kcal/day. With regards to the precise type of diet that should be consumed, there is no consensus in what is suggested; a variety of dietary regimens have been advocated including low to moderate fat and moderate to high carbohydrate intakes; low-carbohydrate, ketogenic diets; and high-protein Mediterranean diets. Currently, very low calorie diets (VLCDs) are not recommended by the Asia-Pacific guidelines as they are considered unsustainable (381), but there is growing interest in their use as a treatment strategy in metabolic disease.

A number of studies have been undertaken investigating the effect of a hypocaloric diet on liver fat content, which we have previously reviewed (382). Briefly, studies have been undertaken with calorie restriction (between 600–1500 kcal/day) for periods between 2 weeks and 2 years, using a variety of dietary regimes including high-carbohydrate, low-fat and low-carbohydrate; high-fat diets. Regardless of the dietary intervention, all studies found a significant decrease in liver fat content to varying degrees, which will be influenced by the phenotype of the participants (these studies are typically undertaken in individuals with a BMI > 25 kg/m2), the length and possibly composition of the dietary intervention and the actual calorie deficit achieved (382). With the hypocaloric diets that have been used, there are 2 different approaches: a relatively modest reduction in energy intake of 600 to 800 kcal/day for longer periods (6–11 months) (383–385) or a more severe reduction in energy intake so that the participant is consuming a VLCD of 450 to 800 kcal/day for period of 6 to 8 weeks (386–388). When compared, the changes in body mass and liver fat content are greater on the VLCD compared to the more modest changes; however, whether these changes are maintained long-term remains to be determined. In the study by Lim et al. (389) where individuals with T2D consumed a VLCD for 8 weeks, they identified rapid and significant changes in liver fat content. During the first week of the intervention, liver fat decreased by 30% and continued to decline to achieve a total reduction of 70% (along with a 15% reduction in body mass), with the liver fat content then being comparable to control subjects (liver fat content 2.9%) (389). These data demonstrate how rapidly a decrease in liver fat content can be achieved with aggressive calorie restriction. Following on from this observation, the Diabetes Remission Clinical Trial (DiRECT) utilized a VLCD for an average of 4 months and found that although some individuals achieved remission of their diabetes (“responders”) and some did not (“nonresponders”), both groups achieved comparable decreases in body mass of 16 kg and 13 kg, respectively, and in liver fat content of 13% and 12%, respectively (388). After 4 months of VLCD, participants in DiRECT went through a stepped food reintroduction program and entered a weight maintenance diet for a further 8 months. At 12 months, the decrease in liver fat content was maintained in the responders and had increased slightly in the nonresponders; the authors noted that the increase in liver fat content during the weight maintenance period was related to degree of weight gain (388).

Recently, Schwenger et al. (390) determined the effect of the prebariatric VLCD in 139 obese individuals (median BMI 47 kg/m2) and found that with a median duration on a VLCD of 3 weeks, the median weight loss was approximately 7 kg, and when liver histology was assessed the prevalence with normal histology, simple steatosis, and NASH was 24%, 81%, and 19%, respectively. As the authors noted, prevalence was lower than predicted, which may be due to the fact that individuals had followed the VLCD prior to having the liver biopsy (390). It is currently unclear as to why some individuals respond to a VLCD and others don’t and how these diets influence NASH. A calorie deficit is important for weight loss as well as decreasing liver fat content, and this is achieved rapidly using a VLCD; however, what remains to be determined is how well weight loss and the reduction in liver fat content are maintained longer term.

Aside from recommending calorie restriction, there is no consensus between the difference guidelines on the precise gold standard dietary intervention (381). For example, the EASL recommend low- to moderate-fat and moderate- to high-carbohydrate diets along with low-carbohydrate ketogenic diets and high-protein Mediterranean diets while the Italian Association for the Study of the Liver recommends a Mediterranean diet, and others make no specific recommendations (381). Taking the available evidence, findings are mixed as to whether in the context of a hypocaloric diet, macronutrient composition has an effect of the decrease in liver fat content achieved. For example, Haufe et al. (384) observed a similar reduction in liver fat content of approximate 45% when comparing diets that achieved a calorie deficit by reducing either fat or carbohydrate. Thus, it would seem that in the context of a hypocaloric diet, the macronutrient composition has little effect on the achieved decrease in liver fat content, and it is likely that the key factor to achieve a decrease in liver fat content is total calorie deficit. Although the mechanisms related to these changes are not well described, it would be reasonable to assume that changes are mediated through the following: decreased substrate for intrahepatic TAG production due to a reduction in fatty acids entering the liver either from diet or adipose tissue lipolysis, decreased intrahepatic DNL due to decreased substrate availability, which would lead to a repartitioning of intrahepatocellular fatty acids away from esterification and toward oxidation pathways (391). Overall, negative energy balance leading to weight loss is associated with a reduction in liver fat content, and this remains true irrespective of whether the calorie deficit is achieved by acute, VLCD interventions, or more modest reductions in calorie intake over an extended period.

Bariatric surgery

In 1978, bariatric surgery was defined as being metabolic surgery by Varco and Buchwald (392). Indeed, metabolic surgery is broadly the capability of surgery to contribute to proactive health care of which indeed bariatric surgery is only one aspect (392). A meta-analysis undertaken by Buchwald et al. in 2004 (393) found that bariatric operations were effective for weight loss in morbidly obese patients, and in the majority of patients with T2D, hyperlipidemia, hypertension, and OSA, there was either resolution or improvement.

Historically, the 6 dominant procedures in bariatric surgery are jejunoileal bypass, Roux-en-Y gastric bypass (RYGB), vertical banded gastroplasty, biliopancreatic diversion (and the related duodenal switch), adjustable gastric banding (AGB), and sleeve gastrectomy (SG), of which RYGB has been suggested to be the most effective treatment for obesity as it achieves greater weight loss than other procedures; SG and AGB are alternative surgical approaches that minimally alter upper gastrointestinal tract anatomy and reduce gastric volume but still achieve weight loss (394). Maciejewski et al. (395) compared the 10-year weight change in a large, multisite, clinical cohort of veterans who underwent RYGB compared with nonsurgical matches, along with the 4-year weight change in veterans who underwent RYGB, AGB, or SG and found patients who underwent RYGB lost 21% more of their baseline weight at 10 years than nonsurgical matches. At 4 years, the decrease from baseline weight was 16.9% greater in patients who underwent RYGB than patients undergoing AGB, who lost 9.7% more than patients undergoing SG.

Due to the dramatic effects on body weight and also the evidence for substantial improvement in obesity-related metabolic diseases, some, but not all, scientific societies worldwide recommend considering bariatric surgery an option for NAFLD in patients unresponsive to lifestyle changes (and pharmacotherapy) for reducing weight and complications (381). In a recent systematic review and meta-analysis by Lee et al. (396), they concluded that bariatric surgery resulted in biopsy-confirmed resolution of steatosis in 66% of patients, inflammation in 50% of patients, ballooning degeneration in 76% of patients, and fibrosis in 40% of patients. Although these findings clearly highlight the resolution of NAFLD in a large proportion of patients who undergo bariatric surgery, the authors also found that bariatric surgery was related to new or worsening features of NAFLD (eg, fibrosis) in 12% of patients (396). It remains unclear why some patients may have worsening of conditions, and it has been suggested that this may be related to the precise bariatric procedure that was performed and the extent of malnutrition and malabsorption experienced (397). Although there are a vast number of papers now investigating the effects of bariatric surgery on NAFLD and they all show, to varying degrees, that bariatric surgery leads to resolution in NAFLD in obese patients when measured histologically using imaging methods or blood markers (eg, fibrosis score, NAFLD activity score), it is difficult to determine if one surgical procedure has benefit over another due to a lack of randomized trials. Although a number of studies have compared the effects of different bariatric procedures, few have compared the different bariatric procedures on liver-specific parameters. One such study that attempted to do this compared liver parameters (using ultrasonography and blood markers) before and then 1 year after individuals had undergone either RYGB or SG (398). Weight loss at 1-year postoperatively was significantly greater in the RYGB group when compared to those who underwent SG. In addition, there were also differences in other metabolic markers including cholesterol and insulin. However, blood transaminase levels were significantly higher in the RYGB group compared the SG group. Clearly more studies robustly comparing the different bariatric procedures are required. Moreover, there are no randomized clinical trials in this area nor are there studies that have assessed patients at a similar point of weight loss, rather than time after surgery, so that the effect of the procedures cannot be robustly compared. It is clear that RYGB results in a more rapid weight loss, however, whether the longer-term effects on liver metabolism are similar needs to be determined. In addition, the mechanisms by which bariatric surgery alters liver fat content (and presumably metabolism) need to be elucidated. Although it is assumed that the change in liver fat after bariatric surgery is due to food restriction or malabsorption or a combination, it is plausible that changes in gut hormones or changes in other metabolic tissues, such as adipose tissue, also play a role (399).

Pharmacotherapy to Treat NAFLD

The therapeutic landscape in NAFLD is rapidly evolving. There are still no currently licensed therapies although several treatment modalities are currently in phase 3 development, and it is likely that within the next few years we will see the first therapies granted a license specifically for the treatment of NAFLD. A summary of agents currently being evaluated for the treatment of NAFLD are presented in Tables 1–3.

Table 1.

The Impact of Anti-Diabetic, Glucose Lowering Agents as Pharmacotherapy to Treat NAFLD

AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
MetforminMultiple mechanisms including AMPKinase activationSome but not all studies, have shown improvements in liver chemistry and hepatic steatosisMeta-analyses have shown no significant benefit on liver histologyRemains the first line pharmacotherapy for patients with T2D. Its use specifically for the treatment of NAFLD is not recommended.(400, 401)
PioglitazonePPARγ agonistConsistent improvements in liver chemistry and reductions in hepatic steatosis.Decreased steatosis and inflammation, with resolution of NASH. Some evidence to suggest that fibrosis may improveDespite the histological benefits, widespread use is currently limited, potentially by concerns over adverse effects including weight gain, bone fracture risk and fluid retention.(402–407)
LiraglutideGLP-1 analogImproved liver chemistry and decreased hepatic steatosis.Decreased steatosis, inflammation and resolution of NASH without worsening of fibrosisTo date, only data from small studies have been published and the relative contribution of weight loss and improvement in glycaemic control to the observed benefits in NASH are yet to be determined.(408–415)
ExenatideGLP-1 analogImprovements in liver chemistry and fatty liver indexn/a(416–418)
LixisenatideGLP-1 analogImprovement in ALT in obese and overweight individualsn/a(419)
SemaglutideGLP-1 analogn/an/aPhase 2 studies currently recruiting(420)
DulaglutideLong acting GLP-1 analogImproved liver chemistryHistological report of improvement in a single caseNo dedicated histological studies performed to date. Phase 2 studies are recruiting.(421, 422)
HM1522GLP-1/GIP/Glucagon triple agonistn/an/aEarly phase 2 studies currently recruiting.(423)
SitagliptinDPPIV inhibitorInconsistent results and some studies have failed to show improvements in liver chemistry or liver TAG content.Limited histological evidence of benefit with inconsistent results; improvement in NAS in some studies but not all.Overall the data do not suggest that there is significant benefit when used as a treatment for NAFLD.(412, 424, 425)
VildagliptinDPPIV inhibitorImproved liver chemistry and decreased liver TAG contentn/a(426)
EmpagliflozinSGLT2 inhibitorImprovements in liver chemistry and decreased hepatic TAG contentSome evidence for histological improvement in an open label study.(427–429)
CanagliflozinSGLT2 inhibitorSome evidence of improvements in liver chemistry and surrogate markers of fibrosis in very small numbers of patients. Reduction in liver TAG content and increased hepatic insulin sensitivityImprovement in inflammation and steatosis in a single biopsy study.Uncontrolled liver biopsy study, with very small numbers (n = 9) with no placebo or other comparator arm(430, 431)
IpragliflozinSGLT2 inhibitorSome evidence for improvement in liver chemistry with variable changes in markers of fibrosis.Single case report of histological improvement in steatosis, inflammation and ballooning(406, 432, 433)
DapagliflozinSGLT2 inhibitorImprovements in liver chemistry and reduction in liver TAG content. No impact on increased hepatic insulin sensitivityn/a(434, 435)
LicogliflozinDual SGLT1 and 2 inhibitorn/an/aEarly phase 2 studies currently recruiting.(436)
AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
MetforminMultiple mechanisms including AMPKinase activationSome but not all studies, have shown improvements in liver chemistry and hepatic steatosisMeta-analyses have shown no significant benefit on liver histologyRemains the first line pharmacotherapy for patients with T2D. Its use specifically for the treatment of NAFLD is not recommended.(400, 401)
PioglitazonePPARγ agonistConsistent improvements in liver chemistry and reductions in hepatic steatosis.Decreased steatosis and inflammation, with resolution of NASH. Some evidence to suggest that fibrosis may improveDespite the histological benefits, widespread use is currently limited, potentially by concerns over adverse effects including weight gain, bone fracture risk and fluid retention.(402–407)
LiraglutideGLP-1 analogImproved liver chemistry and decreased hepatic steatosis.Decreased steatosis, inflammation and resolution of NASH without worsening of fibrosisTo date, only data from small studies have been published and the relative contribution of weight loss and improvement in glycaemic control to the observed benefits in NASH are yet to be determined.(408–415)
ExenatideGLP-1 analogImprovements in liver chemistry and fatty liver indexn/a(416–418)
LixisenatideGLP-1 analogImprovement in ALT in obese and overweight individualsn/a(419)
SemaglutideGLP-1 analogn/an/aPhase 2 studies currently recruiting(420)
DulaglutideLong acting GLP-1 analogImproved liver chemistryHistological report of improvement in a single caseNo dedicated histological studies performed to date. Phase 2 studies are recruiting.(421, 422)
HM1522GLP-1/GIP/Glucagon triple agonistn/an/aEarly phase 2 studies currently recruiting.(423)
SitagliptinDPPIV inhibitorInconsistent results and some studies have failed to show improvements in liver chemistry or liver TAG content.Limited histological evidence of benefit with inconsistent results; improvement in NAS in some studies but not all.Overall the data do not suggest that there is significant benefit when used as a treatment for NAFLD.(412, 424, 425)
VildagliptinDPPIV inhibitorImproved liver chemistry and decreased liver TAG contentn/a(426)
EmpagliflozinSGLT2 inhibitorImprovements in liver chemistry and decreased hepatic TAG contentSome evidence for histological improvement in an open label study.(427–429)
CanagliflozinSGLT2 inhibitorSome evidence of improvements in liver chemistry and surrogate markers of fibrosis in very small numbers of patients. Reduction in liver TAG content and increased hepatic insulin sensitivityImprovement in inflammation and steatosis in a single biopsy study.Uncontrolled liver biopsy study, with very small numbers (n = 9) with no placebo or other comparator arm(430, 431)
IpragliflozinSGLT2 inhibitorSome evidence for improvement in liver chemistry with variable changes in markers of fibrosis.Single case report of histological improvement in steatosis, inflammation and ballooning(406, 432, 433)
DapagliflozinSGLT2 inhibitorImprovements in liver chemistry and reduction in liver TAG content. No impact on increased hepatic insulin sensitivityn/a(434, 435)
LicogliflozinDual SGLT1 and 2 inhibitorn/an/aEarly phase 2 studies currently recruiting.(436)

Abbreviations: ALT, alanine aminotransferase; AMPKinase, adenosine 5' adenosine monophosphate-activated protein kinase; DPPIV, dipeptidyl peptidase-4; GIP, gastric inhibitory polypeptide; GLP-1, glucagon‐like peptide‐1, NAFLD, nonalcoholic fatty liver disease; NASH, nonalcoholic steatohepatitis; PPARγ, peroxisome proliferator-activated receptor gamma; SGLT1, sodium-dependent glucose co-transporter 1; SGLT2, sodium-dependent glucose co-transporter 2; T2D, type 2 diabetes mellitus; TAG, triglyceride.

Table 1.

The Impact of Anti-Diabetic, Glucose Lowering Agents as Pharmacotherapy to Treat NAFLD

AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
MetforminMultiple mechanisms including AMPKinase activationSome but not all studies, have shown improvements in liver chemistry and hepatic steatosisMeta-analyses have shown no significant benefit on liver histologyRemains the first line pharmacotherapy for patients with T2D. Its use specifically for the treatment of NAFLD is not recommended.(400, 401)
PioglitazonePPARγ agonistConsistent improvements in liver chemistry and reductions in hepatic steatosis.Decreased steatosis and inflammation, with resolution of NASH. Some evidence to suggest that fibrosis may improveDespite the histological benefits, widespread use is currently limited, potentially by concerns over adverse effects including weight gain, bone fracture risk and fluid retention.(402–407)
LiraglutideGLP-1 analogImproved liver chemistry and decreased hepatic steatosis.Decreased steatosis, inflammation and resolution of NASH without worsening of fibrosisTo date, only data from small studies have been published and the relative contribution of weight loss and improvement in glycaemic control to the observed benefits in NASH are yet to be determined.(408–415)
ExenatideGLP-1 analogImprovements in liver chemistry and fatty liver indexn/a(416–418)
LixisenatideGLP-1 analogImprovement in ALT in obese and overweight individualsn/a(419)
SemaglutideGLP-1 analogn/an/aPhase 2 studies currently recruiting(420)
DulaglutideLong acting GLP-1 analogImproved liver chemistryHistological report of improvement in a single caseNo dedicated histological studies performed to date. Phase 2 studies are recruiting.(421, 422)
HM1522GLP-1/GIP/Glucagon triple agonistn/an/aEarly phase 2 studies currently recruiting.(423)
SitagliptinDPPIV inhibitorInconsistent results and some studies have failed to show improvements in liver chemistry or liver TAG content.Limited histological evidence of benefit with inconsistent results; improvement in NAS in some studies but not all.Overall the data do not suggest that there is significant benefit when used as a treatment for NAFLD.(412, 424, 425)
VildagliptinDPPIV inhibitorImproved liver chemistry and decreased liver TAG contentn/a(426)
EmpagliflozinSGLT2 inhibitorImprovements in liver chemistry and decreased hepatic TAG contentSome evidence for histological improvement in an open label study.(427–429)
CanagliflozinSGLT2 inhibitorSome evidence of improvements in liver chemistry and surrogate markers of fibrosis in very small numbers of patients. Reduction in liver TAG content and increased hepatic insulin sensitivityImprovement in inflammation and steatosis in a single biopsy study.Uncontrolled liver biopsy study, with very small numbers (n = 9) with no placebo or other comparator arm(430, 431)
IpragliflozinSGLT2 inhibitorSome evidence for improvement in liver chemistry with variable changes in markers of fibrosis.Single case report of histological improvement in steatosis, inflammation and ballooning(406, 432, 433)
DapagliflozinSGLT2 inhibitorImprovements in liver chemistry and reduction in liver TAG content. No impact on increased hepatic insulin sensitivityn/a(434, 435)
LicogliflozinDual SGLT1 and 2 inhibitorn/an/aEarly phase 2 studies currently recruiting.(436)
AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
MetforminMultiple mechanisms including AMPKinase activationSome but not all studies, have shown improvements in liver chemistry and hepatic steatosisMeta-analyses have shown no significant benefit on liver histologyRemains the first line pharmacotherapy for patients with T2D. Its use specifically for the treatment of NAFLD is not recommended.(400, 401)
PioglitazonePPARγ agonistConsistent improvements in liver chemistry and reductions in hepatic steatosis.Decreased steatosis and inflammation, with resolution of NASH. Some evidence to suggest that fibrosis may improveDespite the histological benefits, widespread use is currently limited, potentially by concerns over adverse effects including weight gain, bone fracture risk and fluid retention.(402–407)
LiraglutideGLP-1 analogImproved liver chemistry and decreased hepatic steatosis.Decreased steatosis, inflammation and resolution of NASH without worsening of fibrosisTo date, only data from small studies have been published and the relative contribution of weight loss and improvement in glycaemic control to the observed benefits in NASH are yet to be determined.(408–415)
ExenatideGLP-1 analogImprovements in liver chemistry and fatty liver indexn/a(416–418)
LixisenatideGLP-1 analogImprovement in ALT in obese and overweight individualsn/a(419)
SemaglutideGLP-1 analogn/an/aPhase 2 studies currently recruiting(420)
DulaglutideLong acting GLP-1 analogImproved liver chemistryHistological report of improvement in a single caseNo dedicated histological studies performed to date. Phase 2 studies are recruiting.(421, 422)
HM1522GLP-1/GIP/Glucagon triple agonistn/an/aEarly phase 2 studies currently recruiting.(423)
SitagliptinDPPIV inhibitorInconsistent results and some studies have failed to show improvements in liver chemistry or liver TAG content.Limited histological evidence of benefit with inconsistent results; improvement in NAS in some studies but not all.Overall the data do not suggest that there is significant benefit when used as a treatment for NAFLD.(412, 424, 425)
VildagliptinDPPIV inhibitorImproved liver chemistry and decreased liver TAG contentn/a(426)
EmpagliflozinSGLT2 inhibitorImprovements in liver chemistry and decreased hepatic TAG contentSome evidence for histological improvement in an open label study.(427–429)
CanagliflozinSGLT2 inhibitorSome evidence of improvements in liver chemistry and surrogate markers of fibrosis in very small numbers of patients. Reduction in liver TAG content and increased hepatic insulin sensitivityImprovement in inflammation and steatosis in a single biopsy study.Uncontrolled liver biopsy study, with very small numbers (n = 9) with no placebo or other comparator arm(430, 431)
IpragliflozinSGLT2 inhibitorSome evidence for improvement in liver chemistry with variable changes in markers of fibrosis.Single case report of histological improvement in steatosis, inflammation and ballooning(406, 432, 433)
DapagliflozinSGLT2 inhibitorImprovements in liver chemistry and reduction in liver TAG content. No impact on increased hepatic insulin sensitivityn/a(434, 435)
LicogliflozinDual SGLT1 and 2 inhibitorn/an/aEarly phase 2 studies currently recruiting.(436)

Abbreviations: ALT, alanine aminotransferase; AMPKinase, adenosine 5' adenosine monophosphate-activated protein kinase; DPPIV, dipeptidyl peptidase-4; GIP, gastric inhibitory polypeptide; GLP-1, glucagon‐like peptide‐1, NAFLD, nonalcoholic fatty liver disease; NASH, nonalcoholic steatohepatitis; PPARγ, peroxisome proliferator-activated receptor gamma; SGLT1, sodium-dependent glucose co-transporter 1; SGLT2, sodium-dependent glucose co-transporter 2; T2D, type 2 diabetes mellitus; TAG, triglyceride.

Table 2.

Liver-Targeted Therapies Currently in Development for the Treatment of NAFLD

AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
Liver-targeted therapies (anti-inflammatory and antifibrotic)
Obeticholic acidFXR agonistImprovements in liver chemistryResolution of NASH and improvements in fibrosisPhase 3 studies are currently recruiting (including patients with NASH cirrhosis).(437)
GS-9674FXR agonistn/an/aBiochemical improvements in patients with PSC, dedicated studies in patients with NAFLD are currently recruiting.(438)
Tropifexor (LJN452)Non-bile acid FXR agonistn/an/aPhase 2 studies currently recruiting(439)
ElafibranorDual PPAR α/δ agonistImprovements in liver chemistry as well as lipid profiles and glucose homeostasis in patients with T2D.Resolution of NASH without worsening in fibrosisPhase 3 studies are currently recruiting(440)
Lanifibranor (IVA337)Pan-PPAR agonistn/an/aPhase 2 studies are currently actively recruiting(441)
SaroglitazarDual PPARα/γ agonistn/an/aPhase 2 studies are currently actively recruiting(442)
CenicrivirocDual CCR2/5 antagonistNo changes in liver chemistry or steatosisPrimary end point for resolution of NASH was not met in the phase 2 studies, but improvements in fibrosis were observed.Some improvement in systemic markers of inflammation. A large phase 3 is currently recruiting(443)
Vitamin EMultiple mechanisms including anti-oxidant actionsInconsistent improvements in liver chemistry and steatosisResolution of NASH in some studies, but not all; no impact on fibrosis(432, 444)
GR-MD-02Glactin-3 protein inhibitionNo changes in liver chemistry in the early phase studies published to date.No significant change in NASH, although ballooning decreased. No change in fibrosisLiver biopsy studies performed in patients with NASH cirrhosis.(445, 446)
NGM282FGF19 analogImprovements in liver chemistry and steatosisEarly phase 2 study has demonstrated improvements in steatosis, inflammation and fibrosisA larger phase 2 study is actively recruiting.(447, 448)
LY2405319FGF21 analogn/an/aEvidence for lipid lowering in patients with type 2 diabetes(449)
Pegbelfermin (BMS-986036)FGF21 analogDecreased hepatic steatosisn/aA series of phase 2 studies are currently recruiting.(450)
Resmetirom, MGL-3196Thyroid Hormone Receptor-β agonistDecreased hepatic steatosisResolution of NASH and decreased steatosisA phase 3 study is actively recruiting(50, 451)
SelonsertibApoptosis signal-regulating kinase 1 inhibitorImprovements in liver chemistry.Improved fibrosis at the highest doses.(452)
Simtuzumablysl oxidase-like-2 inhibitionNo significant changes in liver biochemistry.No improvement in NASH or fibrosis.No benefit on histological analysis or on clinical outcomes(453)
PentoxifyllineMultiple mechanisms including anti-inflammatory actions and reduction in free oxygen radial generationNo change in liver chemistryEvidence for improvement in NASH and fibrosis(454, 455)
EmricasanPan-caspase inhibitionImproved liver chemistryn/aShort duration early phase studies only have been reported(456, 457)
SGM-1019Inflammasome inhibitionn/an/aSafe and well tolerated in a phase 1 study. A phase 2 study is actively recruiting(458, 459)
JKB-121Toll-like receptor 4 antagonismNo impact on hepatic steatosisn/aRelatively high adverse event rate leading to drug withdrawal(460)
IMM-124ELimitation of endotoxin exposuren/an/aPhase 2 study completed, but results not yet reported(461)
AmlexanoxIKKb and TANK-binding kinase 1 inhibitionSome evidence for decreased hepatic steatosis in patients with diabetesn/a(462)
Tipelukast (MN-001)Phosphodiesterase and 5-lipoxygenase inhibition and leukotriene receptor antagonismn/an/aPhase 2 studies are currently recruiting(463)
DS1025-lipoxygenase inhibitionn/an/aPhase 2 studies are currently recruiting(464, 465)
BI 1467335Vascular adhesion protein-1 inhibitionn/an/aEarly phase 2 studies are currently recruiting(466)
MSDC-0602KMitochondrial pyruvate carrier inhibitionEvidence for improved metabolic as well as surrogates of hepatic inflammation and fibrosis on interim analysisn/aLarge phase 2 study (EMMINENCE) currently recruiting(467, 468)
Hepatic lipid metabolism
GS-0976Acetyl-Coenzyme A Carboxylase InhibitionImproved liver chemistry and hepatic steatosisn/a(469, 470)
AramcholStearoyl coenzyme A desaturase 1 inhibitionNo change in liver chemistry, but decreased hepatic steatosisn/aPhase 2b completed although not reported.(471)
Gut microbiome
ProbioticsAlteration in gut microbiomeImprovements in liver chemistry and steatosisn/a(472)
SolithromycinMacrolide antibioticn/aImproved NAS in small proof-of-concept study(473)
Fecal microbiome transplantationn/an/aSeveral phase 2 studies currently recruiting(474–476)
AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
Liver-targeted therapies (anti-inflammatory and antifibrotic)
Obeticholic acidFXR agonistImprovements in liver chemistryResolution of NASH and improvements in fibrosisPhase 3 studies are currently recruiting (including patients with NASH cirrhosis).(437)
GS-9674FXR agonistn/an/aBiochemical improvements in patients with PSC, dedicated studies in patients with NAFLD are currently recruiting.(438)
Tropifexor (LJN452)Non-bile acid FXR agonistn/an/aPhase 2 studies currently recruiting(439)
ElafibranorDual PPAR α/δ agonistImprovements in liver chemistry as well as lipid profiles and glucose homeostasis in patients with T2D.Resolution of NASH without worsening in fibrosisPhase 3 studies are currently recruiting(440)
Lanifibranor (IVA337)Pan-PPAR agonistn/an/aPhase 2 studies are currently actively recruiting(441)
SaroglitazarDual PPARα/γ agonistn/an/aPhase 2 studies are currently actively recruiting(442)
CenicrivirocDual CCR2/5 antagonistNo changes in liver chemistry or steatosisPrimary end point for resolution of NASH was not met in the phase 2 studies, but improvements in fibrosis were observed.Some improvement in systemic markers of inflammation. A large phase 3 is currently recruiting(443)
Vitamin EMultiple mechanisms including anti-oxidant actionsInconsistent improvements in liver chemistry and steatosisResolution of NASH in some studies, but not all; no impact on fibrosis(432, 444)
GR-MD-02Glactin-3 protein inhibitionNo changes in liver chemistry in the early phase studies published to date.No significant change in NASH, although ballooning decreased. No change in fibrosisLiver biopsy studies performed in patients with NASH cirrhosis.(445, 446)
NGM282FGF19 analogImprovements in liver chemistry and steatosisEarly phase 2 study has demonstrated improvements in steatosis, inflammation and fibrosisA larger phase 2 study is actively recruiting.(447, 448)
LY2405319FGF21 analogn/an/aEvidence for lipid lowering in patients with type 2 diabetes(449)
Pegbelfermin (BMS-986036)FGF21 analogDecreased hepatic steatosisn/aA series of phase 2 studies are currently recruiting.(450)
Resmetirom, MGL-3196Thyroid Hormone Receptor-β agonistDecreased hepatic steatosisResolution of NASH and decreased steatosisA phase 3 study is actively recruiting(50, 451)
SelonsertibApoptosis signal-regulating kinase 1 inhibitorImprovements in liver chemistry.Improved fibrosis at the highest doses.(452)
Simtuzumablysl oxidase-like-2 inhibitionNo significant changes in liver biochemistry.No improvement in NASH or fibrosis.No benefit on histological analysis or on clinical outcomes(453)
PentoxifyllineMultiple mechanisms including anti-inflammatory actions and reduction in free oxygen radial generationNo change in liver chemistryEvidence for improvement in NASH and fibrosis(454, 455)
EmricasanPan-caspase inhibitionImproved liver chemistryn/aShort duration early phase studies only have been reported(456, 457)
SGM-1019Inflammasome inhibitionn/an/aSafe and well tolerated in a phase 1 study. A phase 2 study is actively recruiting(458, 459)
JKB-121Toll-like receptor 4 antagonismNo impact on hepatic steatosisn/aRelatively high adverse event rate leading to drug withdrawal(460)
IMM-124ELimitation of endotoxin exposuren/an/aPhase 2 study completed, but results not yet reported(461)
AmlexanoxIKKb and TANK-binding kinase 1 inhibitionSome evidence for decreased hepatic steatosis in patients with diabetesn/a(462)
Tipelukast (MN-001)Phosphodiesterase and 5-lipoxygenase inhibition and leukotriene receptor antagonismn/an/aPhase 2 studies are currently recruiting(463)
DS1025-lipoxygenase inhibitionn/an/aPhase 2 studies are currently recruiting(464, 465)
BI 1467335Vascular adhesion protein-1 inhibitionn/an/aEarly phase 2 studies are currently recruiting(466)
MSDC-0602KMitochondrial pyruvate carrier inhibitionEvidence for improved metabolic as well as surrogates of hepatic inflammation and fibrosis on interim analysisn/aLarge phase 2 study (EMMINENCE) currently recruiting(467, 468)
Hepatic lipid metabolism
GS-0976Acetyl-Coenzyme A Carboxylase InhibitionImproved liver chemistry and hepatic steatosisn/a(469, 470)
AramcholStearoyl coenzyme A desaturase 1 inhibitionNo change in liver chemistry, but decreased hepatic steatosisn/aPhase 2b completed although not reported.(471)
Gut microbiome
ProbioticsAlteration in gut microbiomeImprovements in liver chemistry and steatosisn/a(472)
SolithromycinMacrolide antibioticn/aImproved NAS in small proof-of-concept study(473)
Fecal microbiome transplantationn/an/aSeveral phase 2 studies currently recruiting(474–476)

Abbreviations: FGF19, fibroblast growth factor 19; FGF21, fibroblast growth factor 21; FXR, farnesoid X receptor; NAFLD, nonalcoholic fatty liver disease; NASH, nonalcoholic steatohepatitis; PPARα/δ, peroxisome proliferator-activated receptor alpha/delta, PPARα/γ, peroxisome proliferator-activated receptor alpha/gamma; PSC, primary sclerosing cholangitis.

Table 2.

Liver-Targeted Therapies Currently in Development for the Treatment of NAFLD

AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
Liver-targeted therapies (anti-inflammatory and antifibrotic)
Obeticholic acidFXR agonistImprovements in liver chemistryResolution of NASH and improvements in fibrosisPhase 3 studies are currently recruiting (including patients with NASH cirrhosis).(437)
GS-9674FXR agonistn/an/aBiochemical improvements in patients with PSC, dedicated studies in patients with NAFLD are currently recruiting.(438)
Tropifexor (LJN452)Non-bile acid FXR agonistn/an/aPhase 2 studies currently recruiting(439)
ElafibranorDual PPAR α/δ agonistImprovements in liver chemistry as well as lipid profiles and glucose homeostasis in patients with T2D.Resolution of NASH without worsening in fibrosisPhase 3 studies are currently recruiting(440)
Lanifibranor (IVA337)Pan-PPAR agonistn/an/aPhase 2 studies are currently actively recruiting(441)
SaroglitazarDual PPARα/γ agonistn/an/aPhase 2 studies are currently actively recruiting(442)
CenicrivirocDual CCR2/5 antagonistNo changes in liver chemistry or steatosisPrimary end point for resolution of NASH was not met in the phase 2 studies, but improvements in fibrosis were observed.Some improvement in systemic markers of inflammation. A large phase 3 is currently recruiting(443)
Vitamin EMultiple mechanisms including anti-oxidant actionsInconsistent improvements in liver chemistry and steatosisResolution of NASH in some studies, but not all; no impact on fibrosis(432, 444)
GR-MD-02Glactin-3 protein inhibitionNo changes in liver chemistry in the early phase studies published to date.No significant change in NASH, although ballooning decreased. No change in fibrosisLiver biopsy studies performed in patients with NASH cirrhosis.(445, 446)
NGM282FGF19 analogImprovements in liver chemistry and steatosisEarly phase 2 study has demonstrated improvements in steatosis, inflammation and fibrosisA larger phase 2 study is actively recruiting.(447, 448)
LY2405319FGF21 analogn/an/aEvidence for lipid lowering in patients with type 2 diabetes(449)
Pegbelfermin (BMS-986036)FGF21 analogDecreased hepatic steatosisn/aA series of phase 2 studies are currently recruiting.(450)
Resmetirom, MGL-3196Thyroid Hormone Receptor-β agonistDecreased hepatic steatosisResolution of NASH and decreased steatosisA phase 3 study is actively recruiting(50, 451)
SelonsertibApoptosis signal-regulating kinase 1 inhibitorImprovements in liver chemistry.Improved fibrosis at the highest doses.(452)
Simtuzumablysl oxidase-like-2 inhibitionNo significant changes in liver biochemistry.No improvement in NASH or fibrosis.No benefit on histological analysis or on clinical outcomes(453)
PentoxifyllineMultiple mechanisms including anti-inflammatory actions and reduction in free oxygen radial generationNo change in liver chemistryEvidence for improvement in NASH and fibrosis(454, 455)
EmricasanPan-caspase inhibitionImproved liver chemistryn/aShort duration early phase studies only have been reported(456, 457)
SGM-1019Inflammasome inhibitionn/an/aSafe and well tolerated in a phase 1 study. A phase 2 study is actively recruiting(458, 459)
JKB-121Toll-like receptor 4 antagonismNo impact on hepatic steatosisn/aRelatively high adverse event rate leading to drug withdrawal(460)
IMM-124ELimitation of endotoxin exposuren/an/aPhase 2 study completed, but results not yet reported(461)
AmlexanoxIKKb and TANK-binding kinase 1 inhibitionSome evidence for decreased hepatic steatosis in patients with diabetesn/a(462)
Tipelukast (MN-001)Phosphodiesterase and 5-lipoxygenase inhibition and leukotriene receptor antagonismn/an/aPhase 2 studies are currently recruiting(463)
DS1025-lipoxygenase inhibitionn/an/aPhase 2 studies are currently recruiting(464, 465)
BI 1467335Vascular adhesion protein-1 inhibitionn/an/aEarly phase 2 studies are currently recruiting(466)
MSDC-0602KMitochondrial pyruvate carrier inhibitionEvidence for improved metabolic as well as surrogates of hepatic inflammation and fibrosis on interim analysisn/aLarge phase 2 study (EMMINENCE) currently recruiting(467, 468)
Hepatic lipid metabolism
GS-0976Acetyl-Coenzyme A Carboxylase InhibitionImproved liver chemistry and hepatic steatosisn/a(469, 470)
AramcholStearoyl coenzyme A desaturase 1 inhibitionNo change in liver chemistry, but decreased hepatic steatosisn/aPhase 2b completed although not reported.(471)
Gut microbiome
ProbioticsAlteration in gut microbiomeImprovements in liver chemistry and steatosisn/a(472)
SolithromycinMacrolide antibioticn/aImproved NAS in small proof-of-concept study(473)
Fecal microbiome transplantationn/an/aSeveral phase 2 studies currently recruiting(474–476)
AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
Liver-targeted therapies (anti-inflammatory and antifibrotic)
Obeticholic acidFXR agonistImprovements in liver chemistryResolution of NASH and improvements in fibrosisPhase 3 studies are currently recruiting (including patients with NASH cirrhosis).(437)
GS-9674FXR agonistn/an/aBiochemical improvements in patients with PSC, dedicated studies in patients with NAFLD are currently recruiting.(438)
Tropifexor (LJN452)Non-bile acid FXR agonistn/an/aPhase 2 studies currently recruiting(439)
ElafibranorDual PPAR α/δ agonistImprovements in liver chemistry as well as lipid profiles and glucose homeostasis in patients with T2D.Resolution of NASH without worsening in fibrosisPhase 3 studies are currently recruiting(440)
Lanifibranor (IVA337)Pan-PPAR agonistn/an/aPhase 2 studies are currently actively recruiting(441)
SaroglitazarDual PPARα/γ agonistn/an/aPhase 2 studies are currently actively recruiting(442)
CenicrivirocDual CCR2/5 antagonistNo changes in liver chemistry or steatosisPrimary end point for resolution of NASH was not met in the phase 2 studies, but improvements in fibrosis were observed.Some improvement in systemic markers of inflammation. A large phase 3 is currently recruiting(443)
Vitamin EMultiple mechanisms including anti-oxidant actionsInconsistent improvements in liver chemistry and steatosisResolution of NASH in some studies, but not all; no impact on fibrosis(432, 444)
GR-MD-02Glactin-3 protein inhibitionNo changes in liver chemistry in the early phase studies published to date.No significant change in NASH, although ballooning decreased. No change in fibrosisLiver biopsy studies performed in patients with NASH cirrhosis.(445, 446)
NGM282FGF19 analogImprovements in liver chemistry and steatosisEarly phase 2 study has demonstrated improvements in steatosis, inflammation and fibrosisA larger phase 2 study is actively recruiting.(447, 448)
LY2405319FGF21 analogn/an/aEvidence for lipid lowering in patients with type 2 diabetes(449)
Pegbelfermin (BMS-986036)FGF21 analogDecreased hepatic steatosisn/aA series of phase 2 studies are currently recruiting.(450)
Resmetirom, MGL-3196Thyroid Hormone Receptor-β agonistDecreased hepatic steatosisResolution of NASH and decreased steatosisA phase 3 study is actively recruiting(50, 451)
SelonsertibApoptosis signal-regulating kinase 1 inhibitorImprovements in liver chemistry.Improved fibrosis at the highest doses.(452)
Simtuzumablysl oxidase-like-2 inhibitionNo significant changes in liver biochemistry.No improvement in NASH or fibrosis.No benefit on histological analysis or on clinical outcomes(453)
PentoxifyllineMultiple mechanisms including anti-inflammatory actions and reduction in free oxygen radial generationNo change in liver chemistryEvidence for improvement in NASH and fibrosis(454, 455)
EmricasanPan-caspase inhibitionImproved liver chemistryn/aShort duration early phase studies only have been reported(456, 457)
SGM-1019Inflammasome inhibitionn/an/aSafe and well tolerated in a phase 1 study. A phase 2 study is actively recruiting(458, 459)
JKB-121Toll-like receptor 4 antagonismNo impact on hepatic steatosisn/aRelatively high adverse event rate leading to drug withdrawal(460)
IMM-124ELimitation of endotoxin exposuren/an/aPhase 2 study completed, but results not yet reported(461)
AmlexanoxIKKb and TANK-binding kinase 1 inhibitionSome evidence for decreased hepatic steatosis in patients with diabetesn/a(462)
Tipelukast (MN-001)Phosphodiesterase and 5-lipoxygenase inhibition and leukotriene receptor antagonismn/an/aPhase 2 studies are currently recruiting(463)
DS1025-lipoxygenase inhibitionn/an/aPhase 2 studies are currently recruiting(464, 465)
BI 1467335Vascular adhesion protein-1 inhibitionn/an/aEarly phase 2 studies are currently recruiting(466)
MSDC-0602KMitochondrial pyruvate carrier inhibitionEvidence for improved metabolic as well as surrogates of hepatic inflammation and fibrosis on interim analysisn/aLarge phase 2 study (EMMINENCE) currently recruiting(467, 468)
Hepatic lipid metabolism
GS-0976Acetyl-Coenzyme A Carboxylase InhibitionImproved liver chemistry and hepatic steatosisn/a(469, 470)
AramcholStearoyl coenzyme A desaturase 1 inhibitionNo change in liver chemistry, but decreased hepatic steatosisn/aPhase 2b completed although not reported.(471)
Gut microbiome
ProbioticsAlteration in gut microbiomeImprovements in liver chemistry and steatosisn/a(472)
SolithromycinMacrolide antibioticn/aImproved NAS in small proof-of-concept study(473)
Fecal microbiome transplantationn/an/aSeveral phase 2 studies currently recruiting(474–476)

Abbreviations: FGF19, fibroblast growth factor 19; FGF21, fibroblast growth factor 21; FXR, farnesoid X receptor; NAFLD, nonalcoholic fatty liver disease; NASH, nonalcoholic steatohepatitis; PPARα/δ, peroxisome proliferator-activated receptor alpha/delta, PPARα/γ, peroxisome proliferator-activated receptor alpha/gamma; PSC, primary sclerosing cholangitis.

Table 3.

Lipid Lowering Drugs as Pharmacotherapy to Treat NAFLD

AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
SimvastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo evidence of benefit in a small (n = 16) studyStudy may have been underpowered to detect histological improvement.(477, 478)
AtorvastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo improvement on inflammation or fibrosis on liver histology(479–481)
PitavastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo improvement on liver histology(482, 483)
Omega3-fatty acidsSome evidence for improved liver chemistry and decreased steatosisLittle evidence to suggest significant improvements in inflammation or fibrosis(484–488)
FenofibratePPARα agonistSome evidence for improved liver chemistry and steatosis on ultrasound scanning, but no change in intra-hepatic TAG in detailed mechanistic studiesn/aSome conflicting data, unlikely to have any significant impact on liver TAG content.(489–492)
PemafibratePPARα agonistn/an/aPhase 2 studies are currently actively recruiting.(493)
AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
SimvastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo evidence of benefit in a small (n = 16) studyStudy may have been underpowered to detect histological improvement.(477, 478)
AtorvastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo improvement on inflammation or fibrosis on liver histology(479–481)
PitavastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo improvement on liver histology(482, 483)
Omega3-fatty acidsSome evidence for improved liver chemistry and decreased steatosisLittle evidence to suggest significant improvements in inflammation or fibrosis(484–488)
FenofibratePPARα agonistSome evidence for improved liver chemistry and steatosis on ultrasound scanning, but no change in intra-hepatic TAG in detailed mechanistic studiesn/aSome conflicting data, unlikely to have any significant impact on liver TAG content.(489–492)
PemafibratePPARα agonistn/an/aPhase 2 studies are currently actively recruiting.(493)

Abbreviations: HMG CoA, 3-hydroxy-3-methylglutaryl-coenzyme A; NAFLD, nonalcoholic fatty liver disease; PPARα, proliferator-activated receptor alpha; TAG, triglyceride.

Table 3.

Lipid Lowering Drugs as Pharmacotherapy to Treat NAFLD

AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
SimvastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo evidence of benefit in a small (n = 16) studyStudy may have been underpowered to detect histological improvement.(477, 478)
AtorvastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo improvement on inflammation or fibrosis on liver histology(479–481)
PitavastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo improvement on liver histology(482, 483)
Omega3-fatty acidsSome evidence for improved liver chemistry and decreased steatosisLittle evidence to suggest significant improvements in inflammation or fibrosis(484–488)
FenofibratePPARα agonistSome evidence for improved liver chemistry and steatosis on ultrasound scanning, but no change in intra-hepatic TAG in detailed mechanistic studiesn/aSome conflicting data, unlikely to have any significant impact on liver TAG content.(489–492)
PemafibratePPARα agonistn/an/aPhase 2 studies are currently actively recruiting.(493)
AgentMechanism of actionImpact on liver biochemistry and / or non-invasive assessments of NAFLD disease statusImpact on liver histologyCommentsReferences
SimvastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo evidence of benefit in a small (n = 16) studyStudy may have been underpowered to detect histological improvement.(477, 478)
AtorvastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo improvement on inflammation or fibrosis on liver histology(479–481)
PitavastatinHMG CoA reductase inhibitorSome evidence for improved liver chemistryNo improvement on liver histology(482, 483)
Omega3-fatty acidsSome evidence for improved liver chemistry and decreased steatosisLittle evidence to suggest significant improvements in inflammation or fibrosis(484–488)
FenofibratePPARα agonistSome evidence for improved liver chemistry and steatosis on ultrasound scanning, but no change in intra-hepatic TAG in detailed mechanistic studiesn/aSome conflicting data, unlikely to have any significant impact on liver TAG content.(489–492)
PemafibratePPARα agonistn/an/aPhase 2 studies are currently actively recruiting.(493)

Abbreviations: HMG CoA, 3-hydroxy-3-methylglutaryl-coenzyme A; NAFLD, nonalcoholic fatty liver disease; PPARα, proliferator-activated receptor alpha; TAG, triglyceride.

Glucose lowering agents in the treatment of NAFLD

While there are many agents currently at various stages of development, there are many studies that have used established glucose lowering, antidiabetic agents with an existing license for glucose control as potential treatments for NAFLD (Table 1). The use of these agents can, therefore, be considered not only to optimize glycemic control, but also potentially to convey a beneficial impact on the liver.

Biguanides

Metformin

Metformin is established as the first-line pharmacotherapy therapy in the treatment of T2D. It has multiple mechanisms of action including the activation of AMP kinase as well as altering mitochondrial function and cellular redox state (494). Rodent studies have provided a considerable body of evidence to suggest that it may have utility in the treatment of NAFLD. However, clinical studies have been less convincing (495). While there was initially evidence to suggest that it may have a benefit on lipid accumulation and inflammation, more recent meta-analysis of data has suggested a lack of convincing histological benefit and therefore as a therapy specifically for NAFLD, although safe, it is not currently advocated (400, 401).

Peroxisome proliferator-activated receptor gamma agonists (thiazolidinediones)

The thiazolidinediones are a class of glucose-lowering agents that act predominantly, although not exclusively, as peroxisome proliferator-activated receptor gamma agonists. These agents act to improve insulin sensitivity; much of their activity is directed toward adipose tissue where they increase adiponectin expression and drive adipocyte differentiation. Clinical studies have consistently demonstrated clinical benefits including reductions in hepatic steatosis and resolution of NASH (161, 402, 403). The landmark PIVENS study (in patients with NASH, but without T2D) compared pioglitazone to vitamin E and placebo (404). While vitamin E appeared superior to placebo in terms of the primary end-point (although this was complicated by the lack of hepatocyte ballooning, a defined part of the composite primary end-point, in many patients at baseline), pioglitazone reduced steatosis and inflammation and resolved NASH in 47% of patients (compared with 34% on vitamin E and 18% on placebo). More recent studies including patients with T2D have continued to demonstrate histological benefits of pioglitazone treatment with the additional suggestion that fibrosis may improve (161).

Despite the demonstrable benefits to liver histology and published clinical guidance that advocates consideration of its use, it remains relatively underutilized as a therapeutic option. This may reflect concerns over its adverse effect profile, including fluid retention, weight gain (up to 5 kg over 3 years), and increased risk of bone fractures (405). Concerns over cardiovascular risk have been raised. This perhaps reflects previous studies using rosiglitazone suggesting increased myocardial infarction risk, although subsequently the RECORD study showed no overall impact on cardiovascular morbidity or mortality (496, 497). Specifically with regards to pioglitazone, data suggest that it may actually reduce cardiovascular and cerebrovascular risk as well as progression to T2D, all of which are important, bearing in mind the adverse cardiovascular risk profile associated with NAFLD (498–500). Its use in patients with compromised cardiac function is contraindicated due to concerns that this may worsen symptoms potential due to increased fluid retention. The most recent analysis of data has been reassuring with regards to the concerns that had been raised about the risk of bladder cancer associated with pioglitazone use (501).

Glucagon-like peptide 1 analogs

The incretin effect describes the augmentation of the insulin secretion in response to an oral, as opposed to intravenous, glucose load. This is largely mediated by GLP-1, which is released from the intestinal L-cells. GLP-1 agonist therapy is now established as a highly potent and efficacious glucose lowering and weight loss intervention. Retrospective analysis of data from the LEAD series of studies demonstrated that liraglutide caused a dose-dependent decrease in ALT in those individuals in whom liver chemistry was abnormal at baseline (502). There was no impact in those with normal liver blood test. Subsequently, prospective studies have demonstrated histological improvement and resolution of NASH using liraglutide; 36% of patients had resolution of NASH (compared with 9% in the placebo-treated arm) (408). Currently, this is the only study that has looked at histological outcomes, although other studies using liraglutide, exenatide, and lixisenatide have shown improvements in hepatic insulin sensitivity, liver biochemistry, and lipid content (409–414, 416, 419).

The mechanism of action of GLP-1 analogs to improve NAFLD remains to be clarified. Controversy remains as to whether the GLP-1 receptor is expressed in human hepatocytes, although in vitro observations have suggested that GLP-1 agonism can decrease DNL in human hepatocyte primary cultures (503). However, many studies have failed to identify the GLP-1 receptor in human and rodent hepatocytes (504, 505). The confounding issue of weight loss and improvements in glycemic control remain to be disentangled from the potential direct benefits of GLP-1 analog therapy, and therefore a mechanism of action that indirectly benefits the liver remains entirely plausible.

There is now an established body of evidence from the SUSTAIN, LEADER, and REWIND trials to demonstrate deceased cardiovascular risk associated with semaglutide, liraglutide, and dulagluitde use (506–508). Taking into account the increased cardiovascular risk associated with NAFLD, there is a growing body of evidence to suggest that this class of agent has the potential to offer significant clinical benefit to all patients with NAFLD irrespective of whether they do or do not have T2D. A large phase 2 study in patients with NASH using liver biopsy endpoints is currently trialing 3 different doses of semaglutide versus placebo and is due to complete toward the end of 2019 (420). Long-acting, once weekly GLP-1 analog therapy is now widely used. Although there are no histological data, weight and liver chemistry improve with treatment (421, 422) and dedicated phase 2 studies are currently recruiting (509, 510). More recently, novel compounds with multiple agonist properties are being developed; for example, HM1522 is a GLP-1/gastric inhibitory polypeptide/glucagon triple agonist and is currently be trialed in the context of NAFLD (423).

Dipeptidyl peptidase IV inhibitors

The enzyme dipeptidyl peptidase IV (DPP IV) is responsible to the degradation of endogenous GLP-1. Synthetic GLP-1 analogs that are currently used clinically are resistant to the actions of DPP IV, and this is a crucial mechanism that prolongs their half-life and facilitates their clinical utility. Very few histological studies have been performed to examine the impact of DPP IV inhibition in NAFLD; in one study (without a placebo control arm), there was some evidence of histological improvement in NAS after 1 year of treatment, but no benefit was found in a further study (511, 512). Additional studies have looked at liver TAG content and biochemistry (412, 424, 425) and failed to demonstrate significant benefit, although a single study has reported improvement with vildagliptin (426). However, currently, there is no convincing evidence that this class of agent has any beneficial impact upon NAFLD.

Sodium glucose co-transporter 2 inhibitors

The most recent class of glucose-lowering agents to gain approval for clinical use are the sodium glucose co-transporter 2 (SGLT2) inhibitors. Through inhibition of SGLT2, they prevent the reabsorption of the vast majority (>90%) of glucose that has been filtered. They are highly effective as glucose-lowering agents and cause weight loss (as a consequence of glucose loss). In addition, studies have consistently demonstrated a significant improvement in cardiovascular outcome associated with SGLT2 use (513–515), although the mechanisms that underpin this observation are yet to be determined. Controlled clinical studies in the context of NAFLD that have histological endpoints have not been completed; however, several small studies (including open label and uncontrolled studies) have shown improvements in liver chemistry, improved glycemic control, weight loss, and reductions in liver TAG content as assessed using 1H-MRS, although these benefits have not been observed in all studies that have been reported (406, 427, 428, 430, 516, 517). Recently, double-blind placebo controlled studies have been reported using canagliflozin (431) and dapagliflozin (434, 435) in patients with T2D. Hepatic insulin sensitivity improved with canagliflozin, but not dapagliflozin, and with both treatments, there were modest reduction in intrahepatic TAG (431, 434). A single uncontrolled study has reported histological improvements in NAFLD severity after 24 weeks of treatment with canagliflozin (518). Studies in patients without T2D have not been reported.

While there is potential that this class of agents may convey significant clinical benefit (including cardiovascular risk reduction and weight loss), there is a pressing need for well-designed, prospective clinical studies with relevant clinical outcomes. These drugs are generally well tolerated; genitourinary infections are reported in up to 5% of individuals, and there is the potential to develop increased urinary frequency, dehydration, and postural hypotension. Concerns have also been raised about the development of diabetic ketoacidosis, and while recent data seem to be reassuring, there is still a need for vigilance for potentially life-threatening adverse events (519, 520). Licogliflozin is a dual SGLT1 and SGLT2 inhibitor that has been shown to cause weight loss in patients with and without T2D (521), and studies are currently examining its impact on hepatic steatosis (436).

Lipid-lowering drugs

3-hydroxy-3-methyl-glutaryl-coenzyme A reductase inhibitors

It is well established that patients with NAFLD are at increased cardiovascular risk. Furthermore, in patients with T2D, the use of lipid-lowering therapy, usually in the form of 3-hydroxy-3-methyl-glutaryl-coenzyme A (HMG-CoA) reductase inhibitors (statins) is regarded as best clinical practice. Statins are known to have multiple effects that might be beneficial to patients with NAFLD including anti-inflammatory actions as well as reducing oxidative stress. Statin therapy can be associated with elevated liver enzymes in some individuals. However, establishing a temporal relationship is important as elevations in liver enzymes may reflect underlying NAFLD and may not be related to the initiation of statin therapy.

Only a relatively small number of studies have specifically tested the hypothesis that statin therapy may alter the natural history of NAFLD, and many of these have been open label and lacking appropriate controls. Short-duration studies have suggested that liver chemistry may improve in those individuals with elevated liver blood tests at baseline and in those with evidence of elevated hepatic TAG content (482). In addition, there is some evidence to suggest improvements in liver chemistry as well as histological NAS assessment without alteration in fibrosis (479), but study design was compromised by the lack of a suitable placebo control group. Additional very small open label studies using pitavastatin and rosuvastatin have also shown improvements in liver chemistry, but not in liver histology (477, 480, 483). In a single, small, prospective placebo-controlled study, simvastatin (40 mg) used for 12 months failed to show any histological benefit or any impact on liver chemistry (478). However, the study had very small numbers (simvastatin n = 10, placebo n = 6) and may well have been underpowered to detect biochemical or histological changes.

Overall, while statins appear to be safe and reasonably well tolerated (522), there is a lack of appropriately designed clinical studies that have been able to determine if they have a beneficial impact, specifically upon the natural history of NAFLD. However, they have established proven efficacy to improve cardiovascular outcome, and therefore their use in patients with NAFLD should not necessarily be aimed at disease modification within the liver, but rather at more holistic cardiovascular reduction (523).

Omega-3 poly-unsaturated fatty acids

The beneficial effects of the omega-3 long chain fatty acids on circulating hypertriglyceridemia are well established. Many studies have been performed investigating whether these agents may have utility as a treatment for NAFLD, although relatively few have had histological primary end-points. The largest study included 243 patients with biopsy-proven NASH. Treatment did not impact upon liver steatosis, fibrosis, or inflammation (484, 485). Other studies have shown similar results (486), although a very small number of studies have shown modest improvements in NAFLD severity (487). Overall, there seems little compelling evidence that these agents significantly alter the natural history and progression of the more advanced stages of NAFLD but may have a role to limit hepatic TAG accumulation (488).

Peroxisome proliferator-activated receptor alpha agonists

Fibrates are widely used as lipid-lowering therapy, acting through a peroxisome proliferator-activated receptor alpha (PPARα)-dependent mechanism to drive lipid oxidation and utilization. There are very few clinical data that have systematically examined the impact of fibrate therapy in the context of NAFLD. An open label study compared fibrate therapy (fenofibrate 200 mg daily) to atorvastatin 20 mg/day or combination therapy and demonstrated improvements in liver biochemistry and ultrasonographic appearance of the liver (489). A further open label, randomized study has also shown improvements in liver biochemistry in comparison with pioglitazone treatment (490). However, detailed mechanistic studies performed as part of a randomized controlled trail (including a nicotinic acid treatment arm), failed to show any improvement in intrahepatic TAG content by either nicotinic acid or fenofibrate. However, both treatments decreased VLDL TAG levels—fenofibrate through increased clearance and nicotinic acid through decreased secretion. In addition, fenofibrate had no impact on peripheral, hepatic, or adipose tissue insulin sensitivity (491, 492). There are currently no published histological data from clinical studies. As a result, it is hard to draw any significant conclusion as to the utility of fibrates in the treatment of NAFLD. Pemafibrate is a novel selective PPARα agonist with evidence from preclinical models to suggest that it can improve NASH (524). It is currently being used in phase 2 clinical studies in patients with NAFLD (493).

Proprotein convertase subtilisin kexin type 9 inhibitors

Proprotein convertase subtilisin kexin type 9 (PCSK9) is a fundamental regulator of lipid homeostasis. It is secreted from the liver into the circulation where it binds to the low-density lipoprotein (LDL) receptor and enhances its lysosomal degradation and, as a consequence, raises plasma LDL cholesterol levels. PCSK9 inhibitors disrupt this pathway, lower LDL cholesterol, and decrease cardiovascular events (525). The relationship of PCSK9 with the pathogenesis of NAFLD and NASH is yet to be fully elucidated, and some studies have suggested a positive correlation of circulating PCSK9 levels with markers of NASH severity (526), while other have found no relationship with either liver expression or circulating levels (527). The question as to whether this class of agent can improve the histological features of NASH has not been addressed in clinical studies. However, PCSK9 inhibitors seem well tolerated, and there is no evidence that they cause deterioration in liver chemistry (525, 528) although dedicated studies to examine the impact of NAFLD/NASH have not been performed.

Liver-targeted therapies

Farnesoid X receptor agonists

Obeticholic acid

Signaling through FXR regulates multiple processes within the liver, including carbohydrate and lipid metabolisms as well as regeneration and repair. There is an extensive body of preclinical literature suggesting that FXR agonism may have beneficial effects in human NAFLD. Obeticholic acid (OCA) is a synthetic FXR agonist that is already licensed for the treatment of primary biliary cholangitis, where it has been shown to decrease bilirubin and alkaline phosphatase levels (529).

Proof-of-concept and phase 2b studies using OCA have been completed and shown improvements in liver histology in 45% of individuals (compared to 21% in the placebo arm) and resolution of NASH in 22% (vs. 13% in the placebo arm). Thirty-five percent of patents had improvements in liver fibrosis (vs. 19% in the placebo arm) (437). A post hoc analysis suggested that most benefit of OCA was derived in those patients who also lost weight (530). These encouraging data led to the initiation of a global, multicenter phase 3 study that is currently recruiting (REGENERATE) (531). A recent planned interim analysis after 18 months of therapy including the 931 patients recruited so far has suggested that the study has met the primary end-point of improvement in fibrosis, and the adverse effect profile has not identified anything unexpected. Cardiovascular events reported to date are no difference across the 2 active treatment arms or placebo (532).

The REGERATE study is recruiting patients with mild to moderate fibrosis (F2–F3). A further phase 3 study (REVERSE) has also been initiated in which OCA is being trialed in patients with established (but fully compensated) NAFLD cirrhosis (533).

Other FXR agonists

Other FXR agonists are currently in development although at a much earlier stage. GS-9674 has passed through phase 1 studies, and currently a phase 2 study in the context of NASH is ongoing (534). Recently, published data have shown biochemical improvements in liver chemistry in patients with primary sclerosing cholangitis (438). Tropifexor (LJN452) is a nonbile acid-derived FXR agonist and is currently being trialed in phase 2 studies in patients with NASH (439).

Dual and pan-peroxisome proliferator-activated receptor agonists

Elafibranor (PPAR α/δ dual agonist).

PPARα and PPARδ are nuclear transcription factors that are potent regulators of lipid metabolism and inflammation in hepatocytes as well as other tissues. A substantial body of evidence not only implicated their role in the pathogenesis of NAFLD but also highlighted their potential role as a therapeutic target. The dual PPARα and PPARδ agonist, elafibranor, has been trialed in phase 2 studies. In a 52-week randomized, double-blind, placebo-controlled study, 19% of patients in the active treatment arm resolved NASH compared to 13% in the placebo arm without worsening of fibrosis (440). The primary end-point was not met in this study, but in a post hoc analysis in those patients with more advanced disease, there did appear to be significant benefit. In addition, there were metabolic benefits in patients with diabetes (at the highest doses tested) that included reduction in fasting plasma glucose, circulating FFAs, C-peptide and glycated hemoglobin. A large phase 3 study (RESOLVE-IT) is currently recruiting (535).

Lanifibranor (IVA337, pan-PPAR agonist)

Lanifibranor is able to activate all 3 (α, δ, and γ) PPARs. In preclinical studies, it has been reported to decrease tissue fibrosis in nonliver (536, 537) as well as liver models (538). A phase 2, randomized, double-blind, placebo-controlled study in patients with NASH with histological end-points is currently recruiting (441).

Saroglitazar (dual α/γ agonist)

Dual PPARα/γ agonism using saroglitazar has been shown to improve liver histology in preclinical rodent models (539). Clinical studies have shown benefits on circulating lipid profiles in those patients inadequately treated on atorvastatin therapy (442). Currently, an early phase 2 study in patients with NAFLD looking at changes in liver biochemistry with 3 doses of saroglitazar versus placebo is actively recruiting (540).

C–C chemokine receptor type 2 and type 5 dual antagonist

Cenicriviroc

Targeting the inflammatory response within the liver has been proposed as a strategy to limit cellular damage that may ultimately drive fibrosis. In response to hepatocyte injury or insult, KCs within in the liver secrete C–C chemokine ligand 2 (CCR2; also known as monocyte chemo-attractant protein 1), which can lead to monocyte recruitment driving the inflammatory response and, as a consequence, causing HSC activation and the resulting fibrotic response. The key signaling receptors mediating these actions appear to be CCR2 and CCR5, and dual antagonism in preclinical rodent models validated this strategy as a therapeutic target. In a phase 2b study including 289 patients with biopsy-proven NASH (51% with T2D), after 48 weeks of treatment with the dual CCR2/CCR5 antagonist, cenicriviroc (150 mg once daily), the primary end-point was not met (2-point improvement in NAS score without worsening of fibrosis); however, there was a significant increase in the proportion of patients who demonstrated improvements in liver fibrosis on biopsy (20.0% vs. 10.4%, P = .023) (443). There were no changes in steatosis, liver biochemistry, or metabolic variables, but systemic markers of inflammation (including C-reactive protein and IL-6) were decreased. The drug was well tolerated with few significant side effects.

Despite the modest benefits that were observed in the phase 2b study, a phase 3 study (AURORA) is currently underway with the aim of recruiting 2000 patients with a primary completion date during 2021 (541).

Vitamin E

Vitamin E is a potent antioxidant with the potential to exert anti-inflammatory actions. In the PIVENS trial (patients without T2D or cirrhosis), vitamin E, at a dose of 800 mg/ day, was superior to pioglitazone (at the intermediate dose of 30 mg/day) in terms of the predefined primary end-point of the study leading to an improvement in NASH in 43% versus 19% in the placebo arm) (404). Using end-points that have adopted in recent studies (NASH resolution), vitamin E achieved borderline significance over placebo (36% vs. 21%, P = .05) but was inferior to pioglitazone, There was no impact on fibrosis. In the TONIC trial, in which children and adolescents were treated with 300 mg/day, there was no overall benefit in terms of improvements in liver chemistry, steatosis, inflammation, or fibrosis (444). Concerns have been raised with respect to the long-term safety of higher doses of vitamin E (542), although this remains a contentious area with continued debate in the published literature with regards to the precise methodologies that have been used in the analyses (543). There is some evidence to suggest increased all-cause mortality, increased risk of hemorrhagic stroke although the data on the relationship to prostate cancer risk are less clear (544, 545).

Galectin-3 protein inhibitors

GR-MD-02

Galectin-3 protein is believed to play a crucial role in orchestrating the fibrotic response to inflammation with in the liver. The evidence base for its role as a therapeutic target is relatively limited; rodent models with genetic deletion have shown increased lipid accumulation, but either increased or decreased inflammation and fibrosis is dependent upon the precise model used. The GT-020 study was a first-in-class, phase 1 study using the galectin-3 inhibitor, GR-MD-02 (445); the drug was well tolerated with no major treatment emergent adverse effects. An early phase 2 study has been completed (NASH-CX) in patients with NASH cirrhosis. The primary end-point was a reduction in portal pressure, but liver biopsies were taken. The data are not formally published in the peer-reviewed literature (446), but there was no significant change in NAS assessment (although decreased hepatocyte ballooning was observed), and no change was observed in fibrosis score or collagen content on morphometric analysis.

Fibroblast growth factor 19 analogs

NGM282

. FGF19 has been implicated in the pathogenesis of NAFLD. It has a crucial role in bile acid synthesis as well as having important metabolic actions and in rodents administration has been shown to enhance energy expenditure, promote weight loss and improve lipid profiles (546). Transgenic over expression limits lipid synthesis (547).

NGM282 is a potent FGF19 analog, and in a phase 2 study including 166 patients with biopsy-proven NASH, the impact of 2 different doses of NGM282 were examined (3 mg or 6 mg administered subcutaneously) (447). Over the 12-week duration of the study, there was a significant reduction in liver fat (48% reduction with 3 mg and 60% reduction with 6 mg) as well as improvements in liver biochemistry. Adverse events were relatively common with injection site reactions as well as gastrointestinal disturbance being reported. Serum LDL cholesterol also increased (with a parallel decrease in high-density lipoprotein [HDL]-cholesterol), but recent data have shown that this can be lowered by co-administration with rosuvastatin without significant additional adverse effects (548). More recently, liver biopsy findings from an open label, 12-week study (doses 1mg and 3 mg) have been reported demonstrating significant reductions in steatosis, inflammation, and fibrosis (448). A larger phase 2 study aiming to recruit 250 participants with biopsy-proven NASH is currently actively recruiting (549).

FGF21 analogs

Fibroblast growth factor 21 is highly expressed in the liver, and it has an established role in the regulation of many metabolic features including weight, energy expenditure, and insulin sensitivity. The relationship between circulating levels of FGF21 and metabolic phenotype are complex, and there is evidence to suggest that levels rise with worsening metabolic phenotype that is consistent to resistance to the effects of FGF21. Specifically, in the context of NAFLD, recently published data suggest that FGF21 levels correlated most closely with adipose and skeletal muscle insulin resistance (but not hepatic) and that levels were positively associated with inflammation and ballooning (but not steatosis) as well as fibrosis stage (550). In a proof-of-concept study, 4 weeks of treatment with LY240531, a variant of FGF21, in patients with obesity and T2D, improved circuiting lipid profiles (decreased LDL cholesterol and TAG and increased in HDL cholesterol) (449).

More recently, pegbelfermin (BMS-986036), a PEGylated human FGF21 analog has been trialed in 80 overweight/obese individuals with NASH. Patients were recruited into a double-blind, randomized, placebo-controlled study investigating 2 doses of BMS-986036 (10 and 20 mg). The study was terminated early due to a larger than anticipated impact on the primary end-point at the planned interim analysis after 8 weeks of treatment. There was a significant reduction in hepatic steatosis as measured by MRI PDFF, and the drug was well tolerated with no serious adverse events or treatment-related drug withdrawals (450). A series of phase 2 studies are currently actively recruiting (551–558).

Thyroid hormone receptor-β agonists

MGL-3196 (Resmetirom)

MGL-3196 was developed as a liver-specific agonist of the thyroid hormone receptor-β (THR-β). Original studies aimed to explore its role in the treatment of hyperlipidemia and demonstrated improvements in LDL cholesterol, non-HDL cholesterol, lipoprotein(a), and a trend to improved circulating TAG levels importantly, without alteration in circulating thyroid hormone levels (559, 560). A randomized double blind, placebo-controlled study with 36 weeks of treatment in 116 patients with biopsy-proven NASH has been reported in abstract form only. The primary end-point was reduction in hepatic fat as measured by MRI PDFF, and this was achieved with a relative reduction of 36.3% versus 9.6% in the placebo arm (451). Further, data have now been released (although not peer-reviewed) suggest that after 36 weeks of treatment, there were significant improvements in the resolution of NASH alongside a sustained improvement in steatosis (50, 561). The drug was well tolerated in this cohort with no serious adverse events. On the basis of these results, a program of phase 3 studies is now being initiated.

Apoptosis signal-regulating kinase 1 inhibitors

Selonsertib.

Apoptosis signal-regulating kinase 1 (ASK1) is induced by cellular stress and is an important regulator of both inflammation and fibrosis. Preclinical data in rodent models has suggesting that inhibition of ASK1 has the potential in rodent models to modify the natural history and progression of NASH. Selonsertib (GS-4997) is a selective ASK1 inhibitor and has been used in an open label phase 2 study in 72 patients treated with 6 or 18 mg orally for 24 weeks (alone or in combination with simtuzumab—a monoclonal antibody directed against lysl oxidase-like-2 monoclonal that subsequently has been shown to have no efficacy as monotherapy) (453). In the final analysis, liver histology comparison was made with all selonsertib treated patients (with or without simtuzumab) versus simtuzumab alone. At the highest doses (18 mg), fibrosis decreased in 43% or patients (30% in the 6 mg dose arm and 20% in the simtuzmab alone treatment arm) (562). Changes in noninvasive markers, including imaging and serum tests paralleled the changes in liver histology. A further analysis of the data from this study aimed to look at quality of life; where there was histological improvement in NAS assessment (14 out of 68 patients), there was also improvement in quality-of-life scores (563).

Pentoxifyline

The mechanisms by which pentoxifylline may improve NASH are not fully understood, although it has an inhibitory action on proinflammatory cytokines as well as reducing the generation of free oxygen radicals. Treatment with pentoxifylline for 1 year resulted in histological improvement in NAS assessment as well as fibrosis (454). A further study also showed histological improvements, but without significant changes in liver chemistry (455). Currently, there are no actively recruiting studies registered on clinicaltrials.gov investigating the role of pentoxifylline in the treatment of NASH.

Caspase inhibition

Emricasan.

Emricasan is a pan-caspase inhibitor that has been reported to decrease apoptosis and inflammation in liver disease. In patients with cirrhosis (due to NASH and HCV), there was improved liver function after 3 months of treatment that was observed not only in the whole cohort, but in a subgroup analysis, the improvement was statistically significant in those patients with NASH cirrhosis, although the numbers treated were very small (placebo n = 9; emricasan n = 11) (456). Specifically in NAFLD patients without cirrhosis, but with elevated ALT, emricasan decreased ALT and AST as well as reducing circulating cleaved and full-length cytokeratin-18. The most dramatic effects were seen within 7 days of treatment and some, but not all, of the effects persisted throughout the full 28 days of treatment (457). Further studies in NASH are planned although currently not actively recruiting (564).

Targeting hepatic lipid metabolism

Acetyl-coenzyme A carboxylase inhibition

GS-0976.

Acetyl-coenzyme A carboxylase (ACC) is the rate-limiting step in DNL and has therefore become a target to limit hepatic TAG accumulation and potentially enhance lipid oxidation. GS-0976 is a small molecular inhibitor of ACC and has been investigated in a small number of clinical studies. In an open label, uncontrolled, prospective study in 10 individuals treated for 12 weeks, DNL (measured by deuterated water incorporation into palmitate) was reduced by 22%. Hepatic steatosis measured by MRI PDFF decreased as did ALT and liver stiffness (469). In a much larger, randomized, placebo-controlled study (n = 126), GS-0976 at the highest doses only (20 mg) decreased hepatic steatosis, but there was no change in liver stiffness although tissue inhibitor of metalloproteinase 1 levels decreased (470). The drug was safe and well tolerated, although serum TAG levels increased in all groups treated with GS-0976.

Stearoyl coenzyme A desaturase 1 inhibition

Aramchol.

Aramchol is a novel fatty acid–bile acid conjugate molecule (3β-arachidyl-amido, 7α-12α-dihydroxy, 5β-cholan-24-oic acid). Its mechanism of action is likely to be through inhibition of stearoyl coenzyme A desaturase 1, which is an important regulatory step in lipid synthesis within the liver. In an early phase 2 trial for 3 months, there was a significant reduction in liver fat as measured by 1H-MRS when compared to placebo (only in the higher 300 mg/day dose). There were no significant changes in liver chemistry when compared to placebo (471). The drug was safe and well tolerated, and a larger phase 2b study has been completed although data are not formally reported in the peer-reviewed literature. There were significant improvement in liver chemistry as well as resolution of NASH, but no significant change in fibrosis (565).

Modulation of the gut microbiome

It is now established that the gut microbiome has the potential to have a profound regulatory impact upon metabolic phenotype (see The gut microbiome section). Therefore, various strategies have been employed to try and modify the microbiome composition in such a way as to promote a beneficial metabolic phenotype. A small number of randomized, controlled trials have been performed that have administered differing probiotics. Improvements in liver chemistry as well as markers of inflammation and insulin sensitivity have improved; however, additional studies are clearly required with demonstrable histological benefit before their widespread use can be recommended in patients with NAFLD (472).

A very small (n = 6) proof-of-concept study has been performed using solithromycin, a macrolide antibiotic. After 13 weeks of treatment, there was histological improvement in NAS in all patients as well as improvements in liver chemistry in almost all individuals (473). However, treatment of patients with biopsy-proven NASH and elevated aminotransferases (n = 15) with rifaximin (400 mg twice daily for 6 weeks) had no impact on ALT levels, hepatic insulin sensitivity, or liver TAG content (566). Ongoing studies are also looking at the potential of fecal microbiome transplantation as a therapeutic strategy in NAFLD (474–476, 567, 568).

Other treatments in early development

Several other treatment targets are at the very early stages of development and have yet to be formally trialed with dedicated histological or robust and meaningful clinical end points. SGM-1019 is an inflammasome inhibitor and has been hypothesized to limit hepatocyte injury. Early phase clinical studies have been performed, and it is safe and well tolerated; phase 2 studies are currently in progress (458, 459).

Additional modulation and limitation of the inflammatory response could be mediated through antagonism of the TLR-4. JKB-121 is a weak molecular antagonist of TLR-4 receptor and while preclinical data suggest promise in its ability to prevent the development of NASH, an early phase 2 study has shown no difference when compared to placebo in reductions in liver fat content and relatively high (dose-dependent) adverse events that lead to drug withdrawal (460).

IMM-124E is a product derived from bovine colostrum that contains high levels of anti-E.coli-LPS IgG. It has been suggested that this may limit exposure to bacterial endotoxins and has been shown to limit acute colitis in preclinical models (569). A phase 2 study in 133 patients with NASH has been completed using an oral preparation at 2 different doses although results have not yet been reported (461).

Amlexanox is an inhibitor of IKKb and TANK-binding kinase 1 both of which are important signaling molecules that coordinate the inflammatory response. In a small proof-of-concept study in patients with T2D, there was improvement in glycemic control as well as evidence for decreased hepatic steatosis (462).

Tipelukast (MN-001) is an anti-inflammatory and antifibrotic agent that inhibits phosphodiesterase and 5-lipoxygenase activity as well as being a leukotriene receptor antagonist. It has potential as a lipid-lowering agent and is currently being trialed in patients with NASH/NAFLD and hypertriglyceridemia (570). The stated primary outcomes relates to cholesterol metabolism and circulating TAG levels as opposed to direct assessments of liver (463). DS102 is a bioactive lipid that also inhibits 5-lipoxygenase activity. Phase 1 studies have been completed (464), and a phase 2 study in patients with NAFLD is actively recruiting (463).

Vascular adhesion protein-1 (VAP) is a membrane-bound amine oxidase that is thought to have a key role in recruiting inflammatory cells into the liver. Circulating levels of the soluble form of VAP are elevated in patients with NASH (199) and an early phase 2 study trialing a novel VAP inhibitor (BI 1467335) is currently recruiting (466).

Inhibition the mitochondrial pyruvate carrier has been proposed as a potential treatment strategy for NAFLD. This limits pyruvate flux into the tricarboxylic acid and is associated with metabolically beneficial effects including enhanced insulin sensitivity and decreased DNL alongside decreased stellate cell activation (571). A large phase 2 study (EMMINENCE) is currently recruiting using the mitochondrial pyruvate carrier inhibitor MSDC-0602K (467). An interim analysis has shown improvement in metabolic parameters as well as surrogate markers of inflammation and fibrosis (468).

DGAT2 catalyzes the final step in hepatic TAG synthesis. Antisense-mediated inhibition of DGAT2 may have potential to limit hepatic steatosis, although data have currently been presented in abstract form only (572).

CF102 is an adenosine A3 receptor agonist that in preclinical models has shown anti-inflammatory and proapoptotic actions in the liver (573). It has been trialed in small numbers of patients with HCC (574) and is currently under evaluation as a potential treatment for NASH (575).

Conclusions

The magnitude of the clinical burden associated with NAFLD has heightened the need to understand the natural history and risks associated with the condition. While in many individuals, in the absence of significant fibrosis, this may be a relatively benign condition, there is no doubt that with advancing fibrosis there is significant morbidity and mortality. A significant challenge continues to be the development of noninvasive biomarkers (to replace liver biopsy) that can accurately detect and stage disease (and therefore risk) and that can also be used to track changes over time and the impact of specific interventions.

Our understanding of the determinants of disease progression is developing rapidly, but we are still some way from being able to categorically identify those patients who will progress to advanced disease and those who will not. This is clearly important if we are to try to target appropriate therapies to individuals who are most likely to benefit. The future therapeutic landscape is relatively rich, with an impressive array of compounds with mechanisms of action targeting differing elements of the pathogenesis of NAFLD. The future of NAFLD pharmacotherapy will undoubtedly include combination therapies (paralleling the current treatment strategies for hypertension and T2D)—for example, combining limitation of hepatic TAG accumulation alongside therapies to reduce fibrosis and those to reduce cardiovascular risk. NAFLD epitomizes the multisystem disease and necessitates a holistic approach to its management. While we wait for the development of novel drug therapies, adopting a multidisciplinary approach to its management and combining the skills and expertise of multiple health-care professions would appear to set the current gold-standard of care.

Acknowledgments

This work was supported by the National Institute for Health Research (NIHR) Oxford Biomedical Research Centre, the Medical Research Council UK (program grant MR/P011462/1 to JWT); by the British Heart Foundation (senior fellowship to LH) and a University of Oxford/Novo Nordisk clinical research training fellowship (awarded to AM). The views expressed are those of the authors and not necessarily those of the NIHR or the Department of Health and Social Care UK.

Financial Support: Medical Research Council (program grant to JWT ref. MR/P011462/1); NIHR Oxford Biomedical Research Centre (principal investigator award to JT); University of Oxford/Novo Nordisk clinical fellowship (AM); British Heart Foundation (senior fellowship to LH ref. FS/15/56/31645).

Additional Information

Disclosure Summary: JFC has been an advisory board member for Novo Nordisk and Intercept; JWT has been an advisory board member for Novo Nordisk, Pfizer and Poxel. TM, AM and LH have nothing to declare.

References

1.

Younossi
 
ZM
,
Koenig
AB
,
Abdelatif
D
,
Fazel
Y
,
Henry
L
,
Wymer
M
.
Global epidemiology of nonalcoholic fatty liver disease-meta-analytic assessment of prevalence, incidence, and outcomes
.
Hepatology.
2016
;
64
(
1
):
73
84
.

2.

Kanwal
 
F
,
Kramer
JR
,
Duan
Z
,
Yu
X
,
White
D
,
El-Serag
HB
.
Trends in the burden of nonalcoholic fatty liver disease in a United States cohort of veterans
.
Clin Gastroenterol Hepatol.
2016
;
14
(
2
):
301
308.e1
.

3.

Allen
 
AM
,
Therneau
TM
,
Larson
JJ
,
Coward
A
,
Somers
VK
,
Kamath
PS
.
Nonalcoholic fatty liver disease incidence and impact on metabolic burden and death: a 20 year-community study
.
Hepatology.
2018
;
67
(
5
):
1726
1736
.

4.

El-Serag
 
HB
.
Epidemiology of viral hepatitis and hepatocellular carcinoma
.
Gastroenterology.
2012
;
142
(
6
):
1264
1273.e1
.

5.

Bray
 
F
,
Ferlay
J
,
Soerjomataram
I
,
Siegel
RL
,
Torre
LA
,
Jemal
A
.
Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries
.
CA Cancer J Clin.
2018
;
68
(
6
):
394
424
.

6.

Dyson
 
J
,
Jaques
B
,
Chattopadyhay
D
, et al.  
Hepatocellular cancer: the impact of obesity, type 2 diabetes and a multidisciplinary team
.
J Hepatol.
2014
;
60
(
1
):
110
117
.

7.

Rich
 
NE
,
Oji
S
,
Mufti
AR
, et al.  
Racial and ethnic disparities in nonalcoholic fatty liver disease prevalence, severity, and outcomes in the united states: a systematic review and meta-analysis
.
Clin Gastroenterol Hepatol.
2018
;
16
(
2
):
198
210.e2
.

8.

Lomonaco
 
R
,
Ortiz-Lopez
C
,
Orsak
B
, et al.  
Role of ethnicity in overweight and obese patients with nonalcoholic steatohepatitis
.
Hepatology.
2011
;
54
(
3
):
837
845
.

9.

Bril
 
F
,
Portillo-Sanchez
P
,
Liu
IC
,
Kalavalapalli
S
,
Dayton
K
,
Cusi
K
.
Clinical and histologic characterization of nonalcoholic steatohepatitis in African American patients
.
Diabetes Care.
2018
;41(1):187–192.

10.

Menke
 
A
,
Casagrande
S
,
Geiss
L
,
Cowie
CC
.
Prevalence of and trends in diabetes among adults in the United States, 1988-2012
.
JAMA.
2015
;
314
(
10
):
1021
1029
.

11.

Romeo
 
S
,
Kozlitina
J
,
Xing
C
, et al.  
Genetic variation in PNPLA3 confers susceptibility to nonalcoholic fatty liver disease
.
Nat Genet.
2008
;
40
(
12
):
1461
1465
.

12.

Li
 
J
,
Zou
B
,
Yeo
YH
, et al.  
Prevalence, incidence, and outcome of non-alcoholic fatty liver disease in Asia, 1999–2019: a systematic review and meta-analysis
.
Lancet Gastroenterol Hepatol.
2019;4(5):389–398
.

13.

Alazawi
 
W
,
Mathur
R
,
Abeysekera
K
, et al.  
Ethnicity and the diagnosis gap in liver disease: a population-based study
.
Br J Gen Pract.
2014
;
64
(
628
):
e694
e702
.

14.

Fan
 
JG
,
Kim
SU
,
Wong
VW
.
New trends on obesity and NAFLD in Asia
.
J Hepatol.
2017
;
67
(
4
):
862
873
.

15.

Misra
 
A
,
Soares
MJ
,
Mohan
V
, et al.  
Body fat, metabolic syndrome and hyperglycemia in South Asians
.
J Diabetes Complications.
2018
;
32
(
11
):
1068
1075
.

16.

Bedogni
 
G
,
Miglioli
L
,
Masutti
F
, et al.  
Incidence and natural course of fatty liver in the general population: the Dionysos study
.
Hepatology.
2007
;
46
(
5
):
1387
1391
.

17.

Machado
 
M
,
Marques-Vidal
P
,
Cortez-Pinto
H
.
Hepatic histology in obese patients undergoing bariatric surgery
.
J Hepatol.
2006
;
45
(
4
):
600
606
.

18.

World Health

Organization
.
Global Health Observatory (GHO) Data: Overweight and Obesity
. Geneva: World Health Organization;
2014
.

19.

Pradeepa
 
R
,
Anjana
RM
,
Joshi
SR
, et al. ;
ICMR-INDIAB Collaborative Study Group
.
Prevalence of generalized & abdominal obesity in urban & rural India–the ICMR-INDIAB Study (Phase-I) [ICMR–NDIAB-3]
.
Indian J Med Res.
2015
;
142
(
2
):
139
150
.

20.

Li
 
Z
,
Xue
J
,
Chen
P
,
Chen
L
,
Yan
S
,
Liu
L
.
Prevalence of nonalcoholic fatty liver disease in mainland of China: a meta-analysis of published studies
.
J Gastroenterol Hepatol.
2014
;
29
(
1
):
42
51
.

21.

Majumdar
 
A
,
Misra
P
,
Sharma
S
,
Kant
S
,
Krishnan
A
,
Pandav
CS
.
Prevalence of nonalcoholic fatty liver disease in an adult population in a rural community of Haryana, India
.
Indian J Public Health.
2016
;
60
(
1
):
26
33
.

22.

Anurag
 
L
,
Aniket
S
,
Shalik
J
,
Amarja
L
,
Dhananjay
R
,
Sachin
J
.
Non-alcoholic fatty liver disease prevalence and associated risk factors–a study from rural sector of Maharashtra
.
Trop Gastroenterol.
2015
;
36
(
1
):
25
30
.

23.

Soresi
 
M
,
Noto
D
,
Cefalù
AB
, et al. ;
Metabolic Syndrome Study Group
.
Nonalcoholic fatty liver and metabolic syndrome in Italy: results from a multicentric study of the Italian Arteriosclerosis society
.
Acta Diabetol.
2013
;
50
(
2
):
241
249
.

24.

Targher
 
G
,
Bertolini
L
,
Padovani
R
, et al.  
Prevalence of nonalcoholic fatty liver disease and its association with cardiovascular disease among type 2 diabetic patients
.
Diabetes Care.
2007
;
30
(
5
):
1212
1218
.

25.

Williamson
 
RM
,
Price
JF
,
Glancy
S
, et al. ;
Edinburgh Type 2 Diabetes Study Investigators
.
Prevalence of and risk factors for hepatic steatosis and nonalcoholic fatty liver disease in people with type 2 diabetes: the Edinburgh Type 2 Diabetes Study
.
Diabetes Care.
2011
;
34
(
5
):
1139
1144
.

26.

Dai
 
W
,
Ye
L
,
Liu
A
, et al.  
Prevalence of nonalcoholic fatty liver disease in patients with type 2 diabetes mellitus: a meta-analysis
.
Medicine (Baltimore).
2017
;
96
(
39
):
e8179
.

27.

Younossi
 
Z
,
Anstee
QM
,
Marietti
M
, et al.  
Global burden of NAFLD and NASH: trends, predictions, risk factors and prevention
.
Nat Rev Gastroenterol Hepatol.
2018
;
15
(
1
):
11
20
.

28.

Wattacheril
 
J
,
Sanyal
AJ
.
Lean NAFLD: an underrecognized outlier
.
Curr Hepatol Rep.
2016
;
15
(
2
):
134
139
.

29.

Younes
 
R
,
Bugianesi
E
.
NASH in lean individuals
.
Semin Liver Dis.
2019
;
39
(
1
):
86
95
.

30.

Younossi
 
ZM
,
Stepanova
M
,
Negro
F
, et al.  
Nonalcoholic fatty liver disease in lean individuals in the United States
.
Medicine (Baltimore).
2012
;
91
(
6
):
319
327
.

31.

Sookoian
 
S
,
Pirola
CJ
.
Systematic review with meta-analysis: risk factors for non-alcoholic fatty liver disease suggest a shared altered metabolic and cardiovascular profile between lean and obese patients
.
Aliment Pharmacol Ther.
2017
;
46
(
2
):
85
95
.

32.

Stefan
 
N
,
Schick
F
,
Häring
HU
.
Causes, characteristics, and consequences of metabolically unhealthy normal weight in humans
.
Cell Metab.
2017
;
26
(
2
):
292
300
.

33.

Wei
 
JL
,
Leung
JC
,
Loong
TC
, et al.  
Prevalence and severity of nonalcoholic fatty liver disease in non-obese patients: a population study using proton-magnetic resonance spectroscopy
.
Am J Gastroenterol.
2015
;
110
(
9
):
1306
1314; quiz 1315
.

34.

Honda
 
Y
,
Yoneda
M
,
Kessoku
T
, et al.  
Characteristics of non-obese non-alcoholic fatty liver disease: effect of genetic and environmental factors
.
Hepatol Res.
2016
;
46
(
10
):
1011
1018
.

35.

Fracanzani
 
AL
,
Petta
S
,
Lombardi
R
, et al.  
Liver and cardiovascular damage in patients with lean nonalcoholic fatty liver disease, and association with visceral obesity
.
Clin Gastroenterol Hepatol.
2017
;
15
(
10
):
1604
1611.e1
.

36.

Leung
 
JC
,
Loong
TC
,
Wei
JL
, et al.  
Histological severity and clinical outcomes of nonalcoholic fatty liver disease in nonobese patients
.
Hepatology.
2017
;
65
(
1
):
54
64
.

37.

Kumar
 
R
,
Rastogi
A
,
Sharma
MK
, et al.  
Clinicopathological characteristics and metabolic profiles of non-alcoholic fatty liver disease in Indian patients with normal body mass index: Do they differ from obese or overweight non-alcoholic fatty liver disease?
 
Indian J Endocrinol Metab.
 
2013
;
17
(
4
):
665
671
.

38.

Marchesini
 
G
,
Bugianesi
E
,
Forlani
G
, et al.  
Nonalcoholic fatty liver, steatohepatitis, and the metabolic syndrome
.
Hepatology.
2003
;
37
(
4
):
917
923
.

39.

Hagström
 
H
,
Nasr
P
,
Ekstedt
M
, et al.  
Risk for development of severe liver disease in lean patients with nonalcoholic fatty liver disease: a long-term follow-up study
.
Hepatol Commun.
2018
;
2
(
1
):
48
57
.

40.

Cruz
 
AC Dela
,
Bugianesi
E
,
George
J
, et al.  
Characteristics and long-term prognosis of lean patients with nonalcoholic fatty liver disease (379)
.
Gastroenterology.
2014
;
146
(
5
):
S909
.

41.

Cusi
 
K
.
Nonalcoholic steatohepatitis in nonobese patients: not so different after all
.
Hepatology.
2017
;65(1):4–7.

42.

Hazlehurst
 
JM
,
Tomlinson
JW
.
Non-alcoholic fatty liver disease in common endocrine disorders
.
Eur J Endocrinol.
2013
;
169
(
2
):
R27
R37
.

43.

Kim
 
S
,
Kwon
H
,
Park
JH
, et al.  
A low level of serum total testosterone is independently associated with nonalcoholic fatty liver disease
.
BMC Gastroenterol.
2012;12:69
.

44.

Haider
 
A
,
Gooren
LJG
,
Padungtod
P
,
Saad
F
.
Improvement of the metabolic syndrome and of non-alcoholic liver steatosis upon treatment of hypogonadal elderly men with parenteral testosterone undecanoate
.
Exp Clin Endocrinol Diabetes.
2010;118(3):167–171
.

45.

Hazlehurst
 
JM
,
Oprescu
AI
,
Nikolaou
N
, et al.  
Dual-5α-reductase inhibition promotes hepatic lipid accumulation in man
.
J Clin Endocrinol Metab.
2016;101(1):103–113
.

46.

Kauffman
 
RP
,
Baker
TE
,
Baker
V
,
Kauffman
MM
,
Castracane
VD
.
Endocrine factors associated with non-alcoholic fatty liver disease in women with polycystic ovary syndrome: do androgens play a role?
Gynecol Endocrinol.
2010
;
26
(
1
):
39
46
.

47.

Rockall
 
AG
,
Sohaib
SA
,
Evans
D
, et al.  
Hepatic steatosis in Cushing’s syndrome: a radiological assessment using computed tomography
.
Eur J Endocrinol.
2003149(6):543–548
.

48.

Stefan
 
N
,
Ramsauer
M
,
Jordan
P
, et al.  
Inhibition of 11β-HSD1 with RO5093151 for non-alcoholic fatty liver disease: a multicentre, randomised, double-blind, placebo-controlled trial
.
Lancet Diabetes Endocrinol.
2014;2(5):406–416
.

49.

Ittermann
 
T
,
Haring
R
,
Wallaschofski
H
, et al.  
Inverse association between serum free thyroxine levels and hepatic steatosis: results from the Study of Health in Pomerania
.
Thyroid.
2012
;
22
(
6
):
568
574
.

50.

Madrigal pharmaceuticals initiates phase 3, multinational, double-blind, randomized, placebo-controlled study of MGL-3196 (resmetirom) in patients with non-alcoholic steatohepatitis (NASH) and fibrosis to resolve NASH and reduce progression to cirrhosis a [press release]
.
Madrigal Pharmecueticals .
2019, March 28
. http://www.madrigalpharma.com/wp-content/uploads/2019/03/Madrigal-Pharmaceuticals-Initiates-Phase-3-NASH-Study.FINAL_.20190328.pdf. Accessed
September 16, 2019
.

51.

Gutierrez-Grobe
 
Y
,
Ponciano-Rodríguez
G
,
Ramos
MH
,
Uribe
M
,
Méndez-Sánchez
N
.
Prevalence of non alcoholic fatty liver disease in premenopausal, posmenopausal and polycystic ovary syndrome women: the role of estrogens
.
Ann Hepatol.
2010
;
9
(
4
):
402
409
.

52.

Florentino
 
G
,
Cotrim
HP
,
Florentino
A
, et al.  
Hormone replacement therapy in menopausal women: risk factor or protection to nonalcoholic fatty liver disease?
 
Ann Hepatol.
 
2012
;
11
(
1
):
147
149
.

53.

Sites
 
CK
,
L’Hommedieu
GD
,
Toth
MJ
,
Brochu
M
,
Cooper
BC
,
Fairhurst
PA
.
The effect of hormone replacement therapy on body composition, body fat distribution, and insulin sensitivity in menopausal women: a randomized, double-blind, placebo-controlled trial
.
J Clin Endocrinol Metab.
2005
;
90
(
5
):
2701
2707
.

54.

Nishizawa
 
H
,
Iguchi
G
,
Murawaki
A
, et al.  
Nonalcoholic fatty liver disease in adult hypopituitary patients with GH deficiency and the impact of GH replacement therapy
.
Eur J Endocrinol.
2012
;
167
(
1
):
67
74
.

55.

Ballestri
 
S
,
Nascimbeni
F
,
Baldelli
E
,
Marrazzo
A
,
Romagnoli
D
,
Lonardo
A
.
NAFLD as a sexual dimorphic disease: role of gender and reproductive status in the development and progression of nonalcoholic fatty liver disease and inherent cardiovascular risk
.
Adv Ther.
2017
;
34
(
6
):
1291
1326
.

56.

Browning
 
JD
,
Szczepaniak
LS
,
Dobbins
R
, et al.  
Prevalence of hepatic steatosis in an urban population in the United States: impact of ethnicity
.
Hepatology.
2004
;
40
(
6
):
1387
1395
.

57.

Ruhl
 
CE
,
Everhart
JE
.
Determinants of the association of overweight with elevated serum alanine aminotransferase activity in the United States
.
Gastroenterology.
2003
;
124
(
1
):
71
79
.

58.

Pramfalk
 
C
,
Pavlides
M
,
Banerjee
R
, et al.  
Sex-specific differences in hepatic fat oxidation and synthesis may explain the higher propensity for NAFLD in men
.
J Clin Endocrinol Metab.
2015
;
100
(
12
):
4425
4433
.

59.

Park
 
SH
,
Jeon
WK
,
Kim
SH
, et al.  
Prevalence and risk factors of non-alcoholic fatty liver disease among Korean adults
.
J Gastroenterol Hepatol.
2006
;
21
(
1 Pt 1
):
138
143
.

60.

Fan
 
JG
,
Zhu
J
,
Li
XJ
, et al.  
Prevalence of and risk factors for fatty liver in a general population of Shanghai, China
.
J Hepatol.
2005
;
43
(
3
):
508
514
.

61.

Florentino
 
GS
,
Cotrim
HP
,
Vilar
CP
,
Florentino
AV
,
Guimarães
GM
,
Barreto
VS
.
Nonalcoholic fatty liver disease in menopausal women
.
Arq Gastroenterol.
2013
;
50
(
3
):
180
185
.

62.

Clark
 
JM
,
Brancati
FL
,
Diehl
AM
.
Nonalcoholic fatty liver disease
.
Gastroenterology.
2002
;
122
(
6
):
1649
1657
.

63.

Völzke
 
H
,
Schwarz
S
,
Baumeister
SE
, et al.  
Menopausal status and hepatic steatosis in a general female population
.
Gut.
2007
;
56
(
4
):
594
595
.

64.

Matsuo
 
K
,
Gualtieri
MR
,
Cahoon
SS
, et al.  
Surgical menopause and increased risk of nonalcoholic fatty liver disease in endometrial cancer
.
Menopause.
2016
;
23
(
2
):
189
196
.

65.

Bruno
 
S
,
Maisonneuve
P
,
Castellana
P
, et al.  
Incidence and risk factors for non-alcoholic steatohepatitis: prospective study of 5408 women enrolled in Italian tamoxifen chemoprevention trial
.
Bmj.
2005
;
330
(
7497
):
932
.

66.

Argo
 
CK
,
Northup
PG
,
Al-Osaimi
AM
,
Caldwell
SH
.
Systematic review of risk factors for fibrosis progression in non-alcoholic steatohepatitis
.
J Hepatol.
2009
;
51
(
2
):
371
379
.

67.

Tapper
 
EB
,
Krajewski
K
,
Lai
M
, et al.  
Simple non-invasive biomarkers of advanced fibrosis in the evaluation of non-alcoholic fatty liver disease
.
Gastroenterol Rep (Oxf).
2014
;
2
(
4
):
276
280
.

68.

Singh
 
DK
,
Sakhuja
P
,
Malhotra
V
,
Gondal
R
,
Sarin
SK
.
Independent predictors of steatohepatitis and fibrosis in Asian Indian patients with non-alcoholic steatohepatitis
.
Dig Dis Sci.
2008
;
53
(
7
):
1967
1976
.

69.

Klair
 
JS
,
Yang
JD
,
Abdelmalek
MF
, et al. ;
Nonalcoholic Steatohepatitis Clinical Research Network
.
A longer duration of estrogen deficiency increases fibrosis risk among postmenopausal women with nonalcoholic fatty liver disease
.
Hepatology.
2016
;
64
(
1
):
85
91
.

70.

Yang
 
JD
,
Abdelmalek
MF
,
Pang
H
, et al.  
Gender and menopause impact severity of fibrosis among patients with nonalcoholic steatohepatitis
.
Hepatology.
2014
;
59
(
4
):
1406
1414
.

71.

Yasui
 
K
,
Hashimoto
E
,
Komorizono
Y
, et al. ;
Japan NASH Study Group, Ministry of Health, Labour, and Welfare of Japan
.
Characteristics of patients with nonalcoholic steatohepatitis who develop hepatocellular carcinoma
.
Clin Gastroenterol Hepatol.
2011
;
9
(
5
):
428
433; quiz e50
.

72.

Yang
 
D
,
Hanna
DL
,
Usher
J
, et al.  
Impact of sex on the survival of patients with hepatocellular carcinoma: a surveillance, epidemiology, and end results analysis
.
Cancer.
2014
;
120
(
23
):
3707
3716
.

73.

Wang
 
S
,
Sun
H
,
Xie
Z
, et al.  
Improved survival of patients with hepatocellular carcinoma and disparities by age, race, and socioeconomic status by decade, 1983-2012
.
Oncotarget.
2016
;
7
(
37
):
59820
59833
.

74.

Ong
 
JP
,
Pitts
A
,
Younossi
ZM
.
Increased overall mortality and liver-related mortality in non-alcoholic fatty liver disease
.
J Hepatol.
2008
;
49
(
4
):
608
612
.

75.

Senaratna
 
CV
,
Perret
JL
,
Lodge
CJ
, et al.  
Prevalence of obstructive sleep apnea in the general population: a systematic review
.
Sleep Med Rev.
2017
;
34
:
70
81
.

76.

Young
 
T
,
Shahar
E
,
Nieto
FJ
, et al. ;
Sleep Heart Health Study Research Group
.
Predictors of sleep-disordered breathing in community-dwelling adults: the Sleep Heart Health Study
.
Arch Intern Med.
2002
;
162
(
8
):
893
900
.

77.

Baguet
 
JP
,
Barone-Rochette
G
,
Tamisier
R
,
Levy
P
,
Pépin
JL
.
Mechanisms of cardiac dysfunction in obstructive sleep apnea
.
Nat Rev Cardiol.
2012
;
9
(
12
):
679
688
.

78.

Tasali
 
E
,
Ip
MS
.
Obstructive sleep apnea and metabolic syndrome: alterations in glucose metabolism and inflammation
.
Proc Am Thorac Soc.
2008
;
5
(
2
):
207
217
.

79.

Fu
 
Y
,
Xia
Y
,
Yi
H
,
Xu
H
,
Guan
J
,
Yin
S
.
Meta-analysis of all-cause and cardiovascular mortality in obstructive sleep apnea with or without continuous positive airway pressure treatment
.
Sleep Breath.
2017
;
21
(
1
):
181
189
.

80.

Jin
 
S
,
Jiang
S
,
Hu
A
.
Association between obstructive sleep apnea and non-alcoholic fatty liver disease: a systematic review and meta-analysis
.
Sleep Breath.
2018
;
22
(
3
):
841
851
.

81.

Aron-Wisnewsky
 
J
,
Clement
K
,
Pépin
JL
.
Nonalcoholic fatty liver disease and obstructive sleep apnea
.
Metabolism.
2016
;
65
(
8
):
1124
1135
.

82.

Denko
 
NC
.
Hypoxia, HIF1 and glucose metabolism in the solid tumour
.
Nat Rev Cancer.
2008
;
8
(
9
):
705
713
.

83.

Regueira
 
T
,
Lepper
PM
,
Brandt
S
, et al.  
Hypoxia inducible factor-1 alpha induction by tumour necrosis factor-alpha, but not by toll-like receptor agonists, modulates cellular respiration in cultured human hepatocytes
.
Liver Int.
2009
;
29
(
10
):
1582
1592
.

84.

Koshikawa
 
N
,
Hayashi
J
,
Nakagawara
A
,
Takenaga
K
.
Reactive oxygen species-generating mitochondrial DNA mutation up-regulates hypoxia-inducible factor-1alpha gene transcription via phosphatidylinositol 3-kinase-Akt/protein kinase C/histone deacetylase pathway
.
J Biol Chem.
2009
;
284
(
48
):
33185
33194
.

85.

Savransky
 
V
,
Bevans
S
,
Nanayakkara
A
, et al. .
Chronic intermittent hypoxia causes hepatitis in a mouse model of diet-induced fatty liver
.
Am J Physiol Gastrointest Liver Physiol.
2007
;
293
(
4
):
G871
G877
.

86.

Li
 
J
,
Thorne
LN
,
Punjabi
NM
, et al.  
Intermittent hypoxia induces hyperlipidemia in lean mice
.
Circ Res.
2005
;
97
(
7
):
698
706
.

87.

Savransky
 
V
,
Nanayakkara
A
,
Vivero
A
, et al.  
Chronic intermittent hypoxia predisposes to liver injury
.
Hepatology.
2007
;
45
(
4
):
1007
1013
.

88.

Nobili
 
V
,
Alisi
A
,
Cutrera
R
, et al.  
Altered gut-liver axis and hepatic adiponectin expression in OSAS: novel mediators of liver injury in paediatric non-alcoholic fatty liver
.
Thorax.
2015
;
70
(
8
):
769
781
.

89.

Reinke
 
C
,
Bevans-Fonti
S
,
Drager
LF
,
Shin
MK
,
Polotsky
VY
.
Effects of different acute hypoxic regimens on tissue oxygen profiles and metabolic outcomes
.
J Appl Physiol (1985).
2011
;
111
(
3
):
881
890
.

90.

Kim
 
D
,
Ahmed
A
,
Kushida
C
.
Continuous positive airway pressure therapy on nonalcoholic fatty liver disease in patients with obstructive sleep apnea
.
J Clin Sleep Med.
2018
;
14
(
8
):
1315
1322
.

91.

Marin
 
JM
,
Carrizo
SJ
,
Vicente
E
,
Agusti
AG
.
Long-term cardiovascular outcomes in men with obstructive sleep apnoea-hypopnoea with or without treatment with continuous positive airway pressure: an observational study
.
Lancet.
2005
;
365
(
9464
):
1046
1053
.

92.

Jullian-Desayes
 
I
,
Joyeux-Faure
M
,
Tamisier
R
, et al.  
Impact of obstructive sleep apnea treatment by continuous positive airway pressure on cardiometabolic biomarkers: a systematic review from sham CPAP randomized controlled trials
.
Sleep Med Rev.
2015
;
21
:
23
38
.

93.

Jullian-Desayes
 
I
,
Tamisier
R
,
Zarski
JP
, et al.  
Impact of effective versus sham continuous positive airway pressure on liver injury in obstructive sleep apnoea: data from randomized trials
.
Respirology.
2016
;
21
(
2
):
378
385
.

94.

Labarca
 
G
,
Cruz
R
,
Jorquera
J
.
Continuous positive airway pressure in patients with obstructive sleep apnea and non-alcoholic steatohepatitis: a systematic review and meta-analysis
.
J Clin Sleep Med.
2018
;
14
(
1
):
133
139
.

95.

Kalafati
 
IP
,
Borsa
D
,
Dimitriou
M
,
Revenas
K
,
Kokkinos
A
,
Dedoussis
GV
.
Dietary patterns and non-alcoholic fatty liver disease in a Greek case-control study
.
Nutrition.
2019
;
61
:
105
110
.

96.

Yang
 
CQ
,
Shu
L
,
Wang
S
, et al.  
Dietary patterns modulate the risk of non-alcoholic fatty liver disease in chinese adults
.
Nutrients.
2015
;
7
(
6
):
4778
4791
.

97.

Liu
 
X
,
Peng
Y
,
Chen
S
,
Sun
Q
.
An observational study on the association between major dietary patterns and non-alcoholic fatty liver disease in Chinese adolescents
.
Medicine (Baltimore).
2018
;
97
(
17
):
e0576
.

98.

Trovato
 
FM
,
Martines
GF
,
Brischetto
D
,
Trovato
G
,
Catalano
D
.
Neglected features of lifestyle: their relevance in non-alcoholic fatty liver disease
.
World J Hepatol.
2016
;
8
(
33
):
1459
1465
.

99.

Oddy
 
WH
,
Herbison
CE
,
Jacoby
P
, et al.  
The Western dietary pattern is prospectively associated with nonalcoholic fatty liver disease in adolescence
.
Am J Gastroenterol.
2013
;
108
(
5
):
778
785
.

100.

Fakhoury-Sayegh
 
N
,
Younes
H
,
Heraoui
GNHA
,
Sayegh
R
.
Nutritional profile and dietary patterns of lebanese non-alcoholic fatty liver disease patients: a case-control study
.
Nutrients.
2017
;
9
(
11
): E124.

101.

Trovato
 
FM
,
Martines
GF
,
Catalano
D
.
Addressing Western dietary pattern in obesity and NAFLD
.
Nutrire.
2018
;
43
(
1
):
11
.

102.

Kechagias
 
S
,
Ernersson
A
,
Dahlqvist
O
, et al. ;
Fast Food Study Group
.
Fast-food-based hyper-alimentation can induce rapid and profound elevation of serum alanine aminotransferase in healthy subjects
.
Gut.
2008
;
57
(
5
):
649
654
.

103.

Wehmeyer
 
MH
,
Zyriax
BC
,
Jagemann
B
, et al.  
Nonalcoholic fatty liver disease is associated with excessive calorie intake rather than a distinctive dietary pattern
.
Medicine (Baltimore).
2016
;
95
(
23
):
e3887
.

104.

Baratta
 
F
,
Pastori
D
,
Polimeni
L
, et al.  
Adherence to mediterranean diet and non-alcoholic fatty liver disease: effect on insulin resistance
.
Am J Gastroenterol.
2017
;
112
(
12
):
1832
1839
.

105.

Chan
 
R
,
Wong
VW
,
Chu
WC
, et al.  
Diet-quality scores and prevalence of nonalcoholic fatty liver disease: a population study using proton-magnetic resonance spectroscopy
.
PLoS One.
2015
;
10
(
9
):
e0139310
.

106.

Aller
 
R
,
Izaola
O
,
de la Fuente
B
,
De Luis Román
DA
.
Mediterranean diet is associated with liver histology in patients with non alcoholic fatty liver disease
.
Nutr Hosp.
2015
;
32
(
6
):
2518
2524
.

107.

Della Corte
 
C
,
Mosca
A
,
Vania
A
,
Alterio
A
,
Iasevoli
S
,
Nobili
V
.
Good adherence to the Mediterranean diet reduces the risk for NASH and diabetes in pediatric patients with obesity: the results of an Italian study
.
Nutrition.
2017
;
39–40
:
8
14
.

108.

Keys
 
A
,
Menotti
A
,
Aravanis
C
, et al.  
The seven countries study: 2289 deaths in 15 years
.
Prev Med.
1984
;
13
(
2
):
141
154
.

109.

Trichopoulou
 
A
,
Costacou
T
,
Bamia
C
,
Trichopoulos
D
.
Adherence to a Mediterranean diet and survival in a Greek population
.
N Engl J Med.
2003
;
348
(
26
):
2599
2608
.

110.

Mitrou
 
PN
,
Kipnis
V
,
Thiébaut
AC
, et al.  
Mediterranean dietary pattern and prediction of all-cause mortality in a US population: results from the NIH-AARP Diet and Health Study
.
Arch Intern Med.
2007
;
167
(
22
):
2461
2468
.

111.

Fraser
 
A
,
Abel
R
,
Lawlor
DA
,
Fraser
D
,
Elhayany
A
.
A modified Mediterranean diet is associated with the greatest reduction in alanine aminotransferase levels in obese type 2 diabetes patients: results of a quasi-randomised controlled trial
.
Diabetologia.
2008
;
51
(
9
):
1616
1622
.

112.

Ryan
 
MC
,
Itsiopoulos
C
,
Thodis
T
, et al.  
The Mediterranean diet improves hepatic steatosis and insulin sensitivity in individuals with non-alcoholic fatty liver disease
.
J Hepatol.
2013
;
59
(
1
):
138
143
.

113.

Bozzetto
 
L
,
Prinster
A
,
Annuzzi
G
, et al.  
Liver fat is reduced by an isoenergetic MUFA diet in a controlled randomized study in type 2 diabetic patients
.
Diabetes Care.
2012
;
35
(
7
):
1429
1435
.

114.

Anania
 
C
,
Perla
FM
,
Olivero
F
,
Pacifico
L
,
Chiesa
C
.
Mediterranean diet and nonalcoholic fatty liver disease
.
World J Gastroenterol.
2018
;
24
(
19
):
2083
2094
.

115.

European Association for the Study of the Liver (EASL), European Association for the Study of Diabetes (EASD), European Association for the Study of Obesity (EASO)
.
EASL–EASD–EASO clinical practice guidelines for the management of non-alcoholic fatty liver disease
.
J Hepatol.
2016
;
64
(
6
):
1388
1402
.

116.

Welsh
 
JA
,
Sharma
AJ
,
Grellinger
L
,
Vos
MB
.
Consumption of added sugars is decreasing in the United States
.
Am J Clin Nutr.
2011
;
94
(
3
):
726
734
.

117.

Ouyang
 
X
,
Cirillo
P
,
Sautin
Y
, et al.  
Fructose consumption as a risk factor for non-alcoholic fatty liver disease
.
J Hepatol.
2008
;
48
(
6
):
993
999
.

118.

Maersk
 
M
,
Belza
A
,
Stødkilde-Jørgensen
H
, et al.  
Sucrose-sweetened beverages increase fat storage in the liver, muscle, and visceral fat depot: a 6-mo randomized intervention study
.
Am J Clin Nutr.
2012
;
95
(
2
):
283
289
.

119.

Abdelmalek
 
MF
,
Suzuki
A
,
Guy
C
, et al. ;
Nonalcoholic Steatohepatitis Clinical Research Network
.
Increased fructose consumption is associated with fibrosis severity in patients with nonalcoholic fatty liver disease
.
Hepatology.
2010
;
51
(
6
):
1961
1971
.

120.

OʼSullivan
 
TA
,
Oddy
WH
,
Bremner
AP
, et al.  
Lower fructose intake may help protect against development of nonalcoholic fatty liver in adolescents with obesity
.
J Pediatr Gastroenterol Nutr.
2014
;
58
(
5
):
624
631
.

121.

Softic
 
S
,
Cohen
DE
,
Kahn
CR
.
Role of dietary fructose and hepatic de novo lipogenesis in fatty liver disease
.
Dig Dis Sci.
2016
;
61
(
5
):
1282
1293
.

122.

Faeh
 
D
,
Minehira
K
,
Schwarz
JM
, et al.  
Effect of fructose overfeeding and fish oil administration on hepatic de novo lipogenesis and insulin sensitivity in healthy men
.
Diabetes.
2005
;
54
(
7
):
1907
1913
.

123.

Stanhope
 
KL
,
Schwarz
JM
,
Keim
NL
, et al.  
Consuming fructose-sweetened, not glucose-sweetened, beverages increases visceral adiposity and lipids and decreases insulin sensitivity in overweight/obese humans
.
J Clin Invest.
2009
;
119
(
5
):
1322
1334
.

124.

Cox
 
CL
,
Stanhope
KL
,
Schwarz
JM
, et al.  
Consumption of fructose-sweetened beverages for 10 weeks reduces net fat oxidation and energy expenditure in overweight/obese men and women
.
Eur J Clin Nutr.
2012
;
66
(
2
):
201
208
.

125.

Jensen
 
T
,
Abdelmalek
MF
,
Sullivan
S
, et al.  
Fructose and sugar: a major mediator of non-alcoholic fatty liver disease
.
J Hepatol.
2018
;
68
(
5
):
1063
1075
.

126.

Mäenpää
 
PH
,
Raivio
KO
,
Kekomäki
MP
.
Liver adenine nucleotides: fructose-induced depletion and its effect on protein synthesis
.
Science.
1968
;
161
(
3847
):
1253
1254
.

127.

Lanaspa
 
MA
,
Sanchez-Lozada
LG
,
Choi
YJ
, et al.  
Uric acid induces hepatic steatosis by generation of mitochondrial oxidative stress: potential role in fructose-dependent and -independent fatty liver
.
J Biol Chem.
2012
;
287
(
48
):
40732
40744
.

128.

Choi
 
YJ
,
Shin
HS
,
Choi
HS
, et al.  
Uric acid induces fat accumulation via generation of endoplasmic reticulum stress and SREBP-1c activation in hepatocytes
.
Lab Invest.
2014
;
94
(
10
):
1114
1125
.

129.

Perheentupa
 
J
,
Raivio
K
.
Fructose-induced hyperuricaemia
.
Lancet.
1967
;
2
(
7515
):
528
531
.

130.

Zhou
 
Y
,
Wei
F
,
Fan
Y
.
High serum uric acid and risk of nonalcoholic fatty liver disease: a systematic review and meta-analysis
.
Clin Biochem.
2016
;
49
(
7-8
):
636
642
.

131.

Liu
 
Z
,
Que
S
,
Zhou
L
,
Zheng
S
.
Dose-response relationship of serum uric acid with metabolic syndrome and non-alcoholic fatty liver disease incidence: a meta-analysis of prospective studies
.
Sci Rep.
2015
;
5
:
14325
.

132.

Mitchell
 
DC
,
Knight
CA
,
Hockenberry
J
,
Teplansky
R
,
Hartman
TJ
.
Beverage caffeine intakes in the U.S
.
Food Chem Toxicol.
2014
;
63
:
136
142
.

133.

Saab
 
S
,
Mallam
D
,
Cox
GA
2nd
,
Tong
MJ
.
Impact of coffee on liver diseases: a systematic review
.
Liver Int.
2014
;
34
(
4
):
495
504
.

134.

Birerdinc
 
A
,
Stepanova
M
,
Pawloski
L
,
Younossi
ZM
.
Caffeine is protective in patients with non-alcoholic fatty liver disease
.
Aliment Pharmacol Ther.
2012
;
35
(
1
):
76
82
.

135.

Catalano
 
D
,
Martines
GF
,
Tonzuso
A
,
Pirri
C
,
Trovato
FM
,
Trovato
GM
.
Protective role of coffee in non-alcoholic fatty liver disease (NAFLD)
.
Dig Dis Sci.
2010
;
55
(
11
):
3200
3206
.

136.

Gutiérrez-Grobe
 
Y
,
Chávez-Tapia
N
,
Sánchez-Valle
V
, et al.  
High coffee intake is associated with lower grade nonalcoholic fatty liver disease: the role of peripheral antioxidant activity
.
Ann Hepatol.
11
(
3
):
350
355
.

137.

Veronese
 
N
,
Notarnicola
M
,
Cisternino
A
, et al.  
Coffee intake and liver steatosis: a population study in a Mediterranean Area
.
Nutrients.
2018
;
10
(
1
):
89
.

138.

Molloy
 
JW
,
Calcagno
CJ
,
Williams
CD
,
Jones
FJ
,
Torres
DM
,
Harrison
SA
.
Association of coffee and caffeine consumption with fatty liver disease, nonalcoholic steatohepatitis, and degree of hepatic fibrosis
.
Hepatology.
2012
;
55
(
2
):
429
436
.

139.

Anty
 
R
,
Marjoux
S
,
Iannelli
A
, et al.  
Regular coffee but not espresso drinking is protective against fibrosis in a cohort mainly composed of morbidly obese European women with NAFLD undergoing bariatric surgery
.
J Hepatol.
2012
;
57
(
5
):
1090
1096
.

140.

Donnelly
 
KL
,
Smith
CI
,
Schwarzenberg
SJ
,
Jessurun
J
,
Boldt
MD
,
Parks
EJ
.
Sources of fatty acids stored in liver and secreted via lipoproteins in patients with nonalcoholic fatty liver disease
.
J Clin Invest.
2005
;
115
(
5
):
1343
1351
.

141.

Armstrong
 
MJ
,
Hazlehurst
JM
,
Hull
D
, et al.  
Abdominal subcutaneous adipose tissue insulin resistance and lipolysis in patients with non-alcoholic steatohepatitis
.
Diabetes Obes Metab.
2014
;
16
(
7
):
651
660
.

142.

Musso
 
G
,
Cassader
M
,
De Michieli
F
,
Rosina
F
,
Orlandi
F
,
Gambino
R
.
Nonalcoholic steatohepatitis versus steatosis: adipose tissue insulin resistance and dysfunctional response to fat ingestion predict liver injury and altered glucose and lipoprotein metabolism
.
Hepatology.
2012
;
56
(
3
):
933
942
.

143.

Lomonaco
 
R
,
Ortiz-Lopez
C
,
Orsak
B
, et al.  
Effect of adipose tissue insulin resistance on metabolic parameters and liver histology in obese patients with nonalcoholic fatty liver disease
.
Hepatology.
2012
;
55
(
5
):
1389
1397
.

144.

Ortiz-Lopez
 
C
,
Lomonaco
R
,
Orsak
B
, et al.  
Prevalence of prediabetes and diabetes and metabolic profile of patients with nonalcoholic fatty liver disease (NAFLD)
.
Diabetes Care.
2012
;
35
(
4
):
873
878
.

145.

Magkos
 
F
,
Fraterrigo
G
,
Yoshino
J
, et al.  
Effects of moderate and subsequent progressive weight loss on metabolic function and adipose tissue biology in humans with obesity
.
Cell Metab.
2016
;
23
(
4
):
591
601
.

146.

Gastaldelli
 
A
,
Harrison
SA
,
Belfort-Aguilar
R
, et al.  
Importance of changes in adipose tissue insulin resistance to histological response during thiazolidinedione treatment of patients with nonalcoholic steatohepatitis
.
Hepatology.
2009
;
50
(
4
):
1087
1093
.

147.

Byrne
 
CD
.
Dorothy Hodgkin Lecture 2012: non-alcoholic fatty liver disease, insulin resistance and ectopic fat: a new problem in diabetes management
.
Diabet Med.
2012
;
29
(
9
):
1098
1107
.

148.

Safar Zadeh
 
E
,
Lungu
AO
,
Cochran
EK
, et al.  
The liver diseases of lipodystrophy: the long-term effect of leptin treatment
.
J Hepatol.
2013
;
59
(
1
):
131
137
.

149.

Chang
 
Y
,
Jung
HS
,
Cho
J
, et al.  
Metabolically healthy obesity and the development of nonalcoholic fatty liver disease
.
Am J Gastroenterol.
2016
;
111
(
8
):
1133
1140
.

150.

Buechler
 
C
,
Wanninger
J
,
Neumeier
M
.
Adiponectin, a key adipokine in obesity related liver diseases
.
World J Gastroenterol.
2011
;
17
(
23
):
2801
2811
.

151.

Yoon
 
MJ
,
Lee
GY
,
Chung
JJ
,
Ahn
YH
,
Hong
SH
,
Kim
JB
.
Adiponectin increases fatty acid oxidation in skeletal muscle cells by sequential activation of AMP-activated protein kinase, p38 mitogen-activated protein kinase, and peroxisome proliferator-activated receptor alpha
.
Diabetes.
2006
;
55
(
9
):
2562
2570
.

152.

Liu
 
Q
,
Yuan
B
,
Lo
KA
,
Patterson
HC
,
Sun
Y
,
Lodish
HF
.
Adiponectin regulates expression of hepatic genes critical for glucose and lipid metabolism
.
Proc Natl Acad Sci U S A.
2012
;
109
(
36
):
14568
14573
.

153.

Matsumoto
 
H
,
Tamura
S
,
Kamada
Y
, et al.  
Adiponectin deficiency exacerbates lipopolysaccharide/D-galactosamine-induced liver injury in mice
.
World J Gastroenterol.
2006
;
12
(
21
):
3352
3358
.

154.

Ouchi
 
N
,
Walsh
K
.
Adiponectin as an anti-inflammatory factor
.
Clin Chim Acta.
2007
;
380
(
1-2
):
24
30
.

155.

Ding
 
X
,
Saxena
NK
,
Lin
S
, et al.  
The roles of leptin and adiponectin: a novel paradigm in adipocytokine regulation of liver fibrosis and stellate cell biology
.
Am J Pathol.
2005
;
166
(
6
):
1655
1669
.

156.

Adachi
 
M
,
Brenner
DA
.
High molecular weight adiponectin inhibits proliferation of hepatic stellate cells via activation of adenosine monophosphate-activated protein kinase
.
Hepatology.
2008
;
47
(
2
):
677
685
.

157.

Kumar
 
P
,
Raeman
R
,
Chopyk
DM
, et al.  
Adiponectin inhibits hepatic stellate cell activation by targeting the PTEN/AKT pathway
.
Biochim Biophys Acta Mol Basis Dis.
2018
;
1864
(
10
):
3537
3545
.

158.

Pagano
 
C
,
Soardo
G
,
Esposito
W
, et al.  
Plasma adiponectin is decreased in nonalcoholic fatty liver disease
.
Eur J Endocrinol.
2005
;
152
(
1
):
113
118
.

159.

Musso
 
G
,
Gambino
R
,
Biroli
G
, et al.  
Hypoadiponectinemia predicts the severity of hepatic fibrosis and pancreatic Beta-cell dysfunction in nondiabetic nonobese patients with nonalcoholic steatohepatitis
.
Am J Gastroenterol.
2005
;
100
(
11
):
2438
2446
.

160.

Jamali
 
R
,
Razavizade
M
,
Arj
A
,
Aarabi
MH
.
Serum adipokines might predict liver histology findings in non-alcoholic fatty liver disease
.
World J Gastroenterol.
2016
;
22
(
21
):
5096
5103
.

161.

Cusi
 
K
,
Orsak
B
,
Bril
F
, et al.  
Long-term pioglitazone treatment for patients with nonalcoholic steatohepatitis and prediabetes or type 2 diabetes mellitus: a randomized trial
.
Ann Intern Med.
2016
;
165
(
5
):
305
315
.

162.

Xu
 
A
,
Wang
Y
,
Keshaw
H
,
Xu
LY
,
Lam
KS
,
Cooper
GJ
.
The fat-derived hormone adiponectin alleviates alcoholic and nonalcoholic fatty liver diseases in mice
.
J Clin Invest.
2003
;
112
(
1
):
91
100
.

163.

Lambert
 
JE
,
Ramos-Roman
MA
,
Browning
JD
,
Parks
EJ
.
Increased de novo lipogenesis is a distinct characteristic of individuals with nonalcoholic fatty liver disease
.
Gastroenterology.
2014
;
146
(
3
):
726
735
.

164.

Diraison
 
F
,
Moulin
P
,
Beylot
M
.
Contribution of hepatic de novo lipogenesis and reesterification of plasma non esterified fatty acids to plasma triglyceride synthesis during non-alcoholic fatty liver disease
.
Diabetes Metab.
2003
;
29
(
5
):
478
485
.

165.

Dorn
 
C
,
Riener
MO
,
Kirovski
G
, et al.  
Expression of fatty acid synthase in nonalcoholic fatty liver disease
.
Int J Clin Exp Pathol.
2010
;
3
(
5
):
505
514
.

166.

Mitsuyoshi
 
H
,
Yasui
K
,
Harano
Y
, et al.  
Analysis of hepatic genes involved in the metabolism of fatty acids and iron in nonalcoholic fatty liver disease
.
Hepatol Res.
2009
;
39
(
4
):
366
373
.

167.

Ferré
 
P
,
Foufelle
F
.
Hepatic steatosis: a role for de novo lipogenesis and the transcription factor SREBP-1c
.
Diabetes Obes Metab.
2010
;
12
(Suppl 2)
:
83
92
.

168.

Brown
 
MS
,
Goldstein
JL
.
Selective versus total insulin resistance: a pathogenic paradox
.
Cell Metab.
2008
;
7
(
2
):
95
96
.

169.

Schwarz
 
JM
,
Linfoot
P
,
Dare
D
,
Aghajanian
K
.
Hepatic de novo lipogenesis in normoinsulinemic and hyperinsulinemic subjects consuming high-fat, low-carbohydrate and low-fat, high-carbohydrate isoenergetic diets
.
Am J Clin Nutr.
2003
;
77
(
1
):
43
50
.

170.

Shimomura
 
I
,
Matsuda
M
,
Hammer
RE
,
Bashmakov
Y
,
Brown
MS
,
Goldstein
JL
.
Decreased IRS-2 and increased SREBP-1c lead to mixed insulin resistance and sensitivity in livers of lipodystrophic and ob/ob mice
.
Mol Cell.
2000
;
6
(
1
):
77
86
.

171.

Pajvani
 
UB
,
Qiang
L
,
Kangsamaksin
T
,
Kitajewski
J
,
Ginsberg
HN
,
Accili
D
.
Inhibition of Notch uncouples Akt activation from hepatic lipid accumulation by decreasing mTorc1 stability
.
Nat Med.
2013
;
19
(
8
):
1054
1060
.

172.

Wu
 
X
,
Chen
K
,
Williams
KJ
.
The role of pathway-selective insulin resistance and responsiveness in diabetic dyslipoproteinemia
.
Curr Opin Lipidol.
2012
;
23
(
4
):
334
344
.

173.

Vatner
 
DF
,
Majumdar
SK
,
Kumashiro
N
, et al.  
Insulin-independent regulation of hepatic triglyceride synthesis by fatty acids
.
Proc Natl Acad Sci U S A.
2015
;
112
(
4
):
1143
1148
.

174.

Sevastianova
 
K
,
Santos
A
,
Kotronen
A
, et al.  
Effect of short-term carbohydrate overfeeding and long-term weight loss on liver fat in overweight humans
.
Am J Clin Nutr.
2012
;
96
(
4
):
727
734
.

175.

Flannery
 
C
,
Dufour
S
,
Rabøl
R
,
Shulman
GI
,
Petersen
KF
.
Skeletal muscle insulin resistance promotes increased hepatic de novo lipogenesis, hyperlipidemia, and hepatic steatosis in the elderly
.
Diabetes.
2012
;
61
(
11
):
2711
2717
.

176.

DeFronzo
 
RA
,
Tripathy
D
.
Skeletal muscle insulin resistance is the primary defect in type 2 diabetes
.
Diabetes Care.
2009
;
32
(Suppl 2)
:
S157
S163
.

177.

Kim
 
G
,
Lee
SE
,
Lee
YB
, et al.  
Relationship between relative skeletal muscle mass and nonalcoholic fatty liver disease: a 7-year longitudinal study
.
Hepatology.
2018
;
68
(
5
):
1755
1768
.

178.

Cusi
 
K
.
Role of obesity and lipotoxicity in the development of nonalcoholic steatohepatitis: pathophysiology and clinical implications
.
Gastroenterology.
2012
;
142
(
4
):
711
725.e6
.

179.

Petersen
 
MC
,
Shulman
GI
.
Roles of diacylglycerols and ceramides in hepatic insulin resistance
.
Trends Pharmacol Sci.
2017
;
38
(
7
):
649
665
.

180.

Puri
 
P
,
Baillie
RA
,
Wiest
MM
, et al.  
A lipidomic analysis of nonalcoholic fatty liver disease
.
Hepatology.
2007
;
46
(
4
):
1081
1090
.

181.

Neuschwander-Tetri
 
BA
.
Hepatic lipotoxicity and the pathogenesis of nonalcoholic steatohepatitis: the central role of nontriglyceride fatty acid metabolites
.
Hepatology.
2010
;
52
(
2
):
774
788
.

182.

Crespo
 
J
,
Cayón
A
,
Fernández-Gil
P
, et al.  
Gene expression of tumor necrosis factor alpha and TNF-receptors, p55 and p75, in nonalcoholic steatohepatitis patients
.
Hepatology.
2001
;
34
(
6
):
1158
1163
.

183.

Cai
 
D
,
Yuan
M
,
Frantz
DF
, et al.  
Local and systemic insulin resistance resulting from hepatic activation of IKK-beta and NF-kappaB
.
Nat Med.
2005
;
11
(
2
):
183
190
.

184.

Ribeiro
 
PS
,
Cortez-Pinto
H
,
Solá
S
, et al.  
Hepatocyte apoptosis, expression of death receptors, and activation of NF-kappaB in the liver of nonalcoholic and alcoholic steatohepatitis patients
.
Am J Gastroenterol.
2004
;
99
(
9
):
1708
1717
.

185.

Arrese
 
M
,
Cabrera
D
,
Kalergis
AM
,
Feldstein
AE
.
Innate immunity and inflammation in NAFLD/NASH
.
Dig Dis Sci.
2016
;
61
(
5
):
1294
1303
.

186.

Petrasek
 
J
,
Csak
T
,
Szabo
G
.
Toll-like receptors in liver disease
.
Adv Clin Chem.
2013
;
59
:
155
201
.

187.

Miura
 
K
,
Kodama
Y
,
Inokuchi
S
, et al.  
Toll-like receptor 9 promotes steatohepatitis by induction of interleukin-1beta in mice
.
Gastroenterology.
2010
;
139
(
1
):
323
334.e7
.

188.

Szabo
 
G
,
Petrasek
J
.
Inflammasome activation and function in liver disease
.
Nat Rev Gastroenterol Hepatol.
2015
;
12
(
7
):
387
400
.

189.

Wree
 
A
,
McGeough
MD
,
Peña
CA
, et al.  
NLRP3 inflammasome activation is required for fibrosis development in NAFLD
.
J Mol Med (Berl).
2014
;
92
(
10
):
1069
1082
.

190.

Csak
 
T
,
Ganz
M
,
Pespisa
J
,
Kodys
K
,
Dolganiuc
A
,
Szabo
G
.
Fatty acid and endotoxin activate inflammasomes in mouse hepatocytes that release danger signals to stimulate immune cells
.
Hepatology.
2011
;
54
(
1
):
133
144
.

191.

Rensen
 
SS
,
Slaats
Y
,
Nijhuis
J
, et al.  
Increased hepatic myeloperoxidase activity in obese subjects with nonalcoholic steatohepatitis
.
Am J Pathol.
2009
;
175
(
4
):
1473
1482
.

192.

Rensen
 
SS
,
Bieghs
V
,
Xanthoulea
S
, et al.  
Neutrophil-derived myeloperoxidase aggravates non-alcoholic steatohepatitis in low-density lipoprotein receptor-deficient mice
.
Plos One.
2012
;
7
(
12
):
e52411
.

193.

Talukdar
 
S
,
Oh
DY
,
Bandyopadhyay
G
, et al.  
Neutrophils mediate insulin resistance in mice fed a high-fat diet through secreted elastase
.
Nat Med.
2012
;
18
(
9
):
1407
1412
.

194.

Ibusuki
 
R
,
Uto
H
,
Arima
S
, et al.  
Transgenic expression of human neutrophil peptide-1 enhances hepatic fibrosis in mice fed a choline-deficient, L-amino acid-defined diet
.
Liver Int.
2013
;
33
(
10
):
1549
1556
.

195.

Sojka
 
DK
,
Tian
Z
,
Yokoyama
WM
.
Tissue-resident natural killer cells and their potential diversity
.
Semin Immunol.
2014
;
26
(
2
):
127
131
.

196.

Tian
 
Z
,
Chen
Y
,
Gao
B
.
Natural killer cells in liver disease
.
Hepatology.
2013
;
57
(
4
):
1654
1662
.

197.

Ganz
 
M
,
Szabo
G
.
Immune and inflammatory pathways in NASH
.
Hepatol Int.
2013
;
7
(Suppl 2)
:
771
781
.

198.

Henning
 
JR
,
Graffeo
CS
,
Rehman
A
, et al.  
Dendritic cells limit fibroinflammatory injury in nonalcoholic steatohepatitis in mice
.
Hepatology.
2013
;
58
(
2
):
589
602
.

199.

Weston
 
CJ
,
Shepherd
EL
,
Claridge
LC
, et al.  
Vascular adhesion protein-1 promotes liver inflammation and drives hepatic fibrosis
.
J Clin Invest.
2015
;
125
(
2
):
501
520
.

200.

Sutti
 
S
,
Jindal
A
,
Locatelli
I
, et al.  
Adaptive immune responses triggered by oxidative stress contribute to hepatic inflammation in NASH
.
Hepatology.
2014
;
59
(
3
):
886
897
.

201.

Sell
 
H
,
Habich
C
,
Eckel
J
.
Adaptive immunity in obesity and insulin resistance
.
Nat Rev Endocrinol.
2012
;
8
(
12
):
709
716
.

202.

Murphy
 
MP
.
How mitochondria produce reactive oxygen species
.
Biochem J.
2009
;
417
(
1
):
1
13
.

203.

Koliaki
 
C
,
Szendroedi
J
,
Kaul
K
, et al.  
Adaptation of hepatic mitochondrial function in humans with non-alcoholic fatty liver is lost in steatohepatitis
.
Cell Metab.
2015
;
21
(
5
):
739
746
.

204.

Serviddio
 
G
,
Bellanti
F
,
Tamborra
R
, et al.  
Uncoupling protein-2 (UCP2) induces mitochondrial proton leak and increases susceptibility of non-alcoholic steatohepatitis (NASH) liver to ischaemia-reperfusion injury
.
Gut.
2008
;
57
(
7
):
957
965
.

205.

Cortez-Pinto
 
H
,
Chatham
J
,
Chacko
VP
,
Arnold
C
,
Rashid
A
,
Diehl
AM
.
Alterations in liver ATP homeostasis in human nonalcoholic steatohepatitis: a pilot study
.
Jama.
1999
;
282
(
17
):
1659
1664
.

206.

Leclercq
 
IA
,
Farrell
GC
,
Field
J
,
Bell
DR
,
Gonzalez
FJ
,
Robertson
GR
.
CYP2E1 and CYP4A as microsomal catalysts of lipid peroxides in murine nonalcoholic steatohepatitis
.
J Clin Invest.
2000
;
105
(
8
):
1067
1075
.

207.

Marí
 
M
,
Caballero
F
,
Colell
A
, et al.  
Mitochondrial free cholesterol loading sensitizes to TNF- and Fas-mediated steatohepatitis
.
Cell Metab.
2006
;
4
(
3
):
185
198
.

208.

Gusdon
 
AM
,
Song
KX
,
Qu
S
.
Nonalcoholic fatty liver disease: pathogenesis and therapeutics from a mitochondria-centric perspective
.
Oxid Med Cell Longev.
2014
;
2014
:
637027
.

209.

Madrigal-Matute
 
J
,
Cuervo
AM
.
Regulation of liver metabolism by autophagy
.
Gastroenterology.
2016
;
150
(
2
):
328
339
.

210.

González-Rodríguez
 
A
,
Mayoral
R
,
Agra
N
, et al.  
Impaired autophagic flux is associated with increased endoplasmic reticulum stress during the development of NAFLD
.
Cell Death Dis.
2014
;
5
:
e1179
.

211.

Cariou
 
B
,
van Harmelen
K
,
Duran-Sandoval
D
, et al.  
The farnesoid X receptor modulates adiposity and peripheral insulin sensitivity in mice
.
J Biol Chem.
2006
;
281
(
16
):
11039
11049
.

212.

Jiang
 
C
,
Xie
C
,
Li
F
, et al.  
Intestinal farnesoid X receptor signaling promotes nonalcoholic fatty liver disease
.
J Clin Invest.
2015;125(1):386–402
.

213.

Chávez-Talavera
 
O
,
Haas
J
,
Grzych
G
,
Tailleux
A
,
Staels
B
.
Bile acid alterations in nonalcoholic fatty liver disease, obesity, insulin resistance and type 2 diabetes: what do the human studies tell?
Curr Opin Lipidol.
2019;30(3):244–254
.

214.

Puri
 
P
,
Daita
K
,
Joyce
A
, et al.  
The presence and severity of nonalcoholic steatohepatitis is associated with specific changes in circulating bile acids
.
Hepatology.
2018
;
67
(
2
):
534
548
.

215.

Legry
 
V
,
Francque
S
,
Haas
JT
, et al.  
Bile acid alterations are associated with insulin resistance, but not with NASH, in obese subjects
.
J Clin Endocrinol Metab.
2017
;
102
(
10
):
3783
3794
.

216.

Jiao
 
N
,
Baker
SS
,
Chapa-rodriguez
A
, et al.  
Suppressed hepatic bile acid signalling despite elevated production of primary and secondary bile acids in NAFLD
. Gut.
2018;67(10):1881–1891
.

217.

Turnbaugh
 
PJ
,
Ley
RE
,
Mahowald
MA
,
Magrini
V
,
Mardis
ER
,
Gordon
JI
.
An obesity-associated gut microbiome with increased capacity for energy harvest
.
Nature.
2006
;
444
(
7122
):
1027
1031
.

218.

Vrieze
 
A
,
Van Nood
E
,
Holleman
F
, et al.  
Transfer of intestinal microbiota from lean donors increases insulin sensitivity in individuals with metabolic syndrome
.
Gastroenterology.
2012
;
143
(
4
):
913
916.e7
.

219.

Henao-Mejia
 
J
,
Elinav
E
,
Jin
C
, et al.  
Inflammasome-mediated dysbiosis regulates progression of NAFLD and obesity
.
Nature.
2012
;
482
(
7384
):
179
185
.

220.

Leung
 
C
,
Rivera
L
,
Furness
JB
,
Angus
PW
.
The role of the gut microbiota in NAFLD
.
Nat Rev Gastroenterol Hepatol.
2016
;
13
(
7
):
412
425
.

221.

Brandl
 
K
,
Schnabl
B
.
Intestinal microbiota and nonalcoholic steatohepatitis
.
Curr Opin Gastroenterol.
2017
;
33
(
3
):
128
133
.

222.

Mouzaki
 
M
,
Comelli
EM
,
Arendt
BM
, et al.  
Intestinal microbiota in patients with nonalcoholic fatty liver disease
.
Hepatology.
2013
;
58
(
1
):
120
127
.

223.

Jumpertz
 
R
,
Le
DS
,
Turnbaugh
PJ
,
Trinidad
C
,
Bogardus
C
,
Gordon
JI
,
Krakoff
J
.
Energy-balance studies reveal associations between gut microbes, caloric load, and nutrient absorption in humans
.
Am J Clin Nutr.
2011
;
94
(
1
):
58
65
.

224.

Boursier
 
J
,
Mueller
O
,
Barret
M
, et al.  
The severity of nonalcoholic fatty liver disease is associated with gut dysbiosis and shift in the metabolic function of the gut microbiota
.
Hepatology.
2016
;
63
(
3
):
764
775
.

225.

Loomba
 
R
,
Seguritan
V
,
Li
W
, et al.  
Gut microbiome-based metagenomic signature for non-invasive detection of advanced fibrosis in human nonalcoholic fatty liver disease
.
Cell Metab.
2017
;
25
(
5
):
1054
1062.e5
.

226.

Yoshimoto
 
S
,
Loo
TM
,
Atarashi
K
, et al.  
Obesity-induced gut microbial metabolite promotes liver cancer through senescence secretome
.
Nature.
2013
;
499
(
7456
):
97
101
.

227.

Brandi
 
G
,
De Lorenzo
S
,
Candela
M
, et al.  
Microbiota, NASH, HCC and the potential role of probiotics
.
Carcinogenesis.
2017
;
38
(
3
):
231
240
.

228.

Wigg
 
AJ
,
Roberts-Thomson
IC
,
Dymock
RB
,
McCarthy
PJ
,
Grose
RH
,
Cummins
AG
.
The role of small intestinal bacterial overgrowth, intestinal permeability, endotoxaemia, and tumour necrosis factor alpha in the pathogenesis of non-alcoholic steatohepatitis
.
Gut.
2001
;
48
(
2
):
206
211
.

229.

Shanab
 
AA
,
Scully
P
,
Crosbie
O
, et al.  
Small intestinal bacterial overgrowth in nonalcoholic steatohepatitis: association with toll-like receptor 4 expression and plasma levels of interleukin 8
.
Dig Dis Sci.
2011
;
56
(
5
):
1524
1534
.

230.

Miele
 
L
,
Valenza
V
,
La Torre
G
, et al.  
Increased intestinal permeability and tight junction alterations in nonalcoholic fatty liver disease
.
Hepatology.
2009
;
49
(
6
):
1877
1887
.

231.

Zhu
 
L
,
Baker
SS
,
Gill
C
, et al.  
Characterization of gut microbiomes in nonalcoholic steatohepatitis (NASH) patients: a connection between endogenous alcohol and NASH
.
Hepatology.
2013
;
57
(
2
):
601
609
.

232.

Spencer
 
MD
,
Hamp
TJ
,
Reid
RW
,
Fischer
LM
,
Zeisel
SH
,
Fodor
AA
.
Association between composition of the human gastrointestinal microbiome and development of fatty liver with choline deficiency
.
Gastroenterology.
2011
;
140
(
3
):
976
986
.

233.

Wahlström
 
A
,
Sayin
SI
,
Marschall
HU
,
Bäckhed
F
.
Intestinal crosstalk between bile acids and microbiota and its impact on host metabolism
.
Cell Metab.
2016
;
24
(
1
):
41
50
.

234.

Xanthakos
 
S
,
Miles
L
,
Bucuvalas
J
,
Daniels
S
,
Garcia
V
,
Inge
T
.
Histologic spectrum of nonalcoholic fatty liver disease in morbidly obese adolescents
.
Clin Gastroenterol Hepatol.
2006
;
4
(
2
):
226
232
.

235.

Schwimmer
 
JB
,
Celedon
MA
,
Lavine
JE
, et al.  
Heritability of nonalcoholic fatty liver disease
.
Gastroenterology.
2009
;
136
(
5
):
1585
1592
.

236.

Loomba
 
R
,
Schork
N
,
Chen
CH
, et al. ;
Genetics of NAFLD in Twins Consortium
.
Heritability of hepatic fibrosis and steatosis based on a prospective twin study
.
Gastroenterology.
2015
;
149
(
7
):
1784
1793
.

237.

Dai
 
G
,
Liu
P
,
Li
X
,
Zhou
X
,
He
S
.
Association between PNPLA3 rs738409 polymorphism and nonalcoholic fatty liver disease (NAFLD) susceptibility and severity: a meta-analysis
.
Medicine (Baltimore).
2019
;
98
(
7
):
e14324
.

238.

Xu
 
R
,
Tao
A
,
Zhang
S
,
Deng
Y
,
Chen
G
.
Association between patatin-like phospholipase domain containing 3 gene (PNPLA3) polymorphisms and nonalcoholic fatty liver disease: a HuGE review and meta-analysis
.
Sci Rep.
2015
;
5
:
9284
.

239.

Shen
 
JH
,
Li
YL
,
Li
D
,
Wang
NN
,
Jing
L
,
Huang
YH
.
The rs738409 (I148M) variant of the PNPLA3 gene and cirrhosis: a meta-analysis
.
J Lipid Res.
2015
;
56
(
1
):
167
175
.

240.

Huang
 
Z
,
Guo
X
,
Zhang
G
,
Liang
L
,
Nong
B
.
Correlation between PNPLA3 rs738409 polymorphism and hepatocellular carcinoma: a meta-analysis of 10,330 subjects
.
Int J Biol Markers.
2019
:
172460081881247
.

241.

Chen
 
W
,
Chang
B
,
Li
L
,
Chan
L
.
Patatin-like phospholipase domain-containing 3/adiponutrin deficiency in mice is not associated with fatty liver disease
.
Hepatology.
2010
;
52
(
3
):
1134
1142
.

242.

BasuRay
 
S
,
Smagris
E
,
Cohen
JC
,
Hobbs
HH
.
The PNPLA3 variant associated with fatty liver disease (I148M) accumulates on lipid droplets by evading ubiquitylation
.
Hepatology.
2017
;
66
(
4
):
1111
1124
.

243.

Pingitore
 
P
,
Dongiovanni
P
,
Motta
BM
, et al.  
PNPLA3 overexpression results in reduction of proteins predisposing to fibrosis
.
Hum Mol Genet.
2016
;
25
(
23
):
5212
5222
.

244.

Donati
 
B
,
Motta
BM
,
Pingitore
P
, et al.  
The rs2294918 E434K variant modulates patatin-like phospholipase domain-containing 3 expression and liver damage
.
Hepatology.
2016
;
63
(
3
):
787
798
.

245.

Valenti
 
L
,
Dongiovanni
P
.
Mutant PNPLA3 I148M protein as pharmacological target for liver disease
.
Hepatology.
2017
;
66
(
4
):
1026
1028
.

246.

Kim
 
DS
,
Jackson
AU
,
Li
YK
, et al. ;
FinMetSeq Investigators
.
Novel association of TM6SF2 rs58542926 genotype with increased serum tyrosine levels and decreased apoB-100 particles in Finns
.
J Lipid Res.
2017
;
58
(
7
):
1471
1481
.

247.

Pirola
 
CJ
,
Sookoian
S
.
The dual and opposite role of the TM6SF2-rs58542926 variant in protecting against cardiovascular disease and conferring risk for nonalcoholic fatty liver: a meta-analysis
.
Hepatology.
2015
;
62
(
6
):
1742
1756
.

248.

Luukkonen
 
PK
,
Zhou
Y
,
Hyötyläinen
T
, et al.  
The MBOAT7 variant rs641738 alters hepatic phosphatidylinositols and increases severity of non-alcoholic fatty liver disease in humans
.
J Hepatol.
2016
;
65
(
6
):
1263
1265
.

249.

Mancina
 
RM
,
Dongiovanni
P
,
Petta
S
, et al.  
The MBOAT7-TMC4 variant rs641738 increases risk of nonalcoholic fatty liver disease in individuals of european descent
.
Gastroenterology.
2016
;
150
(
5
):
1219
1230.e6
.

250.

Donati
 
B
,
Dongiovanni
P
,
Romeo
S
, et al.  
MBOAT7 rs641738 variant and hepatocellular carcinoma in non-cirrhotic individuals
.
Sci Rep.
2017
;
7
(
1
):
4492
.

251.

Santoro
 
N
,
Zhang
CK
,
Zhao
H
, et al.  
Variant in the glucokinase regulatory protein (GCKR) gene is associated with fatty liver in obese children and adolescents
.
Hepatology.
2012
;
55
(
3
):
781
789
.

252.

Hernaez
 
R
,
McLean
J
,
Lazo
M
, et al. ;
Genetics of Obesity-Related Liver Disease (GOLD) Consortium
.
Association between variants in or near PNPLA3, GCKR, and PPP1R3B with ultrasound-defined steatosis based on data from the third National Health and Nutrition Examination Survey
.
Clin Gastroenterol Hepatol.
2013
;
11
(
9
):
1183
1190.e2
.

253.

Stender
 
S
,
Kozlitina
J
,
Nordestgaard
BG
,
Tybjærg-Hansen
A
,
Hobbs
HH
,
Cohen
JC
.
Adiposity amplifies the genetic risk of fatty liver disease conferred by multiple loci
.
Nat Genet.
2017
;
49
(
6
):
842
847
.

254.

Davis
 
JN
,
KA
,
Walker
RW
, et al.  
Increased hepatic fat in overweight Hispanic youth influenced by interaction between genetic variation in PNPLA3 and high dietary carbohydrate and sugar consumption
.
Am J Clin Nutr.
2010
;
92
(
6
):
1522
1527
.

255.

Nobili
 
V
,
Liccardo
D
,
Bedogni
G
, et al.  
Influence of dietary pattern, physical activity, and I148M PNPLA3 on steatosis severity in at-risk adolescents
.
Genes Nutr.
2014
;
9
(
3
):
392
.

256.

Pirola
 
CJ
,
Garaycoechea
M
,
Flichman
D
, et al.  
Splice variant rs72613567 prevents worst histologic outcomes in patients with nonalcoholic fatty liver disease
.
J Lipid Res.
2019
;
60
(
1
):
176
185
.

257.

Abul-Husn
 
NS
,
Cheng
X
,
Li
AH
, et al.  
A protein-truncating HSD17B13 variant and protection from chronic liver disease
.
N Engl J Med.
2018
;
378
(
12
):
1096
1106
.

258.

Ma
 
Y
,
Belyaeva
OV
,
Brown
PM
, et al. ;
for the Nonalcoholic Steatohepatitis Clinical Research Network
.
17-Beta hydroxysteroid dehydrogenase 13 is a hepatic retinol dehydrogenase associated with histological features of nonalcoholic fatty liver disease
.
Hepatology.
2019
;
69
(
4
):
1504
1519
.

259.

Yang
 
J
,
Trépo
E
,
Nahon
P
, et al.  
A HSD17B13 variant protects from hepatocellular carcinoma development in alcoholic liver disease
.
Hepatology.
2019
;70(1):231–240.

260.

McPherson
 
S
,
Hardy
T
,
Henderson
E
,
Burt
AD
,
Day
CP
,
Anstee
QM
.
Evidence of NAFLD progression from steatosis to fibrosing-steatohepatitis using paired biopsies: implications for prognosis and clinical management
.
J Hepatol.
2015
;
62
(
5
):
1148
1155
.

261.

Angulo
 
P
,
Kleiner
DE
,
Dam-Larsen
S
, et al.  
Liver fibrosis, but no other histologic features, is associated with long-term outcomes of patients with nonalcoholic fatty liver disease
.
Gastroenterology.
2015
;
149
(
2
):
389
97.e10
.

262.

Bhala
 
N
,
Angulo
P
,
van der Poorten
D
, et al.  
The natural history of nonalcoholic fatty liver disease with advanced fibrosis or cirrhosis: an international collaborative study
.
Hepatology.
2011
;
54
(
4
):
1208
1216
.

263.

Mcpherson
 
S
,
Pais
R
,
Valenti
L
, et al.  
Further delineation of fibrosis progression in NAFLD: evidence from a large cohort of patients with sequential biopsies
.
J Hepatol.
2017
;
66
(
1
):
S593
.

264.

Singh
 
S
,
Allen
AM
,
Wang
Z
,
Prokop
LJ
,
Murad
MH
,
Loomba
R
.
Fibrosis progression in nonalcoholic fatty liver vs nonalcoholic steatohepatitis: a systematic review and meta-analysis of paired-biopsy studies
.
Clin Gastroenterol Hepatol.
2015
;
13
(
4
):
643
54.e1
.

265.

Ekstedt
 
M
,
Hagström
H
,
Nasr
P
, et al.  
Fibrosis stage is the strongest predictor for disease-specific mortality in NAFLD after up to 33 years of follow-up
.
Hepatology.
2015
;
61
(
5
):
1547
1554
.

266.

Hagström
 
H
,
Nasr
P
,
Ekstedt
M
, et al.  
Fibrosis stage but not NASH predicts mortality and time to development of severe liver disease in biopsy-proven NAFLD
.
J Hepatol.
2017
;
67
(
6
):
1265
1273
.

267.

Adams
 
LA
,
Anstee
QM
,
Tilg
H
,
Targher
G
.
Non-alcoholic fatty liver disease and its relationship with cardiovascular disease and other extrahepatic diseases
.
Gut.
2017
;
66
(
6
):
1138
1153
.

268.

Faasse
 
S
,
Braun
H
,
Vos
M
.
The role of NAFLD in cardiometabolic disease: an update
.
F1000Res.
2018
;
7
:
170
.

269.

Pais
 
R
,
Giral
P
,
Khan
JF
, et al.  
Fatty liver is an independent predictor of early carotid atherosclerosis
.
J Hepatol.
2016;65(1):95–102
.

270.

Lonardo
 
A
,
Ballestri
S
,
Guaraldi
G
, et al.  
Fatty liver is associated with an increased risk of diabetes and cardiovascular disease - evidence from three different disease models: NAFLD, HCV and HIV
.
World J Gastroenterol.
2016
;
22
(
44
):
9674
9693
.

271.

Sunbul
 
M
,
Kivrak
T
,
Durmus
E
, et al.  
Nonalcoholic steatohepatitis score is an independent predictor of right ventricular dysfunction in patients with nonalcoholic fatty liver disease
.
Cardiovasc Ther.
2015
;
33
(
5
):
294
299
.

272.

Targher
 
G
,
Byrne
CD
,
Lonardo
A
,
Zoppini
G
,
Barbui
C
.
Non-alcoholic fatty liver disease and risk of incident cardiovascular disease: a meta-analysis
.
J Hepatol.
2016
;
65
(
3
):
589
600
.

273.

Mahfood Haddad
 
T
,
Hamdeh
S
,
Kanmanthareddy
A
,
Alla
VM
.
Nonalcoholic fatty liver disease and the risk of clinical cardiovascular events: a systematic review and meta-analysis
.
Diabetes Metab Syndr.
2017
;
11
(Suppl 1)
:
S209
S216
.

274.

Wild
 
SH
,
Walker
JJ
,
Morling
JR
, et al. ;
Scottish Diabetes Research Network Epidemiology Group
.
Cardiovascular disease, cancer, and mortality among people with type 2 diabetes and alcoholic or nonalcoholic fatty liver disease hospital admission
.
Diabetes Care.
2018
;
41
(
2
):
341
347
.

275.

Targher
 
G
,
Marchesini
G
,
Byrne
CD
.
Risk of type 2 diabetes in patients with non-alcoholic fatty liver disease: causal association or epiphenomenon?
Diabetes Metab.
2016
;
42
(
3
):
142
156
.

276.

Ekstedt
 
M
,
Franzén
LE
,
Mathiesen
UL
, et al.  
Long-term follow-up of patients with NAFLD and elevated liver enzymes
.
Hepatology.
2006
;
44
(
4
):
865
873
.

277.

Mantovani
 
A
,
Zaza
G
,
Byrne
CD
, et al.  
Nonalcoholic fatty liver disease increases risk of incident chronic kidney disease: a systematic review and meta-analysis
.
Metabolism.
2018;79:64–76
.

278.

Musso
 
G
,
Gambino
R
,
Tabibian
JH
, et al.  
Association of non-alcoholic fatty liver disease with chronic kidney disease: a systematic review and meta-analysis
.
PloS Med.
2014
;
11
(
7
):
e1001680
.

279.

Targher
 
G
,
Chonchol
MB
,
Byrne
CD
.
CKD and nonalcoholic fatty liver disease
.
Am J Kidney Dis.
2014
;
64
(
4
):
638
652
.

280.

Sanna
 
C
,
Rosso
C
,
Marietti
M
,
Bugianesi
E
.
Non-alcoholic fatty liver disease and extra-hepatic cancers
.
Int J Mol Sci.
2016;17(5):E717
.

281.

Wong
 
MCS
,
Ching
JYL
,
Chan
VCW
, et al.  
Screening strategies for colorectal cancer among patients with nonalcoholic fatty liver disease and family history
.
Int J Cancer.
2016;138(3):576–83
.

282.

Kim
 
GA
,
Lee
HC
,
Choe
J
, et al.  
Association between non-alcoholic fatty liver disease and cancer incidence rate
.
J Hepatol.
2018;68(1):140–146
.

283.

Brunt
 
EM
,
Kleiner
DE
,
Wilson
LA
,
Belt
P
,
Neuschwander-Tetri
BA
,
NASH Clinical Research Network
.
The NAS and the histopathologic diagnosis in NAFLD: distinict clinicopathologic meanings
.
Hepatology.
2011;53(3):810–820
.

284.

Kleiner
 
DE
,
Brunt
EM
,
Van Natta
M
, et al. ;
Nonalcoholic Steatohepatitis Clinical Research Network
.
Design and validation of a histological scoring system for nonalcoholic fatty liver disease
.
Hepatology.
2005
;
41
(
6
):
1313
1321
.

285.

Chalasani
 
N
,
Younossi
Z
,
Lavine
JE
, et al.  
The diagnosis and management of nonalcoholic fatty liver disease: practice guidance from the American Association for the Study of Liver Diseases
.
Hepatology.
2018;67(1):328–357
.

286.

Stål
 
P
.
Liver fibrosis in non-alcoholic fatty liver disease-diagnostic challenge with prognostic significance
.
World J Gastroenterol.
2015;21(39):11077–11087
. doi:.

287.

Younossi
 
ZM
,
Stepanova
M
,
Rafiq
N
, et al.  
Pathologic criteria for nonalcoholic steatohepatitis: interprotocol agreement and ability to predict liver-related mortality
.
Hepatology.
2011;54(1):369–370
. doi:.

288.

Dulai
 
PS
,
Singh
S
,
Patel
J
, et al.  
Increased risk of mortality by fibrosis stage in nonalcoholic fatty liver disease: systematic review and meta-analysis
.
Hepatology.
2017
;
65
(
5
):
1557
1565
.

289.

Vilar-Gomez
 
E
,
Calzadilla-Bertot
L
,
Wai-Sun Wong
V
, et al.  
Fibrosis severity as a determinant of cause-specific mortality in patients with advanced nonalcoholic fatty liver disease: a multi-national cohort study
.
Gastroenterology.
2018
;
155
(
2
):
443
457.e17
.

290.

Reig
 
M
,
Gambato
M
,
Man
NK
, et al.  
Should patients with NAFLD/NASH be surveyed for HCC?
 
Transplantation.
 
2019
;
103
(
1
):
39
44
.

291.

Lee
 
TY
,
Wu
JC
,
Yu
SH
,
Lin
JT
,
Wu
MS
,
Wu
CY
.
The occurrence of hepatocellular carcinoma in different risk stratifications of clinically noncirrhotic nonalcoholic fatty liver disease
.
Int J Cancer.
2017;141(7):1307–1314
. doi:.

292.

Desai
 
A
,
Sandhu
S
,
Lai
J-P
,
Sandhu
DS
.
Hepatocellular carcinoma in non-cirrhotic liver: a comprehensive review
.
World J Hepatol.
2019;11(1):1–18
. doi:.

293.

Schütte
 
K
,
Schulz
C
,
Poranzke
J
, et al.  
Characterization and prognosis of patients with hepatocellular carcinoma (HCC) in the non-cirrhotic liver
.
BMC Gastroenterol.
2014;14:117
. doi:.

294.

Burra
 
P
,
Berenguer
M
,
Pomfret
E
.
The ILTS consensus conference on NAFLD/NASH and liver transplantation: setting the stage
.
Transplantation.
2019
;
103
(
1
):
19
21
.

295.

Singal
 
AK
,
Guturu
P
,
Hmoud
B
,
Kuo
YF
,
Salameh
H
,
Wiesner
RH
.
Evolving frequency and outcomes of liver transplantation based on etiology of liver disease
.
Transplantation.
2013
;
95
(
5
):
755
760
.

296.

Zarrinpar
 
A
,
Busuttil
RW
.
Liver transplantation: past, present and future
.
Nat Rev Gastroenterol Hepatol.
2013
;
10
(
7
):
434
440
.

297.

Cholankeril
 
G
,
Wong
RJ
,
Hu
M
, et al.  
Liver transplantation for nonalcoholic steatohepatitis in the US: temporal trends and outcomes
.
Dig Dis Sci.
2017
;
62
(
10
):
2915
2922
.

298.

Mercado-Irizarry
 
A
,
Torres
EA
.
Cryptogenic cirrhosis: current knowledge and future directions
.
Clin Liver Dis (Hoboken).
2016
;
7
(
4
):
69
72
.

299.

Shingina
 
A
,
DeWitt
PE
,
Dodge
JL
, et al.  
Future trends in demand for liver transplant: birth cohort effects among patients with NASH and HCC
.
Transplantation.
2019
;
103
(
1
):
140
148
.

300.

World Health Organization
.
WHO European Childhood Obesity Surveillance Initiative: overweight and obesity among 6–9-year-old children. Report of the third round of data collection 2012–2013
. Copenhagen, Denmark, World Health Organization. http://www.euro.who.int/__data/assets/pdf_file/0010/378865/COSI-3.pdf?ua=1. Accessed
September 16, 2019
.

301.

Newsome
 
PN
,
Allison
ME
,
Andrews
PA
, et al. ;
British Transplant Society
.
Guidelines for liver transplantation for patients with non-alcoholic steatohepatitis
.
Gut.
2012
;
61
(
4
):
484
500
.

302.

Tsochatzis
 
E
,
Coilly
A
,
Nadalin
S
, et al.  
International liver transplantation consensus statement on end-stage liver disease due to nonalcoholic steatohepatitis and liver transplantation
.
Transplantation.
2019
;
103
(
1
):
45
56
.

303.

Haldar
 
D
,
Kern
B
,
Hodson
J
, et al.  
Outcomes of liver transplantation for non-alcoholic steatohepatitis: a European liver transplant registry study
.
J Hepatol.
2019;71(2):313–322
. doi:.

304.

Charlton
 
MR
,
Burns
JM
,
Pedersen
RA
,
Watt
KD
,
Heimbach
JK
,
Dierkhising
RA
.
Frequency and outcomes of liver transplantation for nonalcoholic steatohepatitis in the United States
.
Gastroenterology.
2011
;
141
(
4
):
1249
1253
.

305.

Wang
 
X
,
Li
J
,
Riaz
DR
,
Shi
G
,
Liu
C
,
Dai
Y
.
Outcomes of liver transplantation for nonalcoholic steatohepatitis: a systematic review and meta-analysis
.
Clin Gastroenterol Hepatol.
2014
;
12
(
3
):
394
402
.e1.

306.

Dick
 
AA
,
Spitzer
AL
,
Seifert
CF
, et al.  
Liver transplantation at the extremes of the body mass index
.
Liver Transpl.
2009
;
15
(
8
):
968
977
.

307.

VanWagner
 
LB
,
Montag
S
,
Zhao
L
, et al.  
Cardiovascular disease outcomes related to early stage renal impairment after liver transplantation
.
Transplantation.
2018
;
102
(
7
):
1096
1107
.

308.

Germani
 
G
,
Laryea
M
,
Rubbia-Brandt
L
, et al.  
Management of recurrent and de novo NAFLD/NASH after liver transplantation
.
Transplantation.
2019
;
103
(
1
):
57
67
.

309.

Dowman
 
JK
,
Tomlinson
JW
,
Newsome
PN
.
Systematic review: the diagnosis and staging of non-alcoholic fatty liver disease and non-alcoholic steatohepatitis
.
Aliment Pharmacol Ther.
2011
;
33
(
5
):
525
540
.

310.

Portillo-Sanchez
 
P
,
Bril
F
,
Maximos
M
, et al.  
High prevalence of nonalcoholic fatty liver disease in patients with type 2 diabetes mellitus and normal plasma aminotransferase levels
.
J Clin Endocrinol Metab.
2015
;
100
(
6
):
2231
2238
.

311.

Hernaez
 
R
,
Lazo
M
,
Bonekamp
S
, et al.  
Diagnostic accuracy and reliability of ultrasonography for the detection of fatty liver: a meta-analysis
.
Hepatology.
2011;54(3):1082–1090
. doi:.

312.

Fishbein
 
M
,
Castro
F
,
Cheruku
S
, et al.  
Hepatic MRI for fat quantitation: its relationship to fat morphology, diagnosis, and ultrasound
.
J Clin Gastroenterol.
2005;39(7):619–625
. doi:.

313.

Bedogni
 
G
,
Bellentani
S
,
Miglioli
L
, et al.  
The fatty liver index: a simple and accurate predictor of hepatic steatosis in the general population
.
BMC Gastroenterol.
2006
;
6
:
33
.

314.

Kotronen
 
A
,
Peltonen
M
,
Hakkarainen
A
, et al.  
Prediction of non-alcoholic fatty liver disease and liver fat using metabolic and genetic factors
.
Gastroenterology.
2009
;
137
(
3
):
865
872
.

315.

Fedchuk
 
L
,
Nascimbeni
F
,
Pais
R
,
Charlotte
F
,
Housset
C
,
Ratziu
V
;
LIDO Study Group
.
Performance and limitations of steatosis biomarkers in patients with nonalcoholic fatty liver disease
.
Aliment Pharmacol Ther.
2014
;
40
(
10
):
1209
1222
.

316.

Bedossa
 
P
;
FLIP Pathology Consortium
.
Utility and appropriateness of the fatty liver inhibition of progression (FLIP) algorithm and steatosis, activity, and fibrosis (SAF) score in the evaluation of biopsies of nonalcoholic fatty liver disease
.
Hepatology.
2014
;
60
(
2
):
565
575
.

317.

Sanyal
 
AJ
,
Brunt
EM
,
Kleiner
DE
, et al.  
Endpoints and clinical trial design for nonalcoholic steatohepatitis
.
Hepatology.
2011
;
54
(
1
):
344
353
.

318.

Cheung
 
A
,
Neuschwander‐Tetri
BA
,
Kleiner
DE
, et al.  
Defining improvement in nonalcoholic steatohepatitis for treatment trial endpoints: recommendations from the liver forum
.
Hepatology.
2019;70(5):1841-1855
. doi:.

319.

Brunt
 
EM
,
Janney
CG
,
Di Bisceglie
AM
,
Neuschwander-Tetri
BA
,
Bacon
BR
.
Nonalcoholic steatohepatitis: a proposal for grading and staging the histological lesions
.
Am J Gastroenterol.
1999
;
94
(
9
):
2467
2474
.

320.

Buzzetti
 
E
,
Hall
A
,
Ekstedt
M
, et al.  
Collagen proportionate area is an independent predictor of long-term outcome in patients with non-alcoholic fatty liver disease
.
Aliment Pharmacol Ther.
2019;49(9):1214–1222
. doi:.

321.

Verma
 
S
,
Jensen
D
,
Hart
J
,
Mohanty
SR
.
Predictive value of ALT levels for non-alcoholic steatohepatitis (NASH) and advanced fibrosis in non-alcoholic fatty liver disease (NAFLD)
.
Liver Int.
2013
;
33
(
9
):
1398
1405
.

322.

Seeff
 
LB
,
Everson
GT
,
Morgan
TR
, et al. ;
HALT–C Trial Group
.
Complication rate of percutaneous liver biopsies among persons with advanced chronic liver disease in the HALT-C trial
.
Clin Gastroenterol Hepatol.
2010
;
8
(
10
):
877
883
.

323.

Rockey
 
DC
,
Caldwell
SH
,
Goodman
ZD
,
Nelson
RC
,
Smith
AD
.
AASLD position paper: liver biopsy
.
Hepatology.
2008;49(3):1017–1044
. doi:.

324.

Castera
 
L
,
Friedrich-Rust
M
,
Loomba
R
.
Noninvasive assessment of liver disease in patients with nonalcoholic fatty liver disease
.
Gastroenterology.
2019;156(5):1264–1281.e4
. doi:.

325.

Chin
 
JL
,
Pavlides
M
,
Moolla
A
,
Ryan
JD
.
Non-invasive markers of liver fibrosis: adjuncts or alternatives to liver biopsy?
Front Pharmacol.
2016
;
7
:
159
.

326.

Younossi
 
ZM
,
Loomba
R
,
Anstee
QM
, et al.  
Diagnostic modalities for nonalcoholic fatty liver disease, nonalcoholic steatohepatitis, and associated fibrosis
.
Hepatology.
2018;68(1):349–360
. doi:.

327.

Vilar-Gomez
 
E
,
Chalasani
N
.
Non-invasive assessment of non-alcoholic fatty liver disease: clinical prediction rules and blood-based biomarkers
.
J Hepatol.
2018
;
68
(
2
):
305
315
.

328.

Williams
 
ALB
,
Hoofnagle
JH
.
Ratio of serum aspartate to alanine aminotransferase in chronic hepatitis relationship to cirrhosis
.
Gastroenterology.
1988;95(3):734–739
. doi:.

329.

Vallet-Pichard
 
A
,
Mallet
V
,
Nalpas
B
, et al.  
FIB-4: an inexpensive and accurate marker of fibrosis in HCV infection. Comparison with liver biopsy and FibroTest
.
Hepatology.
2007;46(1):32–36
. doi:.

330.

Angulo
 
P
,
Hui
JM
,
Marchesini
G
, et al.  
The NAFLD fibrosis score: a noninvasive system that identifies liver fibrosis in patients with NAFLD
.
Hepatology.
2007
;
45
(
4
):
846
854
.

331.

Harrison
 
SA
,
Oliver
D
,
Arnold
HL
,
Gogia
S
,
Neuschwander-Tetri
BA
.
Development and validation of a simple NAFLD clinical scoring system for identifying patients without advanced disease
.
Gut.
2008
;
57
(
10
):
1441
1447
.

332.

McPherson
 
S
,
Stewart
SF
,
Henderson
E
,
Burt
AD
,
Day
CP
.
Simple non-invasive fibrosis scoring systems can reliably exclude advanced fibrosis in patients with non-alcoholic fatty liver disease
.
Gut.
2010
;
59
(
9
):
1265
1269
.

333.

Guha
 
IN
,
Parkes
J
,
Roderick
P
, et al.  
Noninvasive markers of fibrosis in nonalcoholic fatty liver disease: validating the European liver fibrosis panel and exploring simple markers
.
Hepatology.
2008;47(2):373–375
. doi:.

334.

Ratziu
 
V
,
Massard
J
,
Charlotte
F
, et al. ;
LIDO Study Group; CYTOL study group
.
Diagnostic value of biochemical markers (FibroTest-FibroSURE) for the prediction of liver fibrosis in patients with non-alcoholic fatty liver disease
.
BMC Gastroenterol.
2006
;
6
:
6
.

335.

Munteanu
 
M
,
Tiniakos
D
,
Anstee
Q
, et al. ;
FLIP Consortium and the FibroFrance Group
.
Diagnostic performance of FibroTest, SteatoTest and ActiTest in patients with NAFLD using the SAF score as histological reference
.
Aliment Pharmacol Ther.
2016
;
44
(
8
):
877
889
.

336.

Sandrin
 
L
,
Fourquet
B
,
Hasquenoph
JM
, et al.  
Transient elastography: a new noninvasive method for assessment of hepatic fibrosis
.
Ultrasound Med Biol.
2003;29(12):1705–1713
. doi:.

337.

Eddowes
 
PJ
,
Sasso
M
,
Allison
M
, et al.  
Accuracy of FibroScan controlled attenuation parameter and liver stiffness measurement in assessing steatosis and fibrosis in patients with nonalcoholic fatty liver disease
.
Gastroenterology.
2019
;
156
(
6
):
1717
1730
.

338.

Szczepaniak
 
LS
,
Nurenberg
P
,
Leonard
D
, et al.  
Magnetic resonance spectroscopy to measure hepatic triglyceride content: prevalence of hepatic steatosis in the general population
.
Am J Physiol Endocrinol Metab.
2005
;
288
(
2
):
E462
E468
.

339.

Gu
 
J
,
Liu
S
,
Du
S
, et al.  
Diagnostic value of MRI-PDFF for hepatic steatosis in patients with non-alcoholic fatty liver disease: a meta-analysis
.
Eur Radiol.
2019;29(7):3564–3573
. doi:.

340.

Tang
 
A
,
Desai
A
,
Hamilton
G
, et al.  
Accuracy of MR imaging–estimated proton density fat fraction for classification of dichotomized histologic steatosis grades in nonalcoholic fatty liver disease
.
Radiology.
2014;274(2):416–425
. doi:.

341.

Ajmera
 
V
,
Park
CC
,
Caussy
C
, et al.  
Magnetic resonance imaging proton density fat fraction associates with progression of fibrosis in patients with nonalcoholic fatty liver disease
.
Gastroenterology.
2018
;
155
(
2
):
307
310.e2
.

342.

Chen
 
J
,
Talwalkar
JA
,
Yin
M
,
Glaser
KJ
,
Sanderson
SO
,
Ehman
RL
.
Early detection of nonalcoholic steatohepatitis in patients with nonalcoholic fatty liver disease by using MR elastography
.
Radiology.
2011
;
259
(
3
):
749
756
.

343.

Xiao
 
G
,
Zhu
S
,
Xiao
X
,
Yan
L
,
Yang
J
,
Wu
G
.
Comparison of laboratory tests, ultrasound, or magnetic resonance elastography to detect fibrosis in patients with nonalcoholic fatty liver disease: a meta-analysis
.
Hepatology.
2017;66(5):1486–1501
. doi:.

344.

Pavlides
 
M
,
Banerjee
R
,
Tunnicliffe
EM
, et al.  
Multiparametric magnetic resonance imaging for the assessment of non-alcoholic fatty liver disease severity
.
Liver Int.
2017
;
37
(
7
):
1065
1073
.

345.

Armstrong
 
MJ
,
Houlihan
DD
,
Bentham
L
, et al.  
Presence and severity of non-alcoholic fatty liver disease in a large prospective primary care cohort
.
J Hepatol.
2012
;
56
(
1
):
234
240
.

346.

Koehler
 
EM
,
Schouten
JN
,
Hansen
BE
, et al.  
Prevalence and risk factors of non-alcoholic fatty liver disease in the elderly: results from the Rotterdam study
.
J Hepatol.
2012
;
57
(
6
):
1305
1311
.

347.

Harman
 
DJ
,
Ryder
SD
,
James
MW
, et al.  
Obesity and type 2 diabetes are important risk factors underlying previously undiagnosed cirrhosis in general practice: a cross-sectional study using transient elastography
.
Aliment Pharmacol Ther.
2018
;
47
(
4
):
504
515
.

348.

Roulot
 
D
,
Roudot-Thoraval
F
,
NKontchou
G
, et al.  
Concomitant screening for liver fibrosis and steatosis in French type 2 diabetic patients using Fibroscan
.
Liver Int.
2017
;
37
(
12
):
1897
1906
.

349.

Morling
 
JR
,
Fallowfield
JA
,
Guha
IN
, et al. ;
Edinburgh Type 2 Diabetes Study investigators
.
Using non-invasive biomarkers to identify hepatic fibrosis in people with type 2 diabetes mellitus: the Edinburgh type 2 diabetes study
.
J Hepatol.
2014
;
60
(
2
):
384
391
.

350.

Newsome
 
PN
,
Cramb
R
,
Davison
SM
, et al.  
Guidelines on the management of abnormal liver blood tests
.
Gut.
2018
;
67
(
1
):
6
19
.

351.

Petta
 
S
,
Wong
VW
,
Cammà
C
, et al.  
Serial combination of non-invasive tools improves the diagnostic accuracy of severe liver fibrosis in patients with NAFLD
.
Aliment Pharmacol Ther.
2017
;
46
(
6
):
617
627
.

352.

Srivastava
 
A
,
Gailer
R
,
Tanwar
S
, et al.  
Prospective evaluation of a primary care referral pathway for patients with non-alcoholic fatty liver disease
.
J Hepatol.
2019
;
71
(
2
):
371
378
.

353.

Hippisley-Cox
 
J
,
Coupland
C
,
Brindle
P
.
Development and validation of QRISK3 risk prediction algorithms to estimate future risk of cardiovascular disease: prospective cohort study
.
Bmj.
2017
;
357
:
j2099
.

354.

D’Agostino
 
RB
,
Vasan
RS
,
Pencina
MJ
, et al.  
General cardiovascular risk profile for use in primary care: the Framingham heart study
.
Circulation.
2008;117(6):743–753
. doi:.

355.

Abeles
 
RD
,
Mullish
BH
,
Forlano
R
, et al.  
Derivation and validation of a cardiovascular risk score for prediction of major acute cardiovascular events in non-alcoholic fatty liver disease; the importance of an elevated mean platelet volume
.
Aliment Pharmacol Ther.
2019;49(8):1077–1085
. doi:.

356.

Cobbold
 
JFL
,
Taylor-Robinson
SD
.
Transient elastography in acute hepatitis: all that’s stiff is not fibrosis
.
Hepatology.
2008;47(2):370–372
. doi:.

357.

Tanajewski
 
L
,
Harris
R
,
Harman
DJ
, et al.  
Economic evaluation of a community-based diagnostic pathway to stratify adults for non-alcoholic fatty liver disease: a Markov model informed by a feasibility study
.
BMJ Open.
2017;7:e015659
. doi:.

358.

Marjot
 
T
,
Sbardella
E
,
Moolla
A
, et al.  
Prevalence and severity of non-alcoholic fatty liver disease are underestimated in clinical practice: impact of a dedicated screening approach at a large university teaching hospital
.
Diabet Med.
2018
;
35
(
1
):
89
98
.

359.

Armstrong
 
MJ
,
Hazlehurst
JM
,
Parker
R
, et al.  
Severe asymptomatic non-alcoholic fatty liver disease in routine diabetes care; a multi-disciplinary team approach to diagnosis and management
.
Qjm.
2014
;
107
(
1
):
33
41
.

360.

Moolla
 
A
,
Motohashi
K
,
Marjot
T
, et al.  
A multidisciplinary approach to the management of NAFLD is associated with improvement in markers of liver and cardio-metabolic health
.
Frontline Gastroenterol.
2019;10:337–346
. doi:.

361.

Cobbold
 
JFL
,
Raveendran
S
,
Peake
CM
,
Anstee
QM
,
Yee
MS
,
Thursz
MR
.
Piloting a multidisciplinary clinic for the management of non-alcoholic fatty liver disease: initial 5-year experience
.
Frontline Gastroenterol.
2013
;
4
(
4
):
263
269
.

362.

Townsend
 
SA
,
Newsome
PN
.
The role of a dedicated non-alcoholic fatty liver disease clinic in 2016
.
Dig Dis.
2017;35(4):371–376
. doi:.

363.

Danford
 
CJ
,
Lai
M
.
NAFLD: a multisystem disease that requires a multidisciplinary approach
.
Frontline Gastroenterol.
2019;10:328–329
. doi:.

364.

Goyale
 
A
,
Guerrero
M
,
Prat
LI
,
Roccarina
D
,
Nair
D
,
Tsochatzis
E
.
THU-309-Cardiovascular risk assessment and management improves outcomes in a specialist multidisciplinary non-alcoholic fatty liver disease clinic
.
J Hepatol.
2019;70
(1 Suppl)
;
e296
. doi:.

365.

Thivel
 
D
,
Tremblay
A
,
Genin
PM
,
Panahi
S
,
Rivière
D
,
Duclos
M
.
Physical activity, inactivity, and sedentary behaviors: definitions and implications in occupational health
.
Front Public Health.
2018
;
6
:
288
.

366.

Ryu
 
S
,
Chang
Y
,
Jung
HS
, et al.  
Relationship of sitting time and physical activity with non-alcoholic fatty liver disease
.
J Hepatol.
2015
;
63
(
5
):
1229
1237
.

367.

Pälve
 
KS
,
Pahkala
K
,
Suomela
E
, et al.  
Cardiorespiratory fitness and risk of fatty liver: the young finns study
.
Med Sci Sports Exerc.
2017
;
49
(
9
):
1834
1841
.

368.

Johnson
 
NA
,
Sachinwalla
T
,
Walton
DW
, et al.  
Aerobic exercise training reduces hepatic and visceral lipids in obese individuals without weight loss
.
Hepatology.
2009
;
50
(
4
):
1105
1112
.

369.

Sullivan
 
S
,
Kirk
EP
,
Mittendorfer
B
,
Patterson
BW
,
Klein
S
.
Randomized trial of exercise effect on intrahepatic triglyceride content and lipid kinetics in nonalcoholic fatty liver disease
.
Hepatology.
2012
;
55
(
6
):
1738
1745
.

370.

Keating
 
SE
,
Hackett
DA
,
George
J
,
Johnson
NA
.
Exercise and non-alcoholic fatty liver disease: a systematic review and meta-analysis
.
J Hepatol.
2012
;
57
(
1
):
157
166
.

371.

Keating
 
SE
,
Hackett
DA
,
Parker
HM
, et al.  
Effect of aerobic exercise training dose on liver fat and visceral adiposity
.
J Hepatol.
2015
;
63
(
1
):
174
182
.

372.

Smart
 
NA
,
King
N
,
McFarlane
JR
,
Graham
PL
,
Dieberg
G
.
Effect of exercise training on liver function in adults who are overweight or exhibit fatty liver disease: a systematic review and meta-analysis
.
Br J Sports Med.
2018
;
52
(
13
):
834
843
.

373.

Hallsworth
 
K
,
Fattakhova
G
,
Hollingsworth
KG
, et al.  
Resistance exercise reduces liver fat and its mediators in non-alcoholic fatty liver disease independent of weight loss
.
Gut.
2011
;
60
(
9
):
1278
1283
.

374.

Zelber-Sagi
 
S
,
Buch
A
,
Yeshua
H
, et al.  
Effect of resistance training on non-alcoholic fatty-liver disease a randomized-clinical trial
.
World J Gastroenterol.
2014
;
20
(
15
):
4382
4392
.

375.

Hashida
 
R
,
Kawaguchi
T
,
Bekki
M
, et al.  
Aerobic vs. resistance exercise in non-alcoholic fatty liver disease: a systematic review
.
J Hepatol.
2017
;
66
(
1
):
142
152
.

376.

Gibala
 
MJ
,
Little
JP
,
Macdonald
MJ
,
Hawley
JA
.
Physiological adaptations to low-volume, high-intensity interval training in health and disease
.
J Physiol.
2012
;
590
(
5
):
1077
1084
.

377.

Hallsworth
 
K
,
Thoma
C
,
Hollingsworth
KG
, et al.  
Modified high-intensity interval training reduces liver fat and improves cardiac function in non-alcoholic fatty liver disease: a randomized controlled trial
.
Clin Sci (Lond).
2015
;
129
(
12
):
1097
1105
.

378.

Cassidy
 
S
,
Thoma
C
,
Hallsworth
K
, et al.  
High intensity intermittent exercise improves cardiac structure and function and reduces liver fat in patients with type 2 diabetes: a randomised controlled trial
.
Diabetologia.
2016
;
59
(
1
):
56
66
.

379.

Thompson
 
D
,
Karpe
F
,
Lafontan
M
,
Frayn
K
.
Physical activity and exercise in the regulation of human adipose tissue physiology
.
Physiol Rev.
2012
;
92
(
1
):
157
191
.

380.

Shojaee-Moradie
 
F
,
Cuthbertson
DJ
,
Barrett
M
, et al.  
Exercise training reduces liver fat and increases rates of VLDL clearance but not VLDL production in NAFLD
.
J Clin Endocrinol Metab.
2016
;
101
(
11
):
4219
4228
.

381.

Leoni
 
S
,
Tovoli
F
,
Napoli
L
,
Serio
I
,
Ferri
S
,
Bolondi
L
.
Current guidelines for the management of non-alcoholic fatty liver disease: a systematic review with comparative analysis
.
World J Gastroenterol.
2018
;
24
(
30
):
3361
3373
.

382.

Parry
 
SA
,
Hodson
L
.
Influence of dietary macronutrients on liver fat accumulation and metabolism
.
J Investig Med.
2017
;
65
(
8
):
1102
1115
.

383.

Bergholm
 
R
,
Sevastianova
K
,
Santos
A
, et al.  
CB(1) blockade-induced weight loss over 48 weeks decreases liver fat in proportion to weight loss in humans
.
Int J Obes (Lond).
2013
;
37
(
5
):
699
703
.

384.

Haufe
 
S
,
Engeli
S
,
Kast
P
, et al.  
Randomized comparison of reduced fat and reduced carbohydrate hypocaloric diets on intrahepatic fat in overweight and obese human subjects
.
Hepatology.
2011
;
53
(
5
):
1504
1514
.

385.

Tiikkainen
 
M
,
Bergholm
R
,
Vehkavaara
S
, et al.  
Effects of identical weight loss on body composition and features of insulin resistance in obese women with high and low liver fat content
.
Diabetes.
2003
;
52
(
3
):
701
707
.

386.

Lewis
 
MC
,
Phillips
ML
,
Slavotinek
JP
,
Kow
L
,
Thompson
CH
,
Toouli
J
.
Change in liver size and fat content after treatment with Optifast very low calorie diet
.
Obes Surg.
2006
;
16
(
6
):
697
701
.

387.

Steven
 
S
,
Hollingsworth
KG
,
Al-Mrabeh
A
, et al.  
Very low-calorie diet and 6 months of weight stability in type 2 diabetes: pathophysiological changes in responders and nonresponders
.
Diabetes Care.
2016
;
39
(
5
):
808
815
.

388.

Taylor
 
R
,
Al-Mrabeh
A
,
Zhyzhneuskaya
S
, et al.  
Remission of human type 2 diabetes requires decrease in liver and pancreas fat content but is dependent upon capacity for β cell recovery
.
Cell Metab.
2018
;
28
(
4
):
667
.

389.

Lim
 
EL
,
Hollingsworth
KG
,
Aribisala
BS
,
Chen
MJ
,
Mathers
JC
,
Taylor
R
.
Reversal of type 2 diabetes: normalisation of beta cell function in association with decreased pancreas and liver triacylglycerol
.
Diabetologia.
2011
;
54
(
10
):
2506
2514
.

390.

Schwenger
 
KJP
,
Fischer
SE
,
Jackson
TD
,
Okrainec
A
,
Allard
JP
.
Non-alcoholic fatty liver disease in morbidly obese individuals undergoing bariatric surgery: prevalence and effect of the pre-bariatric very low calorie diet
.
Obes Surg.
2018
;
28
(
4
):
1109
1116
.

391.

Hodson
 
L
,
Frayn
KN
.
Hepatic fatty acid partitioning
.
Curr Opin Lipidol.
2011
;
22
(
3
):
216
224
.

392.

Buchwald
 
H
.
The evolution of metabolic/bariatric surgery
.
Obes Surg.
2014
;
24
(
8
):
1126
1135
.

393.

Buchwald
 
H
,
Avidor
Y
,
Braunwald
E
, et al.  
Bariatric surgery: a systematic review and meta-analysis
.
JAMA.
2004
;
292
(
14
):
1724
1737
.

394.

Laursen
 
TL
,
Hagemann
CA
,
Wei
C
, et al.  
Bariatric surgery in patients with non-alcoholic fatty liver disease - from pathophysiology to clinical effects
.
World J Hepatol.
2019
;
11
(
2
):
138
149
.

395.

Maciejewski
 
ML
,
Arterburn
DE
,
Van Scoyoc
L
, et al.  
Bariatric surgery and long-term durability of weight loss
.
JAMA Surg.
2016
;
151
(
11
):
1046
1055
.

396.

Lee
 
Y
,
Doumouras
AG
,
Yu
J
, et al.  
Complete resolution of nonalcoholic fatty liver disease after bariatric surgery: a systematic review and meta-analysis
.
Clin Gastroenterol Hepatol.
2019
;
17
(
6
):
1040
1060.e11
.

397.

Verna
 
EC
,
Berk
PD
.
Role of fatty acids in the pathogenesis of obesity and fatty liver: impact of bariatric surgery
.
Semin Liver Dis.
2008
;
28
(
4
):
407
426
.

398.

Ledoux
 
S
,
Sami
O
,
Calabrese
D
,
Le Gall
M
,
Flamant
M
,
Coupaye
M
.
Gastric bypass specifically impairs liver parameters as compared with sleeve gastrectomy, independently of evolution of metabolic disorders
.
Surg Obes Relat Dis.
2019
;
15
(
2
):
220
226
.

399.

Guida
 
C
,
Stephen
SD
,
Watson
M
, et al.  
PYY plays a key role in the resolution of diabetes following bariatric surgery in humans
.
Ebiomedicine.
2019
;
40
:
67
76
.

400.

Li
 
Y
,
Liu
L
,
Wang
B
,
Wang
J
,
Chen
D
.
Metformin in non-alcoholic fatty liver disease: a systematic review and meta-analysis
.
Biomed Rep.
2013
;
1
(
1
):
57
64
.

401.

Musso
 
G
,
Cassader
M
,
Rosina
F
,
Gambino
R
.
Impact of current treatments on liver disease, glucose metabolism and cardiovascular risk in non-alcoholic fatty liver disease (NAFLD): a systematic review and meta-analysis of randomised trials
.
Diabetologia.
2012
;
55
(
4
):
885
904
.

402.

Aithal
 
GP
,
Thomas
JA
,
Kaye
PV
, et al.  
Randomized, placebo-controlled trial of pioglitazone in nondiabetic subjects with nonalcoholic steatohepatitis
.
Gastroenterology.
2008
;
135
(
4
):
1176
1184
.

403.

Belfort
 
R
,
Harrison
SA
,
Brown
K
, et al.  
A placebo-controlled trial of pioglitazone in subjects with nonalcoholic steatohepatitis
.
N Engl J Med.
2006
;
355
(
22
):
2297
2307
.

404.

Sanyal
 
AJ
,
Chalasani
N
,
Kowdley
KV
, et al. ;
NASH CRN
.
Pioglitazone, vitamin E, or placebo for nonalcoholic steatohepatitis
.
N Engl J Med.
2010
;
362
(
18
):
1675
1685
.

405.

Viscoli
 
CM
,
Inzucchi
SE
,
Young
LH
, et al. ;
IRIS Trial Investigators
.
Pioglitazone and risk for bone fracture: safety data from a randomized clinical trial
.
J Clin Endocrinol Metab.
2017
;
102
(
3
):
914
922
.

406.

Ito
 
D
,
Shimizu
S
,
Inoue
K
, et al. .
Comparison of ipragliflozin and pioglitazone effects on nonalcoholic fatty liver disease in patients with type 2 diabetes: a randomized, 24-week, open-label, active-controlled trial
.
Diabetes Care.
2017
;
40
(
10
):
1364
1372
.

407.

Bril
 
F
,
Biernacki
DM
,
Kalavalapalli
S
, et al.  
Role of vitamin E for nonalcoholic steatohepatitis in patients with type 2 diabetes: a randomized controlled trial
.
Diabetes Care.
2019
;
42
(
8
):
1481
1488
.

408.

Armstrong
 
MJ
,
Gaunt
P
,
Aithal
GP
, et al. ;
LEAN Trial Team
.
Liraglutide safety and efficacy in patients with non-alcoholic steatohepatitis (LEAN): a multicentre, double-blind, randomised, placebo-controlled phase 2 study
.
Lancet.
2016
;
387
(
10019
):
679
690
.

409.

Feng
 
W
,
Gao
C
,
Bi
Y
, et al.  
Randomized trial comparing the effects of gliclazide, liraglutide, and metformin on diabetes with non-alcoholic fatty liver disease
.
J Diabetes.
2017;9(8):800–809
. doi:.

410.

Bouchi
 
R
,
Nakano
Y
,
Fukuda
T
, et al.  
Reduction of visceral fat by liraglutide is associated with ameliorations of hepatic steatosis, albuminuria, and micro-inflammation in type 2 diabetic patients with insulin treatment: a randomized control trial
.
Endocr J.
2017
;
64
(
3
):
269
281
.

411.

Petit
 
JM
,
Cercueil
JP
,
Loffroy
R
, et al.  
Effect of liraglutide therapy on liver fat content in patients with inadequately controlled type 2 diabetes: the Lira-NAFLD study
.
J Clin Endocrinol Metab.
2017
;
102
(
2
):
407
415
.

412.

Smits
 
MM
,
Tonneijck
L
,
Muskiet
MH
, et al.  
Twelve week liraglutide or sitagliptin does not affect hepatic fat in type 2 diabetes: a randomised placebo-controlled trial
.
Diabetologia.
2016
;
59
(
12
):
2588
2593
.

413.

Vanderheiden
 
A
,
Harrison
LB
,
Warshauer
JT
, et al.  
Mechanisms of action of liraglutide in patients with type 2 diabetes treated with high-dose insulin
.
J Clin Endocrinol Metab.
2016
;
101
(
4
):
1798
1806
.

414.

Frøssing
 
S
,
Nylander
M
,
Chabanova
E
, et al.  
Effect of liraglutide on ectopic fat in polycystic ovary syndrome: a randomized clinical trial
.
Diabetes Obes Metab.
2018
;
20
(
1
):
215
218
.

415.

Eguchi
 
Y
,
Kitajima
Y
,
Hyogo
H
, et al. ;
Japan Study Group for NAFLD (JSG-NAFLD)
.
Pilot study of liraglutide effects in non-alcoholic steatohepatitis and non-alcoholic fatty liver disease with glucose intolerance in Japanese patients (LEAN-J)
.
Hepatol Res.
2015
;
45
(
3
):
269
278
.

416.

Gastaldelli
 
A
,
Gaggini
M
,
Daniele
G
, et al.  
Exenatide improves both hepatic and adipose tissue insulin resistance: a dynamic positron emission tomography study
.
Hepatology.
2016
;
64
(
6
):
2028
2037
.

417.

Shao
 
N
,
Kuang
HY
,
Hao
M
,
Gao
XY
,
Lin
WJ
,
Zou
W
.
Benefits of exenatide on obesity and non-alcoholic fatty liver disease with elevated liver enzymes in patients with type 2 diabetes
.
Diabetes Metab Res Rev.
2014
;
30
(
6
):
521
529
.

418.

Blaslov
 
K
,
Zibar
K
,
Bulum
T
,
Duvnjak
L
.
Effect of exenatide therapy on hepatic fat quantity and hepatic biomarkers in type 2 diabetic patients
.
Clin Res Hepatol Gastroenterol.
2014
;
38
(
3
):
e61
e63
.

419.

Gluud
 
LL
,
Knop
FK
,
Vilsbøll
T
.
Effects of lixisenatide on elevated liver transaminases: systematic review with individual patient data meta-analysis of randomised controlled trials on patients with type 2 diabetes
.
BMJ Open.
2014
;
4
(
12
):
e005325
.

420.

Investigation of Efficacy and Safety of Three Dose Levels of Subcutaneous Semaglutide Once Daily Versus Placebo in Subjects With Non-Alcoholic Steatohepatitis
. ClinicalTrials.gov identifier
NCT02970942
. https://clinicaltrials.gov/ct2/show/NCT02970942. Accessed
September 16, 2019
.

421.

Seko
 
Y
,
Sumida
Y
,
Tanaka
S
, et al.  
Effect of 12-week dulaglutide therapy in Japanese patients with biopsy-proven non-alcoholic fatty liver disease and type 2 diabetes mellitus
.
Hepatol Res.
2017
;
47
(
11
):
1206
1211
.

422.

Cusi
 
K
,
Sattar
N
,
García-Pérez
LE
, et al.  
Dulaglutide decreases plasma aminotransferases in people with Type 2 diabetes in a pattern consistent with liver fat reduction: a post hoc analysis of the AWARD programme
.
Diabet Med.
2018
;
35
(
10
):
1434
1439
.

423.

A Study of Multiple Doses of HM15211 in Obese Subjects With NAFLD and in Subjects With NASH
. ClinicalTrials.gov identifier
NCT03744182
. https://clinicaltrials.gov/ct2/show/NCT03744182. Accessed
September 16, 2019
.

424.

Cui
 
J
,
Philo
L
,
Nguyen
P
, et al.  
Sitagliptin vs. placebo for non-alcoholic fatty liver disease: a randomized controlled trial
.
J Hepatol.
2016
;
65
(
2
):
369
376
.

425.

Fukuhara
 
T
,
Hyogo
H
,
Ochi
H
, et al.  
Efficacy and safety of sitagliptin for the treatment of nonalcoholic fatty liver disease with type 2 diabetes mellitus
.
Hepatogastroenterology.
2014
;
61
(
130
):
323
328
.

426.

Macauley
 
M
,
Hollingsworth
KG
,
Smith
FE
, et al.  
Effect of vildagliptin on hepatic steatosis
.
J Clin Endocrinol Metab.
2015
;
100
(
4
):
1578
1585
.

427.

Sattar
 
N
,
Fitchett
D
,
Hantel
S
,
George
JT
,
Zinman
B
.
Empagliflozin is associated with improvements in liver enzymes potentially consistent with reductions in liver fat: results from randomised trials including the EMPA-REG OUTCOME® trial
.
Diabetologia.
2018
;
61
(
10
):
2155
2163
.

428.

Kuchay
 
MS
,
Krishan
S
,
Mishra
SK
, et al.  
Effect of empagliflozin on liver fat in patients with type 2 diabetes and nonalcoholic fatty liver disease: a randomized controlled trial (E-LIFT Trial)
.
Diabetes Care.
2018
;
41
(
8
):
1801
1808
.

429.

Lai
 
L-L
,
Vethakkan
SR
,
Nik Mustapha
NR
,
Mahadeva
S
,
Chan
W-K
.
Empagliflozin for the treatment of nonalcoholic steatohepatitis in patients with type 2 diabetes mellitus
.
Dig Dis Sci.
2019
. doi:. [Epub ahead of print]

430.

Seko
 
Y
,
Nishikawa
T
,
Umemura
A
, et al.  
Efficacy and safety of canagliflozin in type 2 diabetes mellitus patients with biopsy-proven nonalcoholic steatohepatitis classified as stage 1-3 fibrosis
.
Diabetes Metab Syndr Obes.
2018
;
11
:
835
843
.

431.

Cusi
 
K
,
Bril
F
,
Barb
D
, et al.  
Effect of canagliflozin treatment on hepatic triglyceride content and glucose metabolism in patients with type 2 diabetes
.
Diabetes Obes Metab.
2019
. [Epub ahead of print]doi:.

432.

Ohki
 
T
,
Isogawa
A
,
Toda
N
,
Tagawa
K
.
Effectiveness of Ipragliflozin, a sodium-glucose co-transporter 2 inhibitor, as a second-line treatment for non-alcoholic fatty liver disease patients with type 2 diabetes mellitus who do not respond to incretin-based therapies including glucagon-like peptide-1 analogs and dipeptidyl peptidase-4 inhibitors
.
Clin Drug Investig.
2016
;
36
(
4
):
313
319
.

433.

Takeda
 
A
,
Irahara
A
,
Nakano
A
, et al.  
The improvement of the hepatic histological findings in a patient with non-alcoholic steatohepatitis with type 2 diabetes after the administration of the sodium-glucose cotransporter 2 inhibitor ipragliflozin
.
Intern Med.
2017
;
56
(
20
):
2739
2744
.

434.

Latva-Rasku
 
A
,
Honka
MJ
,
Kullberg
J
, et al.  
The SGLT2 inhibitor dapagliflozin reduces liver fat but does not affect tissue insulin sensitivity: a randomized, double-blind, placebo-controlled study with 8-week treatment in type 2 diabetes patients
.
Diabetes Care.
2019
;
42
(
5
):
931
937
.

435.

Eriksson
 
JW
,
Lundkvist
P
,
Jansson
PA
, et al.  
Effects of dapagliflozin and n-3 carboxylic acids on non-alcoholic fatty liver disease in people with type 2 diabetes: a double-blind randomised placebo-controlled study
.
Diabetologia.
2018
;
61
(
9
):
1923
1934
.

436.

Effect of LIK066 on Reduction of Fatty Content in Livers of Obese Patients
. ClinicalTrials.gov identifier
NCT03205150
. https://clinicaltrials.gov/ct2/show/NCT03205150. Accessed
September 16, 2019
.

437.

Neuschwander-Tetri
 
BA
,
Loomba
R
,
Sanyal
AJ
, et al. ;
NASH Clinical Research Network
.
Farnesoid X nuclear receptor ligand obeticholic acid for non-cirrhotic, non-alcoholic steatohepatitis (FLINT): a multicentre, randomised, placebo-controlled trial
.
Lancet.
2015
;
385
(
9972
):
956
965
.

438.

Trauner
 
M
,
Gulamhusein
A
,
Hameed
B
, et al.  
The nonsteroidal farnesoid X receptor agonist cilofexor (GS-9674) improves markers of cholestasis and liver injury in patients with primary sclerosing cholangitis
.
Hepatology.
2019;70(3):788–801
. doi:.

439.

Study of Safety and Efficacy of Tropifexor (LJN452) in Patients With Non-alcoholic Steatohepatitis (NASH) (FLIGHT-FXR)
. ClinicalTrials.gov identifier
NCT02855164
. https://clinicaltrials.gov/ct2/show/NCT02855164. Accessed
September 16, 2019
.

440.

Ratziu
 
V
,
Harrison
SA
,
Francque
S
, et al. ;
GOLDEN-505 Investigator Study Group
.
Elafibranor, an agonist of the peroxisome proliferator-activated receptor-α and -δ, induces resolution of nonalcoholic steatohepatitis without fibrosis worsening
.
Gastroenterology.
2016
;
150
(
5
):
1147
1159.e5
.

441.

Phase 2b Study in NASH to Assess IVA337 (NATIVE)
. ClinicalTrials.gov identifier
NCT03008070
. https://clinicaltrials.gov/ct2/show/NCT03008070. Accessed
September 16, 2019
.

442.

Jani
 
RH
,
Pai
V
,
Jha
P
, et al.  
A multicenter, prospective, randomized, double-blind study to evaluate the safety and efficacy of Saroglitazar 2 and 4 mg compared with placebo in type 2 diabetes mellitus patients having hypertriglyceridemia not controlled with atorvastatin therapy (PRESS VI)
.
Diabetes Technol Ther.
2014
;
16
(
2
):
63
71
.

443.

Friedman
 
SL
,
Ratziu
V
,
Harrison
SA
, et al.  
A randomized, placebo-controlled trial of cenicriviroc for treatment of nonalcoholic steatohepatitis with fibrosis
.
Hepatology.
2018
;
67
(
5
):
1754
1767
.

444.

Lavine
 
JE
,
Schwimmer
JB
,
Van Natta
ML
, et al. ;
Nonalcoholic Steatohepatitis Clinical Research Network
.
Effect of vitamin E or metformin for treatment of nonalcoholic fatty liver disease in children and adolescents: the TONIC randomized controlled trial
.
Jama.
2011
;
305
(
16
):
1659
1668
.

445.

Harrison
 
SA
,
Marri
SR
,
Chalasani
N
, et al.  
Randomised clinical study: GR-MD-02, a galectin-3 inhibitor, vs. placebo in patients having non-alcoholic steatohepatitis with advanced fibrosis
.
Aliment Pharmacol Ther.
2016
;
44
(
11-12
):
1183
1198
.

446.

Chalasani
 
N
,
Garcia-Tsao
G
,
Goodman
Z
, et al.  
A multicenter, randomized, double-blind, PLB-controlled trial of Galectin-3 inhibitor (GR-MD-02) in patients with NASH cirrhosis and portal hypertension (LBO-001)
.
J Hepatol.
2018
;
68
:
S100
S101
.

447.

Harrison
 
SA
,
Rinella
ME
,
Abdelmalek
MF
, et al.  
NGM282 for treatment of non-alcoholic steatohepatitis: a multicentre, randomised, double-blind, placebo-controlled, phase 2 trial
.
Lancet.
2018
;
391
(
10126
):
1174
1185
.

448.

Harrison
 
SA
,
Rossi
SJ
,
Paredes
AH
, et al.  
NGM282 improves liver fibrosis and histology in 12 weeks in patients with nonalcoholic steatohepatitis
.
Hepatology.
2019
. doi:. [Epub ahead of print]

449.

Gaich
 
G
,
Chien
JY
,
Fu
H
, et al.  
The effects of LY2405319, an FGF21 analog, in obese human subjects with type 2 diabetes
.
Cell Metab.
2013
;
18
(
3
):
333
340
.

450.

Sanyal
 
A
,
Charles
ED
,
Neuschwander-Tetri
BA
, et al.  
Pegbelfermin (BMS-986036), a PEGylated fibroblast growth factor 21 analogue, in patients with non-alcoholic steatohepatitis: a randomised, double-blind, placebo-controlled, phase 2a trial
.
Lancet.
2019
;
392
(
10165
):
2705
2717
.

451.

Harrison
 
S
,
Moussa
S
,
Bashir
M
, et al.  
MGL-3196, a selective thyroid hormone receptor-beta agonist significantly decreases hepatic fat in NASH patients at 12 weeks, the primary endpoint in a 36 week serial liver biopsy study (GS-009)
.
J Hepatol.
2018
;
68
:
S38
.

452.

Loomba
 
R
,
Lawitz
E
,
Mantry
PS
, et al. ,
GS-US-384-1497 Investigators
.
The ASK1 inhibitor selonsertib in patients with nonalcoholic steatohepatitis: a randomized, phase 2 trial
.
Hepatology.
2017;67(2):549–559
. doi:.

453.

Harrison
 
SA
,
Abdelmalek
MF
,
Caldwell
S
, et al. ;
GS-US-321-0105 and GS-US-321-0106 Investigators
.
Simtuzumab is ineffective for patients with bridging fibrosis or compensated cirrhosis caused by nonalcoholic steatohepatitis
.
Gastroenterology.
2018
;
155
(
4
):
1140
1153
.

454.

Zein
 
CO
,
Yerian
LM
,
Gogate
P
, et al.  
Pentoxifylline improves nonalcoholic steatohepatitis: a randomized placebo-controlled trial
.
Hepatology.
2011
;
54
(
5
):
1610
1619
.

455.

Van Wagner
 
LB
,
Koppe
SW
,
Brunt
EM
, et al.  
Pentoxifylline for the treatment of non-alcoholic steatohepatitis: a randomized controlled trial
.
Ann Hepatol.
2011
;
10
(
3
):
277
286
.

456.

Frenette
 
CT
,
Morelli
G
,
Shiffman
ML
, et al.  
Emricasan improves liver function in patients with cirrhosis and high model for end-stage liver disease scores compared with placebo
.
Clin Gastroenterol Hepatol.
2019
;
17
(
4
):
774
783.e4
.

457.

Shiffman
 
M
,
Freilich
B
,
Vuppalanchi
R
, et al.  
Randomised clinical trial: emricasan versus placebo significantly decreases ALT and caspase 3/7 activation in subjects with non-alcoholic fatty liver disease
.
Aliment Pharmacol Ther.
2019
;
49
(
1
):
64
73
.

458.

Study of SGM-1019 in Patients With Nonalcoholic Steatohepatitis (NASH)
. ClinicalTrials.gov identifier
NCT03676231
. https://clinicaltrials.gov/ct2/show/NCT03676231. Accessed
September 16, 2019
.

459.

Dabbagh
 
K
,
Steven Dodson
G
,
Yamamoto
L
,
Baeza-Raja
B
,
Goodyear
AW
.
Preclinical and first-in human development of SGM-1019, a first-in-class novel small molecule modulator of inflammasome activity for the treatment of nonalcoholic steatohepatitis (NASH) (PS-111)
.
J Hepatol.
2018
;
68
:
S60
.

460.

Diehl
 
AM
,
Harrison
S
,
Caldwell
S
, et al.  
JKB-121 in patients with nonalcoholic steatohepatitis: a phase 2 double blind randomized placebo control study (LBO-006)
.
J Hepatol.
2018
;
68
:
S103
.

461.

A Phase ll study of IMM-124E for Patients With Non-alcoholic Steatohepatitis
. ClinicalTrials.gov identifier
NCT02316717
. https://clinicaltrials.gov/ct2/show/NCT02316717. Accessed September 16, 2019.

462.

Oral
 
EA
,
Reilly
SM
,
Gomez
AV
, et al.  
Inhibition of IKKɛ and TBK1 improves glucose control in a subset of patients with type 2 diabetes
.
Cell Metab.
2017
;
26
(
1
):
157
170.e7
.

463.

Open-Label Study to Evaluate MN-001 on HDL and triglyceride in NASH & NAFLD Subjects
. ClinicalTrials.gov identifier
NCT02681055
. https://clinicaltrials.gov/ct2/show/NCT02681055. Accessed
September 16, 2019
.

464.

Safety and Efficacy Study of Orally Administered DS102 in Healthy Subjects
. ClinicalTrials.gov identifier
NCT02673593
. https://clinicaltrials.gov/ct2/show/NCT02673593. Accessed
September 16, 2019
.

465.

Safety and Efficacy Study of Orally Administered DS102 in Patients With NAFLD
. ClinicalTrials.gov identifier
NCT02941549
. https://clinicaltrials.gov/ct2/show/NCT02941549. Accessed
September 16, 2019
.

466.

Different Doses of BI 1467335 Compared to Placebo in Patients With Clinical Evidence of NASH
. ClinicalTrials.gov identifier
NCT03166735
. https://clinicaltrials.gov/ct2/show/NCT03166735. Accessed
September 16, 2019
.

467.

A Study to Evaluate the Safety, Tolerability and Efficacy of MSDC 0602K in Patients With NASH (EMMINENCE)
. ClinicalTrials.gov identifier
NCT02784444
. https://clinicaltrials.gov/ct2/show/NCT02784444. Accessed
September 16, 2019
.

468.

Harrison
 
S
,
Neff
G
,
Iwashita
J
, et al.  
Six month interim results of MSDC-0602 K in a large phase 2b NASH study demonstrate significant improvement in liver enzymes and glycemic control (NCT02784444) (PS-111)
.
J Hepatol.
2019
;
70
(
1
):
e70
.

469.

Lawitz
 
EJ
,
Coste
A
,
Poordad
F
, et al.  
Acetyl-CoA carboxylase inhibitor GS-0976 for 12 weeks reduces hepatic de novo lipogenesis and steatosis in patients with nonalcoholic steatohepatitis
.
Clin Gastroenterol Hepatol.
2018
;
16
(
12
):
1983
1991.e3
.

470.

Loomba
 
R
,
Kayali
Z
,
Noureddin
M
, et al.  
GS-0976 reduces hepatic steatosis and fibrosis markers in patients with nonalcoholic fatty liver disease
.
Gastroenterology.
2018
;
155
(
5
):
1463
1473.e6
.

471.

Safadi
 
R
,
Konikoff
FM
,
Mahamid
M
, et al. ;
FLORA Group
.
The fatty acid-bile acid conjugate Aramchol reduces liver fat content in patients with nonalcoholic fatty liver disease
.
Clin Gastroenterol Hepatol.
2014
;
12
(
12
):
2085
2091.e1
.

472.

Koutnikova
 
H
,
Genser
B
,
Monteiro-Sepulveda
M
, et al.  
Impact of bacterial probiotics on obesity, diabetes and non-alcoholic fatty liver disease related variables: a systematic review and meta-analysis of randomised controlled trials
.
BMJ Open.
2019
;
9
(
3
):
e017995
.

473.

Dobbins
 
R
,
Gholam
P
,
Ertel
S
, et al.  
Solithromycin treatment for 13 weeks improves histological parameters in nonalcoholic steatohepatitis: results of an ongoing pilot study (SAT-299)
.
J Hepatol.
2017
;
66
(
1
):
S586
.

474.

Fecal Microbiota Transplantation for the Treatment of Non-alcoholic Steatohepatitis (FMT-NASH)
. ClinicalTrials.gov identifier
NCT03803540
. https://clinicaltrials.gov/ct2/show/NCT03803540. Accessed
September 16, 2019
.

475.

Fecal Microbiota Transplantation (FMT) in Nonalcoholic Steatohepatitis(NASH): A Pilot Study
. ClinicalTrials.gov identifier
NCT02469272
. https://clinicaltrials.gov/ct2/show/NCT02469272. Accessed
September 16, 2019
.

476.

Fecal Microbiota Therapy Versus Standard Therapy in Decompensated NASH Related Cirrhosis: A Randomized Controlled Trial
. ClinicalTrials.gov identifier
NCT02868164
. https://clinicaltrials.gov/ct2/show/NCT02868164. Accessed
September 16, 2019
.

477.

Abel
 
T
,
Fehér
J
,
Dinya
E
,
Eldin
MG
,
Kovács
A
.
Safety and efficacy of combined ezetimibe/simvastatin treatment and simvastatin monotherapy in patients with non-alcoholic fatty liver disease
.
Med Sci Monit.
2009
;
15
(
12
):
MS6
M11
.

478.

Nelson
 
A
,
Torres
DM
,
Morgan
AE
,
Fincke
C
,
Harrison
SA
.
A pilot study using simvastatin in the treatment of nonalcoholic steatohepatitis: a randomized placebo-controlled trial
.
J Clin Gastroenterol.
2009
;
43
(
10
):
990
994
.

479.

Kimura
 
Y
,
Hyogo
H
,
Yamagishi
S
, et al.  
Atorvastatin decreases serum levels of advanced glycation endproducts (AGEs) in nonalcoholic steatohepatitis (NASH) patients with dyslipidemia: clinical usefulness of AGEs as a biomarker for the attenuation of NASH
.
J Gastroenterol.
2010
;
45
(
7
):
750
757
.

480.

Hyogo
 
H
,
Tazuma
S
,
Arihiro
K
, et al.  
Efficacy of atorvastatin for the treatment of nonalcoholic steatohepatitis with dyslipidemia
.
Metabolism.
2008
;
57
(
12
):
1711
1718
.

481.

Cioboată
 
R
,
Găman
A
,
Traşcă
D
, et al.  
Pharmacological management of non-alcoholic fatty liver disease: atorvastatin versus pentoxifylline
.
Exp Ther Med.
2017
;
13
(
5
):
2375
2381
.

482.

Han
 
KH
,
Rha
SW
,
Kang
HJ
, et al.  
Evaluation of short-term safety and efficacy of HMG-CoA reductase inhibitors in hypercholesterolemic patients with elevated serum alanine transaminase concentrations: PITCH study (PITavastatin versus atorvastatin to evaluate the effect on patients with hypercholesterolemia and mild to moderate hepatic damage)
.
J Clin Lipidol.
2012
;
6
(
4
):
340
351
.

483.

Hyogo
 
H
,
Ikegami
T
,
Tokushige
K
, et al.  
Efficacy of pitavastatin for the treatment of non-alcoholic steatohepatitis with dyslipidemia: an open-label, pilot study
.
Hepatol Res.
2011
;
41
(
11
):
1057
1065
.

484.

Sanyal
 
AJ
,
Abdelmalek
MF
,
Suzuki
A
,
Cummings
OW
,
Chojkier
M
;
EPE-A Study Group
.
No significant effects of ethyl-eicosapentanoic acid on histologic features of nonalcoholic steatohepatitis in a phase 2 trial
.
Gastroenterology.
2014
;
147
(
2
):
377
384.e1
.

485.

Nogueira
 
MA
,
Oliveira
CP
,
Ferreira Alves
VA
, et al.  
Omega-3 polyunsaturated fatty acids in treating non-alcoholic steatohepatitis: a randomized, double-blind, placebo-controlled trial
.
Clin Nutr.
2016
;
35
(
3
):
578
586
.

486.

Dasarathy
 
S
,
Dasarathy
J
,
Khiyami
A
, et al.  
Double-blind randomized placebo-controlled clinical trial of omega 3 fatty acids for the treatment of diabetic patients with nonalcoholic steatohepatitis
.
J Clin Gastroenterol.
2015
;
49
(
2
):
137
144
.

487.

Li
 
YH
,
Yang
LH
,
Sha
KH
,
Liu
TG
,
Zhang
LG
,
Liu
XX
.
Efficacy of poly-unsaturated fatty acid therapy on patients with nonalcoholic steatohepatitis
.
World J Gastroenterol.
2015
;
21
(
22
):
7008
7013
.

488.

Parker
 
HM
,
Johnson
NA
,
Burdon
CA
,
Cohn
JS
,
O’Connor
HT
,
George
J
.
Omega-3 supplementation and non-alcoholic fatty liver disease: a systematic review and meta-analysis
.
J Hepatol.
2012
;
56
(
4
):
944
951
.

489.

Athyros
 
VG
,
Mikhailidis
DP
,
Didangelos
TP
, et al.  
Effect of multifactorial treatment on non-alcoholic fatty liver disease in metabolic syndrome: a randomised study
.
Curr Med Res Opin.
2006
;
22
(
5
):
873
883
.

490.

Yaghoubi
 
M
,
Jafari
S
,
Sajedi
B
, et al.  
Comparison of fenofibrate and pioglitazone effects on patients with nonalcoholic fatty liver disease
.
Eur J Gastroenterol Hepatol.
2017
;
29
(
12
):
1385
1388
.

491.

Belfort
 
R
,
Berria
R
,
Cornell
J
,
Cusi
K
.
Fenofibrate reduces systemic inflammation markers independent of its effects on lipid and glucose metabolism in patients with the metabolic syndrome
.
J Clin Endocrinol Metab.
2010
;
95
(
2
):
829
836
.

492.

Fabbrini
 
E
,
Mohammed
BS
,
Korenblat
KM
, et al.  
Effect of fenofibrate and niacin on intrahepatic triglyceride content, very low-density lipoprotein kinetics, and insulin action in obese subjects with nonalcoholic fatty liver disease
.
J Clin Endocrinol Metab.
2010
;
95
(
6
):
2727
2735
.

493.

A Study of Pemafibrate in Patients With Nonalcoholic Fatty Liver Disease (NAFLD)
. ClinicalTrials.gov identifier
NCT03350165
. https://clinicaltrials.gov/ct2/show/NCT03350165. Accessed
September 16, 2019
.

494.

Rena
 
G
,
Hardie
DG
,
Pearson
ER
.
The mechanisms of action of metformin
.
Diabetologia.
2017
;
60
(
9
):
1577
1585
.

495.

Green
 
CJ
,
Marjot
T
,
Tomlinson
JW
,
Hodson
L
.
Of mice and men: Is there a future for metformin in the treatment of hepatic steatosis?
Diabetes Obes Metab.
2018
. doi:. [Epub ahead of print]

496.

Home
 
PD
,
Pocock
SJ
,
Beck-Nielsen
H
, et al.  
Rosiglitazone evaluated for cardiovascular outcomes in oral agent combination therapy for type 2 diabetes (RECORD): a multicentre, randomised, open-label trial
.
Lancet.
2009;373(9681):2125–2135
. doi:.

497.

Nissen
 
SE
,
Wolski
K
.
Effect of rosiglitazone on the risk of myocardial infarction and death from cardiovascular causes
.
N Engl J Med.
2007
;
356
(
24
):
2457
2471
.

498.

Spence
 
JD
,
Viscoli
CM
,
Inzucchi
SE
, et al. ;
IRIS Investigators
.
Pioglitazone therapy in patients with stroke and prediabetes: a post hoc analysis of the IRIS randomized clinical trial
.
JAMA Neurol.
2019
;
76
(
5
):
526
535
.

499.

Young
 
LH
,
Viscoli
CM
,
Curtis
JP
, et al. ;
IRIS Investigators
.
Cardiac outcomes after ischemic stroke or transient ischemic attack: effects of pioglitazone in patients with insulin resistance without diabetes mellitus
.
Circulation.
2017
;
135
(
20
):
1882
1893
.

500.

Kernan
 
WN
,
Viscoli
CM
,
Furie
KL
, et al. ;
IRIS Trial Investigators
.
Pioglitazone after ischemic stroke or transient ischemic attack
.
N Engl J Med.
2016
;
374
(
14
):
1321
1331
.

501.

Filipova
 
E
,
Uzunova
K
,
Kalinov
K
,
Vekov
T
.
Pioglitazone and the risk of bladder cancer: a meta-analysis
.
Diabetes Ther.
2017
;
8
(
4
):
705
726
.

502.

Armstrong
 
MJ
,
Houlihan
DD
,
Rowe
IA
, et al.  
Safety and efficacy of liraglutide in patients with type 2 diabetes and elevated liver enzymes: individual patient data meta-analysis of the LEAD program
.
Aliment Pharmacol Ther.
2013
;
37
(
2
):
234
242
.

503.

Armstrong
 
MJ
,
Hull
D
,
Guo
K
, et al.  
Glucagon-like peptide 1 decreases lipotoxicity in non-alcoholic steatohepatitis
.
J Hepatol.
2016
;
64
(
2
):
399
408
.

504.

Pyke
 
C
,
Heller
RS
,
Kirk
RK
, et al.  
GLP-1 receptor localization in monkey and human tissue: novel distribution revealed with extensively validated monoclonal antibody
.
Endocrinology.
2014
;
155
(
4
):
1280
1290
.

505.

Panjwani
 
N
,
Mulvihill
EE
,
Longuet
C
, et al.  
GLP-1 receptor activation indirectly reduces hepatic lipid accumulation but does not attenuate development of atherosclerosis in diabetic male ApoE(-/-) mice
.
Endocrinology.
2013
;
154
(
1
):
127
139
.

506.

Marso
 
SP
,
Daniels
GH
,
Brown-Frandsen
K
, et al. ;
LEADER Steering Committee; LEADER Trial Investigators
.
Liraglutide and cardiovascular outcomes in type 2 diabetes
.
N Engl J Med.
2016
;
375
(
4
):
311
322
.

507.

Marso
 
SP
,
Bain
SC
,
Consoli
A
, et al. ;
SUSTAIN-6 Investigators
.
Semaglutide and cardiovascular outcomes in patients with type 2 diabetes
.
N Engl J Med.
2016
;
375
(
19
):
1834
1844
.

508.

Gerstein
 
HC
,
Colhoun
HM
,
Dagenais
GR
, et al.  
Dulaglutide and cardiovascular outcomes in type 2 diabetes (REWIND): a double-blind, randomised placebo-controlled trial
.
Lancet.
2019;394(10193):121–130
. doi:.

509.

Effect of Dulaglutide on Liver Fat in Patients With Type 2 Diabetes and Nonalcoholic Fatty Liver Disease
. ClinicalTrials.gov identifier
NCT03590626
. https://clinicaltrials.gov/ct2/show/NCT03590626. Accessed
September 16, 2019
.

510.

Researching an Effect of GLP-1 Agonist on Liver STeatosis (REALIST)
. ClinicalTrials.gov identifier
NCT03648554
. https://clinicaltrials.gov/ct2/show/NCT03648554. Accessed
September 16, 2019
.

511.

Alam
 
S
,
Ghosh
J
,
Mustafa
G
,
Kamal
M
,
Ahmad
N
.
Effect of sitagliptin on hepatic histological activity and fibrosis of nonalcoholic steatohepatitis patients: a 1-year randomized control trial
.
Hepat Med.
2018
;
10
:
23
31
.

512.

Joy
 
TR
,
McKenzie
CA
,
Tirona
RG
, et al.  
Sitagliptin in patients with non-alcoholic steatohepatitis: a randomized, placebo-controlled trial
.
World J Gastroenterol.
2017
;
23
(
1
):
141
150
.

513.

Neal
 
B
,
Perkovic
V
,
Mahaffey
KW
, et al. ;
CANVAS Program Collaborative Group
.
Canagliflozin and cardiovascular and renal events in type 2 diabetes
.
N Engl J Med.
2017
;
377
(
7
):
644
657
.

514.

Zinman
 
B
,
Wanner
C
,
Lachin
JM
, et al. ;
EMPA-REG OUTCOME Investigators
.
Empagliflozin, cardiovascular outcomes, and mortality in type 2 diabetes
.
N Engl J Med.
2015
;
373
(
22
):
2117
2128
.

515.

Wiviott
 
SD
,
Raz
I
,
Bonaca
MP
, et al. ;
DECLARE–TIMI 58 Investigators
.
Dapagliflozin and cardiovascular outcomes in type 2 diabetes
.
N Engl J Med.
2019
;
380
(
4
):
347
357
.

516.

Bajaj
 
HS
,
Brown
RE
,
Bhullar
L
,
Sohi
N
,
Kalra
S
,
Aronson
R
.
SGLT2 inhibitors and incretin agents: associations with alanine aminotransferase activity in type 2 diabetes
.
Diabetes Metab.
2018
;
44
(
6
):
493
499
.

517.

Shibuya
 
T
,
Fushimi
N
,
Kawai
M
, et al.  
Luseogliflozin improves liver fat deposition compared to metformin in type 2 diabetes patients with non-alcoholic fatty liver disease: a prospective randomized controlled pilot study
.
Diabetes Obes Metab.
2018
;
20
(
2
):
438
442
.

518.

Akuta
 
N
,
Kawamura
Y
,
Watanabe
C
, et al.  
Impact of SGLT2 inhibitor to histological features and glucose metabolism of non-alcoholic fatty liver disease complicated by diabetes mellitus
.
Hepatol Res.
2018
. doi:.

519.

Donnan
 
JR
,
Grandy
CA
,
Chibrikov
E
, et al.  
Comparative safety of the sodium glucose co-transporter 2 (SGLT2) inhibitors: a systematic review and meta-analysis
.
BMJ Open.
2019
;
9
(
1
):
e022577
.

520.

Hamblin
 
PS
,
Wong
R
,
Ekinci
EI
, et al.  
SGLT2 inhibitors increase the risk of diabetic ketoacidosis developing in the community and during hospital admission
.
J Clin Endocrinol Metab.
2019
;
104
(
8
):
3077
3087
.

521.

He
 
YL
,
Haynes
W
,
Meyers
CD
, et al.  
The effects of licogliflozin, a dual SGLT1/2 inhibitor, on body weight in obese patients with or without diabetes
.
Diabetes Obes Metab.
2019
;
21
(
6
):
1311
1321
.

522.

Bril
 
F
,
Portillo Sanchez
P
,
Lomonaco
R
, et al.  
Liver safety of statins in prediabetes or T2DM and nonalcoholic steatohepatitis: post hoc analysis of a randomized trial
.
J Clin Endocrinol Metab.
2017
;
102
(
8
):
2950
2961
.

523.

Nascimbeni
 
F
,
Pellegrini
E
,
Lugari
S
, et al.  
Statins and nonalcoholic fatty liver disease in the era of precision medicine: more friends than foes
.
Atherosclerosis.
2019
;
284
:
66
74
.

524.

Honda
 
Y
,
Kessoku
T
,
Ogawa
Y
, et al.  
Pemafibrate, a novel selective peroxisome proliferator-activated receptor alpha modulator, improves the pathogenesis in a rodent model of nonalcoholic steatohepatitis
.
Sci Rep.
2017
;
7
:
42477
.

525.

Sabatine
 
MS
,
Giugliano
RP
,
Keech
AC
, et al. ;
FOURIER Steering Committee and Investigators
.
Evolocumab and clinical outcomes in patients with cardiovascular disease
.
N Engl J Med.
2017
;
376
(
18
):
1713
1722
.

526.

Ruscica
 
M
,
Ferri
N
,
Macchi
C
, et al.  
Liver fat accumulation is associated with circulating PCSK9
.
Ann Med.
2016
;
48
(
5
):
384
391
.

527.

Wargny
 
M
,
Ducluzeau
PH
,
Petit
JM
, et al.  
Circulating PCSK9 levels are not associated with the severity of hepatic steatosis and NASH in a high-risk population
.
Atherosclerosis.
2018
;
278
:
82
90
.

528.

Jones
 
PH
,
Bays
HE
,
Chaudhari
U
, et al.  
Safety of alirocumab (A PCSK9 Monoclonal Antibody) from 14 randomized trials
.
Am J Cardiol.
2016
;
118
(
12
):
1805
1811
.

529.

Manne
 
V
,
Kowdley
KV
.
Obeticholic acid in primary biliary cholangitis: where we stand
.
Curr Opin Gastroenterol.
2019
;
35
(
3
):
191
196
.

530.

Hameed
 
B
,
Terrault
NA
,
Gill
RM
, et al. ;
NASH CRN
.
Clinical and metabolic effects associated with weight changes and obeticholic acid in non-alcoholic steatohepatitis
.
Aliment Pharmacol Ther.
2018
;
47
(
5
):
645
656
.

531.

Randomized Global Phase 3 Study to Evaluate the Impact on NASH With Fibrosis of Obeticholic Acid Treatment (REGENERATE) Title
. ClinicalTrials.gov identifier
NCT02548351
. .https://clinicaltrials.gov/ct2/show/NCT02548351. Accessed
September 16, 2019
.

532.

Younossi
 
Z
,
Ratziu
V
,
Loomba
R
, et al.  
GS-06-positive results from REGENERATE: a phase 3 international, randomized, placebo-controlled study evaluating obeticholic acid treatment for NASH
.
J Hepatol.
2019
;
70
(
1
):
e5
.

533.

Study Evaluating the Efficacy and Safety of Obeticholic Acid in Subjects With Compensated Cirrhosis Due to Nonalcoholic Steatohepatitis (REVERSE)
. ClinicalTrials.gov identifier
NCT03439254
. https://clinicaltrials.gov/ct2/show/NCT03439254. Accessed
September 16, 2019
.

534.

Evaluating the Safety, Tolerability, and Efficacy of GS-9674 in Participants With Nonalcoholic Steatohepatitis (NASH)
. ClinicalTrials.gov identifier
NCT02854605
. https://clinicaltrials.gov/ct2/show/NCT02854605. Accessed
September 16, 2019
.

535.

Phase 3 Study to Evaluate the Efficacy and Safety of Elafibranor Versus Placebo in Patients With Nonalcoholic Steatohepatitis (NASH) (RESOLVE-IT)
. ClinicalTrials.gov identifier
NCT02704403
. https://clinicaltrials.gov/ct2/show/NCT02704403. Accessed
September 16, 2019
.

536.

Avouac
 
J
,
Konstantinova
I
,
Guignabert
C
, et al.  
Pan-PPAR agonist IVA337 is effective in experimental lung fibrosis and pulmonary hypertension
.
Ann Rheum Dis.
2017
;
76
(
11
):
1931
1940
.

537.

Ruzehaji
 
N
,
Frantz
C
,
Ponsoye
M
, et al.  
Pan PPAR agonist IVA337 is effective in prevention and treatment of experimental skin fibrosis
.
Ann Rheum Dis.
2016
;
75
(
12
):
2175
2183
.

538.

Wettstein
 
G
,
Luccarini
JM
,
Poekes
L
, et al.  
The new-generation pan-peroxisome proliferator-activated receptor agonist IVA337 protects the liver from metabolic disorders and fibrosis
.
Hepatol Commun.
2017
;
1
(
6
):
524
537
.

539.

Jain
 
MR
,
Giri
SR
,
Bhoi
B
, et al.  
Dual PPARα/γ agonist saroglitazar improves liver histopathology and biochemistry in experimental NASH models
.
Liver Int.
2018
;
38
(
6
):
1084
1094
.

540.

Saroglitazar Magnesium in Patients With Nonalcoholic Fatty Liver Disease and/or Nonalcoholic Steatohepatitis (EVIDENCES IV)
. ClinicalTrials.gov identifier
NCT03061721
. https://clinicaltrials.gov/ct2/show/NCT03061721. Accessed
September 16, 2019
.

541.

AURORA: Phase 3 Study for the Efficacy and Safety of CVC for the Treatment of Liver Fibrosis in Adults With NASH
. ClinicalTrials.gov identifier
NCT03028740
. https://clinicaltrials.gov/ct2/show/NCT03028740. Accessed
September 16, 2019
.

542.

Miller
 
ER
 3rd ,
Pastor-Barriuso
R
,
Dalal
D
,
Riemersma
RA
,
Appel
LJ
,
Guallar
E
.
Meta-analysis: high-dosage vitamin E supplementation may increase all-cause mortality
.
Ann Intern Med.
2005
;
142
(
1
):
37
46
.

543.

Oliver
 
CJ
,
Myers
SP
.
Validity of a cochrane systematic review and meta-analysis for determining the safety of vitamin E
.
BMC Complement Altern Med.
2017
;
17
(
1
):
408
.

544.

Schürks
 
M
,
Glynn
RJ
,
Rist
PM
,
Tzourio
C
,
Kurth
T
.
Effects of vitamin E on stroke subtypes: meta-analysis of randomised controlled trials
.
Bmj.
2010
;
341
:
c5702
.

545.

Rodriguez
 
C
,
Jacobs
EJ
,
Mondul
AM
,
Calle
EE
,
McCullough
ML
,
Thun
MJ
.
Vitamin E supplements and risk of prostate cancer in U.S. men
.
Cancer Epidemiol Biomarkers Prev.
2004
;
13
(
3
):
378
382
.

546.

Zhou
 
M
,
Learned
RM
,
Rossi
SJ
,
DePaoli
AM
,
Tian
H
,
Ling
L
.
Engineered FGF19 eliminates bile acid toxicity and lipotoxicity leading to resolution of steatohepatitis and fibrosis in mice
.
Hepatol Commun.
2017
;
1
(
10
):
1024
1042
.

547.

Tomlinson
 
E
,
Fu
L
,
John
L
, et al.  
Transgenic mice expressing human fibroblast growth factor-19 display increased metabolic rate and decreased adiposity
.
Endocrinology.
2002
;
143
(
5
):
1741
1747
.

548.

Rinella
 
ME
,
Trotter
JF
,
Abdelmalek
MF
, et al.  
Rosuvastatin improves the FGF19 analogue NGM282-associated lipid changes in patients with non-alcoholic steatohepatitis
.
J Hepatol.
2019
;
70
(
4
):
735
744
.

549.

Study of NGM282 in Patients With Nonalcoholic Steatohepatitis (NASH)
. ClinicalTrials.gov identifier
NCT02443116
. https://clinicaltrials.gov/ct2/show/NCT02443116. Accessed
September 16, 2019
.

550.

Barb
 
D
,
Bril
F
,
Kalavalapalli
S
,
Cusi
K
.
Plasma fibroblast growth factor 21 is associated with severity of nonalcoholic steatohepatitis in patients with obesity and type 2 diabetes
.
J Clin Endocrinol Metab.
2019
;
104
(
8
):
3327
3336
.

551.

An Investigational Study to Evaluate Experimental Medication BMS-986036 in Participants With Different Levels of Kidney Function
. ClinicalTrials.gov identifier
NCT03674476
. https://www.clinicaltrials.gov/ct2/show/NCT03674476. Accessed
September 16, 2019
.

552.

A Sub-study of BMS-986036 in Subjects With Non-Alcoholic Steatohepatitis (NASH)
. ClinicalTrials.gov identifier
NCT03400163
. https://clinicaltrials.gov/ct2/show/NCT03400163. Accessed
September 16, 2019
.

553.

A Study of Experimental Medication BMS-986036 Given to Healthy Participants
. ClinicalTrials.gov identifier
NCT03445208
. https://clinicaltrials.gov/ct2/show/NCT03445208. Accessed
September 16, 2019
.

554.

A Study of BMS-986036 in Subjects With Non-Alcoholic Steatohepatitis (NASH)
. ClinicalTrials.gov identifier
NCT02413372
. https://clinicaltrials.gov/ct2/show/NCT02413372. Accessed
September 16, 2019
.

555.

A Study to Evaluate BMS-986036 in Obese Adults With Type-2 Diabetes
. ClinicalTrials.gov identifier
NCT02097277
. https://clinicaltrials.gov/ct2/show/NCT02097277. Accessed
September 16, 2019
.

556.

A Study to Test the Effects of BMS-986036 on the Body in Healthy Japanese and Non-Japanese Subjects
. ClinicalTrials.gov identifier
NCT03198182
. https://clinicaltrials.gov/ct2/show/NCT03198182. Accessed
September 16, 2019
.

557.

A Study of Experimental Medication BMS-986036 in Adults With Nonalcoholic Steatohepatitis (NASH) and Stage 3 Liver Fibrosis (FALCON 1)
. ClinicalTrials.gov identifier
NCT03486899
. https://clinicaltrials.gov/ct2/show/NCT03486899. Accessed
September 16, 2019
.

558.

A Study of Experimental Medication BMS-986036 in Adults With Nonalcoholic Steatohepatitis (NASH) and Liver Cirrhosis
. ClinicalTrials.gov identifier
NCT03486912
. https://clinicaltrials.gov/ct2/show/NCT03486912. Accessed
September 16, 2019
.

559.

Taub
 
R
,
Chiang
E
,
Chabot-Blanchet
M
, et al.  
Lipid lowering in healthy volunteers treated with multiple doses of MGL-3196, a liver-targeted thyroid hormone receptor-β agonist
.
Atherosclerosis.
2013
;
230
(
2
):
373
380
.

560.

Kelly
 
MJ
,
Pietranico-Cole
S
,
Larigan
JD
, et al.  
Discovery of 2-[3,5-dichloro-4-(5-isopropyl-6-oxo-1,6-dihydropyridazin-3-yloxy)phenyl]-3,5-dioxo-2,3,4,5-tetrahydro[1,2,4]triazine-6-carbonitrile (MGL-3196), a highly selective thyroid hormone receptor β agonist in clinical trials for the treatment of dyslipidemia
.
J Med Chem.
2014
;
57
(
10
):
3912
3923
.

561.

Phase 2 Study of MGL-3196 in Patients With Non-Alcoholic Steatohepatitis (NASH)
. ClinicalTrials.gov identifier
NCT02912260
. https://clinicaltrials.gov/ct2/show/NCT02912260. Accessed
September 16, 2019
.

562.

Dobrilla
 
G
,
De Pretis
G
,
Piazzi
L
, et al.  
Comparison of once-daily bedtime administration of famotidine and ranitidine in the short-term treatment of duodenal ulcer. A multicenter, double-blind, controlled study
.
Scand J Gastroenterol Suppl.
1987
;
134
:
21
28
.

563.

Younossi
 
ZM
,
Stepanova
M
,
Lawitz
E
, et al.  
Improvement of hepatic fibrosis and patient-reported outcomes in non-alcoholic steatohepatitis treated with selonsertib
.
Liver Int.
2018
;
38
(
10
):
1849
1859
.

564.

Emricasan, a Caspase Inhibitor, for Evaluation in Subjects With Non-Alcoholic Steatohepatitis (NASH) Fibrosis (ENCORE-NF)
. ClinicalTrials.gov identifier
NCT02686762
. https://clinicaltrials.gov/ct2/show/NCT02686762. Accessed
September 16, 2019
.

565.

Ratziu
 
V
,
de Guevara
L
,
Safadi
R
, et al.  
One-year results of the global phase 2b randomized placebocontrolled ARREST trial of aramchol, a stearoyl CoA desaturase modulator in NASH patients
.
Hepatology.
2018
;
68
(
Suppl 1(LB-5
).

566.

Cobbold
 
JFL
,
Atkinson
S
,
Marchesi
JR
, et al.  
Rifaximin in non-alcoholic steatohepatitis: an open-label pilot study
.
Hepatol Res.
2018
;
48
(
1
):
69
77
.

567.

Transplantation of Microbes for Treatment of Metabolic Syndrome & NAFLD (FMT)
. ClinicalTrials.gov identifier
NCT02496390
. https://clinicaltrials.gov/ct2/show/NCT02496390. Accessed
September 16, 2019
.

568.

Fecal Microbiota Therapy Versus Standard Therapy in NASH Related Cirrhosis
. ClinicalTrials.gov identifier
NCT02721264
. https://clinicaltrials.gov/ct2/show/NCT02721264. Accessed
September 16, 2019
.

569.

Spalinger
 
MR
,
Atrott
K
,
Baebler
K
, et al.  
Administration of the hyper-immune bovine colostrum extract IMM-124E ameliorates experimental murine colitis
.
J Crohns Colitis.
2018;13(6):785–797
. doi:.

570.

Matsuda
 
K
,
Gutierrez
J
,
Hassanein
T
,
Iwaki
Y
.
The anti-fibrotic agent, tipelukast (MN-001) reduces serum triglyceride significantly in non-alcoholic steatohepatitis and non-alcoholic fatty liver disease patients with hypertriglyceridemia after 4 weeks of treatment, an interim analysis of ongoing clin
.
J Hepatol.
2018
;
68
:
S586
.

571.

McCommis
 
KS
,
Hodges
WT
,
Brunt
EM
, et al.  
Targeting the mitochondrial pyruvate carrier attenuates fibrosis in a mouse model of nonalcoholic steatohepatitis
.
Hepatology.
2017
;
65
(
5
):
1543
1556
.

572.

Loomba
 
R
,
Morgan
E
,
Fung
M
,
Watts
L
,
Geary
R
,
Bhanot
S
.
PS-106-An international, randomized, placebo-controlled phase 2 trial demonstrates novel effects of DGAT2 antisense inhibition in reducing steatosis without causing hypertriglyceridemia in T2DM patients
.
J Hepatol.
2019
;
70
(
1
):
e67
e68
.

573.

Bar-Yehuda
 
S
,
Stemmer
SM
,
Madi
L
, et al.  
The A3 adenosine receptor agonist CF102 induces apoptosis of hepatocellular carcinoma via de-regulation of the Wnt and NF-kappaB signal transduction pathways
.
Int J Oncol.
2008
;
33
(
2
):
287
295
.

574.

Stemmer
 
SM
,
Benjaminov
O
,
Medalia
G
, et al.  
CF102 for the treatment of hepatocellular carcinoma: a phase I/II, open-label, dose-escalation study
.
Oncologist.
2013
;
18
(
1
):
25
26
.

575.

A Study of the Efficacy and Safety of CF102 in the Treatment of Non-Alcoholic Fatty Liver Disease
. ClinicalTrials.gov identifier
NCT02927314
. https://clinicaltrials.gov/ct2/show/NCT02927314. Accessed
September 16, 2019
.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model)