eISSN: 2449-8238
ISSN: 2392-1099
Clinical and Experimental Hepatology
Current issue Archive Manuscripts accepted About the journal Editorial board Abstracting and indexing Subscription Contact Instructions for authors Ethical standards and procedures
Editorial System
Submit your Manuscript
SCImago Journal & Country Rank
3/2018
vol. 4
 
Share:
Share:
Review paper

A systematic review of the present and future of non-alcoholic fatty liver disease

Christiana Lucas
,
Georgia Lucas
,
Nicholas Lucas
,
Joanna Krzowska-Firych
,
Krzysztof Tomasiewicz

Clin Exp HEPATOL 2018; 4, 3: 165–174
Online publish date: 2018/09/10
Article file
- A systematic review.pdf  [0.12 MB]
Get citation
 
PlumX metrics:
 

Introduction

Non-alcoholic fatty liver disease (NAFLD) is estimated to affect up to 30% of the population in developed countries [1]. This makes it the most common cause of chronic liver disease in the Western world, and it is now projected to become a leading indication for liver transplantation, superseding hepatitis C [2]. NAFLD is a clinico-histopathological entity that includes a spectrum of conditions that are histologically characterized by macrovesicular hepatic steatosis. NAFLD, as the name suggests, is caused by factors other than excessive alcohol consumption.
Non-alcoholic steatohepatitis (NASH), an active form of NAFLD, comprises hepatocellular necrosis, liver inflammation, and tissue injury. NASH is associated with rapid progression of fibrosis, which is initially accentuated in zone 3 of the centrilobular location; this pattern of fibrosis is known as chicken wire fibrosis. Eventually, this can progress to the development of cirrhosis, with the potential of a hepatic malignancy, hepatocellular carcinoma (HCC) [3].
Currently, the European Medicines Agency does not have a recommended drug therapy regimen for the management of NASH. The development of new drugs presents challenges due to the slow progression of NASH, the clinically relevant endpoints, and the duration of observation. Therefore, lifestyle modifications remain the first line of therapy, since the reversal of overnutrition and metabolic obesity is known to improve the liver’s condition.
This review article aims to highlight the current developments of NAFLD with regards to pathogenesis, epidemiology, diagnosis, clinical features, and available treatment, including novel targets and therapies.

Pathogenesis

Although the pathogenesis of NAFLD is not yet fully understood, it is believed to be a multifactorial process, which develops in genetically predisposed individuals. NAFLD is attributed to insulin resistance, hepatic lipid accumulation, gut microbiota, sedentary lifestyle, a high fat diet, and obesity [4].
Currently, the multiple parallel-hit hypothesis is used to explain the progression of NAFLD to NASH; this theory focuses on many hits that act in parallel from the gut or adipose tissue to promote liver inflammation.
The first hit is insulin resistance, which is associated with metabolic disturbances. Overnutrition induces chronic inflammation, and also allows for hepatic de-novo lipogenesis that leads to accumulation of excessive triglycerides and free fatty acids, leading to hepatic steatosis [5]. There is an increase in the fatty acid de-novo biosynthesis owing to the liver X receptor (LXR) nuclear receptor; which is a fundamental regulator of cholesterol, fatty acid and glucose homeostasis. LXR induces the expression of target genes such as carbohydrate regulatory element binding protein (ChREBP), sterol regulatory element binding protein 1c (SREBP-1c), and fatty acid synthase [6, 7].
The presence of insulin resistance is an independent predictor of advanced fibrosis in patients with NAFLD; it causes increased cellular uptake of free fatty acids via the CD36 receptor. CD36 is a member of the class B scavenger receptors, and it binds to various ligands such as long chain fatty acids, oxidized low-density lipoproteins and native lipoproteins [8]. The downregulation of the production and secretion of adipokines and inflammatory cytokines is a consequence of adipose tissue dysfunction, resulting from insulin resistance [9].
Adipokines have an important role in the homeostasis of glucose and lipid metabolism. When there is resistance to leptin, an adipokine, there is associated triglyceride accumulation in non-adipose organs such as the liver, muscle and pancreas. Adiponectin levels correlate positively with high-density lipoprotein cholesterol (HDL-C) and negatively with triglycerides. Hypoadiponectinemia leads to a raised level of several inflammatory markers as well as causing the promotion of visceral fat accumulation. As the severity of the NAFLD increases, there is an increase in leptin levels and a decrease in adiponectin levels, leading to liver fat loss, which is seen in cirrhosis and advanced fibrosis [10].
Hepatokines are secreted by the liver and play an important role in NAFLD; usually in NAFLD patients, there is increased secretion of hepatokines with upregulation of gluconeogenesis, decreased glycogen synthesis and inhibition of insulin signaling. The most important hepatokines include fetuin-A, fibroblast growth factor-21 (FGF-21), selenoprotein P, sex hormone-binding globulin (SHBG), angiopoietin-related growth factor, and leukocyte derived chemotaxin 2 (LECT2). Fetuin-A is a biomarker of NAFLD, and is increased in NAFLD patients, this hepatokine suppresses the auto­phosphorylation of insulin receptors in the skeletal muscles and liver, as well as promoting inflammatory cytokines in monocytes and adipocytes whilst simultaneously activating Toll-like receptors (TLRs) 4 [11].
Bile acids are amphipathic steroid molecules which emulsify and absorb dietary fat, cholesterol, and fat soluble vitamins. The farnesoid X receptor (FXR), mainly expressed in the liver, intestine, kidney and adipose tissue, is involved in the regulation of cholesterol metabolism via regulation of bile acid synthesis and conjugation transport; bile acids are endogenous FXR ligands with a high affinity to FXR.
Toll-like receptors and NLRP3 inflammasomes are implicated in inducing liver inflammation via the nonspecific immune system. TLRs are predominantly expressed on sentinel cells, with TLR4 mediating the recruitment of monocytes and macrophages to the liver [12].
Subsequent initiation of the inflammatory cascade stimulates hepatic cells to produce proinflammatory cytokines. NLRP3 is activated by lysosomal damage, mitochondrial dysfunction, and oxidative stress via the recruitment of an apoptosis-associated protein. Oxidative stress is one of the major mechanisms that lead to NASH, due to the elevated production of reactive oxygen species (ROS). ROS induce lipid peroxidation causing inflammation and fibrogenesis via activation of hepatic stellate cells, as well as inhibiting hepatocyte secretion of very-low-density lipoproteins (LDL); this results in liver fat accumulation. NASH with fibrosis is a more severe subset of NAFLD and is associated with a worse prognosis of liver failure [13, 14].
There is a buildup of damaged cellular products deposited in the hepatocytes due to the defective autophagic function. This promotes inflammation and endoplasmic reticulum (ER) stress. ER stress constitutes a cellular process that is triggered by pathologies which disturb the folding of proteins [15]. Hepatocyte apoptosis is a consequence of liver injury in NASH, and therefore the extent of apoptotic cell death is closely associated with the severity of NASH. This hence enables the development of biomarkers that reflect this process [16].
NAFLD is a polygenic and heritable disease according to evidence from population-based and familial-aggregation studies, including twin studies. Single nucleotide polymorphisms have been identified at several loci, and there has been extensive knowledge acquired on the role of patatin-like phospholipase domain-containing protein 3 (PNPLA3). PNPLA3 is believed to play a role in the export of very-low-density lipoprotein (vLDL), and hepatic cell lipid droplet remodeling; this gene may have a polymorphism that results in the conversion of isoleucine into methionine amino acid [17].
A variant of PNPLA3 is rs738409; it is associated with accumulation of fat in the liver, and GG homozygotes have a 73% increase of lipid fat content [18]. This genotype is linked to a 3.24-fold increased risk of a more pronounced necroinflammatory score, and a 3.2-fold increased risk of developing fibrosis [19]. Rs738409 explains a fraction of the sexual dimorphism associated with NAFLD, as a remarkably higher effect was demonstrated in women. Transmembrane 6 superfamily 2 (TM6SF2) affects the secretion of triglyceride-rich lipoproteins (TRLs) and the hepatic lipid droplet content and also aids in liver fat metabolism [20]. The variant rs58542926, on the C allele, exerts a notable effect in modulating lipid traits such as total cholesterol, circulating low-density lipoprotein cholesterol (LDL-C), and triglycerides; these are associated with a higher cardiovascular risk [21]. The T allele of this genotype is particularly associated with NAFLD [22].
There are 7 categories of genes linked with NAFLD which are demonstrated in Table 1 [23, 24]. Patatin-like phospholipase domain-containing protein 3 (PNPLA3), transmembrane 6 superfamily 2 (TM6SF2), nuclear receptor subfamily 1 group I member 2 (NR1I2), peroxisome proliferator-activated receptor alpha (PPAR-alpha), phosphatidylethanolamine N-methyltransferase (PEMT), microsomal triglyceride transfer protein (MTTP), apolipoprotein C-III (APOC3), apolipoprotein E (ApoE), ectonucleotide pyrophosphatase/phosphodiesterase 1 (ENPP1), insulin receptor substrate 1 (IRS-1), glucokinase regulator (GCKR), solute carrier family 2, facilitated glucose transporter member 1 (SLC2A1), ghrelin O acyl transferase (GOAT), transcription factor 7-like 2 (TCF7L2), peroxisome proliferator-activated receptor gamma (PPARG), bile acyl-CoA synthetase (SLC27A5), fatty acid desaturase 1 (FADS1), lipin-1 (LPIN1), human hemochromatosis protein (HFE), glutamate-cysteine ligase catalytic (GCLC), glutamate-cysteine ligase regulatory subunit (GCLM), ATP-binding cassette sub-family C member 2 (ABCC2), superoxide dismutase 2, mitochondrial (SOD2), Toll-like receptor 4 (TLR4), cluster of differentiation 14 (CD14), tumor necrosis factor (TNF), interleukin 6 (IL-6), angiotensin II receptor type 1 (AGTR1), Kruppel-like factor 6 (KLF6).

Epidemiology

Currently, due to the clinical nature of NAFLD, there is limited availability of epidemiological data to determine its incidence rates. Nevertheless, NAFLD is one of the most common causes of chronic liver diseases and it is a growing problem in the Western world.
In a major European study, a prevalence rate in those with both type 2 diabetes mellitus (T2DM) and NAFLD was reported to be 42.6-69.5% [25]. There are limited studies investigating the incident rates of NAFLD in general populations. A study in Italy conducted over an 8.5-year follow-up period diagnosed NAFLD via ultrasonography at a rate of 18.5 per 1000 person-years [26]. In a UK study, NAFLD accounted for 26.4% of cases, and it was found to be the most common cause of having asymptomatic abnormal liver biochemistry [27]. A study in Greece revealed that in autopsied cases of ischemic heart disease or deaths due to traffic accidents, when hepatitis B and any known liver diseases were excluded, NAFLD was discovered in 31% of cases, with 40% of cases having NASH [28].
In the general population of Poland, NAFLD is documented at a rate of 37.2%, with prevalence rates peaking at 51.4% in those aged 65-70; the prevalence subsequently decreases with advancing age [29]. A further study conducted in Poland observed NAFLD in 78% of all obese individuals and documented an increasing trend. This study concluded that there was a strong indication to raise awareness amongst obese individuals about the influence of their lifestyle on their health status [30].

Risk factors

In 1980, the term NASH was first described by Ludwig and his team when a group of morbidly obese patients developed liver failure following surgical jejunoileal bypass [31]. NAFLD is now recognized as a hepatic manifestation of the metabolic syndrome, which is in turn the most common risk factor for NAFLD. Metabolic syndrome is a cluster of cardiovascular risk factors that include the presence of 3 or more of the following criteria which can be seen in Table 2 [32].
Western diets are associated with a greater likelihood for the development of metabolic syndrome, and progression to NAFLD. This diet may include high intakes of red and processed meats, refined sugars, saturated fats, refined grains, high-fat dairy, and sugar-laden fructose drinks [33, 34]. Statistically, 30% of obese patients have a fatty liver, and up to 90% of morbidly obese patients with a BMI > 35 are diagnosed with NAFLD [35]. Obstructive sleep apnea is associated with obesity, and it is hypothesized that intermittent hypoxia causes NASH [36]. The severity of NAFLD is augmented through reduced physical activity, since sedentary behavior is associated with the development of NAFLD and NASH [37]. A further lifestyle risk factor is smoking, as tobacco is known to increase susceptibility to the development of insulin resistance [38]. Sexual dimorphism of NAFLD exists with a higher incidence in males as well as postmenopausal females.
Additionally, early menarche may lead to a higher risk, which is primarily due to the accumulation of adipocytes. Whilst fertile, the risk of liver fibrosis in females is reduced, but there is an increased risk of severe hepatocyte injury and inflammation, in comparison to men and postmenopausal status. During the female reproductive years, a common syndrome that encompasses insulin resistance and obesity is polycystic ovarian syndrome (PCOS); PCOS sufferers can be hyperandrogenic, which leads to a higher risk of developing NAFLD [39].
Moreover, ethnicity is seen as a risk factor, with Hispanic individuals being more susceptible than their Caucasian counterparts. The lowest susceptibility of NAFLD is observed in black individuals [40]. Additionally, NAFLD can be caused by medications, which include amiodarone, nucleoside analogues, aspirin, oestrogens, glucocorticosteroids, methotrexate, tamoxifen and tetracycline [41]. Other risk factors include total parenteral nutrition, severe anemia, inflammatory bowel disease, inborn errors of metabolism, barium salts, chromates, thallium, and phosphorus.

Diagnosis

Typically, NAFLD is clinically silent, with diagnoses being incidental due to abnormal liver enzymes or imaging; most commonly it results from steatosis. Once NAFLD is suspected, the diagnosis is confirmed after other potential causes of steatosis have been excluded; alcoholic hepatitis is clinically indistinguishable from NASH, and therefore it must be ruled out in order to establish a diagnosis. It is necessary to exclude any significant alcohol consumption, which is usually considered as consumption of more than 20 γ daily [42]. The clinical history, dietary history, medication use, occupational exposure to organic solvents and a family history of liver disease should all be investigated. Anti-smooth muscle antibodies and antinuclear antibody (ANA) are often seen in patients with NASH. This often represents a nonspecific antibody response that is not related to the severity or pattern of injury on liver biopsy [43].
Transabdominal ultrasonography is the preferred first line imaging investigation for the diagnosis of hepatic steatosis due to it being inexpensive, non-invasive and widely accessible. On the ultrasound, there is normally a visual decrease of the vascular margins, a loss of definition of the diaphragm, hepatomegaly, and hyperechogenicity of the liver parenchyma, in addition to focal fat deposition on the hyperechoic area. Transabdominal ultrasonography is very effective if the steatosis of hepatocytes is not less than 31% [44]. FibroScan is non-invasive ultrasound-based elastography for detecting liver fibrosis in both adults and children; the result is expressed as kilopascals (kPa), and this transient elastography technique takes less than five minutes to complete [45].
A liver biopsy is not always required to diagnose NAFLD, but it still remains the gold standard examination since it can be used to distinguish steatohepatitis from simple steatosis, provide an assessment of the extent of necroinflammatory activity, and visualize fibrotic and architectural alterations. The most widely used histological grading and staging system for NAFLD is the NAFLD Activity Score (NAS), which can be seen in Table 3 [46]. The SAF score, which includes the assessment of steatosis (S), activity (A) and fibrosis (F), has made it easier to identify a subset of NAFLD: NASH [5]. The histopathological features of NAFLD include lobular and portal inflammation, steatosis, hepatocellular ballooning, glycogenated nuclei, apoptotic hepatocytes (acidophil bodies), deposition, megamitochondria, Mallory-Denk bodies, and fibrosis with the characteristic pattern focused on the perisinusoidal/pericellular area. This fibrotic pattern is known as chicken wire fibrosis, and it typically originates in zone 3 for adults [47].
A score of ≥ 5 with steatosis and hepatocyte ballooning is generally considered diagnostic of NASH, but patients can still have NASH with lower NAS scores if there is presence of steatosis and hepatocyte balloon [46].

Symptoms

In the absence of a chance diagnosis, NAFLD usually presents asymptomatically until liver decompensation commences. However, sometimes signs and symptoms can occur; these are illustrated in Table 4. The most prevalent initial symptoms are fatigue and malaise, which do not seem to correlate with the severity of disease [48]. Hepatosplenomegaly presents in up to 50% of patients. However, as the disease progresses, the liver shrinks in size, whereas the spleen continues to enlarge [49]. Pruritus can also be seen as a result of increased bilirubin levels and jaundice, but pruritus can occur when bilirubin levels are normal. A Cruveilhier-Baumgarten murmur, which is a venous hymn in the epigastric region, may also be heard. There may be presence of hypotension due to reduced total systemic vascular resistance.

Available treatments

A fundamental part of NASH management involves lifestyle modification. The European Association for the Study of the Liver (EASL) recommends that a 7% weight reduction may be adequate in resolving steatosis and inflammation, whereas the American Association for the Study of Liver Diseases (AASLD) states that a decrease of at least 10% may resolve mild to moderate fibrosis, and improve necroinflammation [50, 51]. A large study involving 154 NAFLD patients demonstrated that a lifestyle modification over a one-year period resulted in remissions of 64% in the lifestyle modification group, and only 20% remissions in the control group [52].
One method of lifestyle alteration is via aerobic and resistance training exercise regimens to reduce both liver and visceral fat. Johnson et al. discovered that in previously inactive, overweight or obese NAFLD patients, all exercise doses, regardless of the amount or intensity, were capable of scaling down liver fat and visceral fat. In patients with obesity, exercise is further complicated by musculoskeletal problems that manifest biomechanically [53].
Dietary restriction is recommended in studies by Elias et al., and Lin et al., proving that pursuing a hypo­caloric diet is key in improving NAFLD; a reduction as low as 450 to 1000 kcal/day has proven to be effective and safe. Nutritional recommendations, listed in Table 5, for patients with NAFLD state that carbohydrates should comprise 40-50% of total dietary energy, with an increase in the amount of complex carbohydrates rich in dietary fiber. Furthermore, fat intake should be less than 30% of the daily calorie intake, and there should also be an increase of mono- and polyunsaturated fatty acids [54, 55].
Bariatric surgery (BS) is one of the most reliable options available in yielding a sustained reduction in body weight for individuals classified as severely obese, with a body mass index ≥ 35 kg/m2. Therefore, patients who have severe obesity and NAFLD could benefit from BS due to improvements in glycemic control, quality of life, and long-term survival. Additionally, BS may aid in alleviating some obesity-related comorbidities, such as sleep apnea, which will also increase patients’ exercise tolerance. To date, there have not been any randomized controlled trials examining the effect that BS has on NAFLD patients; thus, BS as an approach for treatment in NAFLD is not definitive. Nonetheless, there has been a meta-analysis including 766 liver biopsies after BS, which showed a resolution rate of 91.6, 81.3 and 65.6% in steatosis, steatohepatitis, and fibrosis respectively. A study involving 109 morbidly obese patients with biopsy-proven NASH, who had undergone BS, showed features of reduced pathology after 1 year of follow-up in 85% of the cohort. Presently, there is a trial observing biomarkers to quantify liver pathology in patients with NASH at baseline and after BS (ClinicalTrials.gov Identifier: NCT03294850) [56].
NAFLD patients have an increased cardiovascular risk (CVR), with the main cause of death in these patients being ischemic heart disease. Therefore, it is paramount to manage and treat any CVR factors; lipid-lowering agents including statins, and some omega-3 polyunsaturated fatty acid supplements are frequently used in patients in order to reduce CVR, hepatic inflammation and low-density lipoprotein cholesterol (LDL-C) levels [57].
The use of statins remains controversial due to statin-induced hepatotoxicity and is sometimes avoided in patients with increased transaminase levels. The Incremental Decrease in End Points Through Aggressive Lipid Lowering (IDEAL) trial, Greek Atorvastatin and Coronary Heart Disease Evaluation (GREACE) trial, and Action to Control Cardiovascular Risk in Diabetes (ACCORD) trial all demonstrated the improvement of liver enzymes in these patients. These trials endorsed the safe use of statins in compensated cirrhosis patients, since the absence of statins in patients with elevated aminotransferases correlated with the worsening of cardiovascular outcomes [58, 59]. Furthermore, 26 randomized controlled trials combined their results and the meta-analysis including more than 1.4 million patients with 4298 cases of HCC revealed that the use of statin was associated with a 37% reduction in HCC [60].
The usefulness of ezetimibe was examined in a 24-month trial with 45 patients newly diagnosed with NAFLD via liver biopsy. The results demonstrated that the drug significantly improved visceral fat areas, fasting insulin, concentration of triglycerides, total cholesterol, mean levels of small LDL and very small LDL, as well as significantly lowering serum alanine aminotransferase (ALT) and C-reactive protein levels. Moreover, the histological features of steatosis grade necroinflammatory grade, ballooning score and NAFLD activity score were notably improved when compared to the baseline [61].
Studies have shown that NAFLD patients documented a reduction in transaminase levels when taking omega-3 polyunsaturated fatty acid supplements. In the Japan EPA Lipid Intervention Study (JELIS), the study group taking both the statin and omega-3 supplement regimen saw a reduction in CVD events by 19%, compared to those who solely took statins [62]. Nicotinic acid is an alternative, but the Atherothrombosis Intervention in Metabolic Syndrome with Low HDL/High triglycerides: Impact on Global Health Outcomes (AIM-HIGH) study revealed that taking nicotinic acid over a 4-year follow-up showed no difference in the incidence rates of CVD events [63].
Peroxisome proliferator-activated receptors (PPARs) are affiliated with cellular proliferation and nutritional metabolism. The randomized placebo controlled Fatty Liver Improvement With Rosiglitazone Therapy (FLIRT) trial resulted in 47% of NASH patients having a significant (> 30%) reduction in steatosis after a period of 12 months of treatment, with 38% achieving normalization of ALT values [65]. The FLIRT 2 extension trial reviewed the patients after a period of 48 months to observe the long-term effects, and found that the ALT levels remained normal, steatosis improved and insulin levels decreased. Saroglitazar, used for diabetic dyslipidemia treatment in mouse models of NASH, was found to reduce both ALT and steatosis, as well as improving liver histology [66]. A phase 2 open-label study of saroglitazar (PRESS VIII) evaluated 32 patients with NASH and after a period of 12 weeks, a result of a 52% decrease in ALT was documented [67]. A prospective experimental dual PPAR α/δ drug is elafibranor, and it has been shown to reverse NASH by preventing the progression of fibrosis; it is currently undergoing a phase III clinical trial, REOLVE-IT, a randomized pivotal double-blind (2 : 1), placebo controlled trial with 250 centers involving 2000 patients worldwide (ClinicalTrials.gov Identifier: NCT02704403). Fibrates are archetypical PPAR-α agonists that have had mainly disappointing results, but fenofibrates are one of the most used fibrates and are used with statins; they have been shown to reduce triglycerides whilst raising HDL-C levels [74].
GLP-1 agonists have been shown to reduce both hepatic steatosis and necroinflammation [66, 67]. In the liraglutide safety and efficacy in patients with NAFLD (LEAN) study, 39% of patients who received liraglutide had resolution of definite NASH at 48 weeks, whereas only 9% of the placebo group experienced resolution [70]. Studies have shown that GLP-1 agonists can reverse the progression of NAFLD indirectly through an incretin effect that improves key parameters, but directly through an effect on lipid metabolism of hepatocytes and inflammation in the liver [71]. A recent study using DDP-4 inhibitors on patients with T2DM, over 6 months, demonstrated an improvement in liver dysfunction; although the effects on NAFLD are unclear, they do deserve some further research [72].
NASH is an inflammatory disorder, and therefore potential strategies that may amend or turn off inflammatory triggers such as cytokines, including tumor necrosis factor (TNF) and chemokines should be investigated. In a non-randomized pilot study, pentoxifylline, a phosphodiesterase inhibitor with antioxidant properties, demonstrated improvements in serum aminotransferases and liver histology, but not in serum TNF. Further controlled trials have established that pentoxifylline failed to reduce hepatocyte ballooning, and had no histological differences; therefore, this drug is not recommended in treatment [73, 74].
A range of illnesses such as obesity and metabolic disorders, including NAFLD, have been associated with the gut microbiome. In a study with NASH patients, it was found that a minute bacterial intestinal overgrowth led to larger intestinal permeability due to a disruption of the gap junctions [75]. A choline-deficient diet can lead to a decrease in vLDL levels and hepatic beta oxidation which results in an inflammatory reaction and accumulation of liver fatty acids. Therefore, a potential therapy could be with the use of pre- and probiotics to manipulate the gut microbes. In 4 randomized control trials of 134 NAFLD/NASH patients receiving probiotics, the results showed that compared to a placebo, there was a decrease in total cholesterol, HDL, ALT, AST and TNF, but there was no change in BMI, glucose or LDLs [76].
Antifibrotic therapy has attained awareness due to fibrosis being a prognostic marker for NAFLD. Lysyl oxidase (LOX) L2 is an enzyme which aids in the cross-linking of extracellular fibrillar collagen I and is thought to be a critical step in the development of fibrosis in the liver [77]. The efficacy and safety of simtuzumab, a human monoclonal anti-LOXL2 antibody, was explored in two phase 2 placebo-controlled studies, which involved patients with NASH cirrhosis, and patients with advanced NASH fibrosis over a 96-week treatment plan. They showed that the drug was safe and well tolerated but had no additional benefits [76].
Another protein that plays a crucial role in organ fibrosis is galectin-3. In an animal model, mice that were treated with this protein became resistant to liver, lung, and kidney fibrosis. In the mouse models with NASH taking a galectin inhibitor, GR-MD-02, the study concluded that there was a significant decrease in fibrosis, and therefore a phase 1 trial for galectin-3 antagonist is being undertaken [79].
Pathological increases in hepatocyte apoptosis in the liver and peripheral tissues have emerged as an important mechanism in the development of NASH. In phase II clinical trials on 38 patients, emricasan, a pan-caspase inhibitor, lowered serum aminotransferase activity, liver injury and fibrosis by inhibiting hepatocytes apoptosis. A placebo-controlled, multicenter, double-blind, randomized trial with NAFLD patients with raised transaminases found the drug to be well tolerated, and the results suggested the use of the drug as a treatment option for NAFLD/NASH patients [80]. Currently, a 72-week phase II trial in NASH patients with fibrosis is being initiated to verify whether there is any histological improvement of NASH resolution with emricasan (ClinicalTrials.gov Identifier: NCT02686762).
Another controversial but attractive new drug is obeticholic acid (OCA), which is a selective agonist of the FXRs due to its 100-fold-higher affinity to FXR in comparison to chenodeoxycholic acid. A recent randomized placebo-controlled trial called FXR Ligand Obeticholic Acid in NASH Treatment (FLINT) demonstrated an improvement in NAFLD patients’ liver histology. The trial showed a decrease in the NAFLD activity score by at least 2 points without a deterioration of fibrosis; 50 patients out of 110 met the primary endpoint at 72 weeks, compared to 23 out of 109 patients in the placebo group. OCA appears to improve hepatic steatosis, inflammation, and fibrosis, and helps to reduce body weight; however, OCA has been shown to aggravate insulin resistance and dyslipidemia. During treatment, pruritus is commonly witnessed, and furthermore, higher rates of pruritus are reported in patients taking OCA therapy for other liver diseases [82].
Due to the increase in the rates and impact of NASH, there is an urgent requirement to develop more effective management strategies, as well as a need for the discovery of biomarkers to aid in the diagnosis and monitoring of NASH patients. A current trial is investigating the HepQuant SHUNT diagnostic kit to look at the monitoring and treatment of NASH (ClinicalTrials.gov identifier NCT03294941). Other studies are investigating the efficacy of Tropifexor (LJN452) in patients with NASH (ClinicalTrials.gov Identifier: NCT02855164).

Conclusions

NAFLD is one of the most common diseases in the general population, and currently it is the most common cause of chronic liver disease in the Western world; NASH is a severe form of NAFLD. Whilst ongoing research has furthered our understanding of NAFLD, the nature of progression of the disease, as well as its impact on morbidity and mortality, is still highly disparate. Furthermore, screening recommendations for NAFLD and NASH are not available due to a lack of a nationally accepted guideline, but they should be considered in those with risk factors for development of the disease. Additionally, available treatments which have been discussed in this systematic review are of limited efficacy, durability and applicability, due to the intricacy of this disease. Today’s emerging challenge is the race to correlate new data with effective treatment regimens to establish a greater understanding of the factors that control the course of NAFLD, and thus enable more appropriate non-invasive prognostic assessments with the ability to focus on at-risk NAFLD populations for tailored individual treatment. In patients with NAFLD, there is also a need for surveillance and management for any complication of cirrhosis which may include HCC. Moreover, the implementation of lifestyle changes is necessary in this population of patients, but this comes with its own difficulties, considering that a rebound to prior unhealthy eating habits and inactivity is common.

Disclosure

The authors report no conflict of interest.

References

1. Torres DM, Harrison SA. Diagnosis and therapy of nonalcoholic steatohepatitis. Gastroenterology 2008; 134: 1682-1698.
2. Calzadilla Bertot L, Adams LA. The natural course of non-alcoholic fatty liver disease. Int J Mol Sci 2016; 17: 774.
3. Sayiner M, Koenig A, Henry L, et al. Epidemiology of nonalcoholic fatty liver disease and nonalcoholic steatohepatitis in the United States and the rest of the world. Clin Liver Dis 2016; 20: 205-214.
4. Buzzetti E, Pinzani M, Tsochatzis EA. The multiple-hit pathogenesis of non-alcoholic fatty liver disease (NAFLD). Metab Clin Exp 2016; 65: 1038-1048.
5. Ndumele CE, Nasir K, Conceiçao RD, et al. Hepatic steatosis, obesity, and the metabolic syndrome are independently and additively associated with increased systemic inflammation. Arterioscler Thromb Vasc Biol 2011; 31: 1927-1932.
6. Kim JY, Song EH, Lee HJ, et al. HBx-induced hepatic steatosis and apoptosis are regulated by TNFR1-and NF-κB-dependent pathways. J Mol Biol 2010; 397: 917-931.
7. Lizuka K, Bruick RK, Liang G, et al. Deficiency of carbohydrate response element-binding protein (ChREBP) reduces lipogenesis as well as glycolysis. Proceedings of the National Academy of Sciences of the United States of America 2004; 101: 7281-7286.
8. Kashyap SR, Ioachimescu AG, Gornik HL, et al. Lipid-induced insulin resistance is associated with increased monocyte expression of scavenger receptor CD36 and internalization of oxidized LDL. Obesity (Silver Spring) 2009; 17: 2142-2148.
9. Drover VA, Nguyen DV, Bastie CC, et al. CD36 mediates both cellular uptake of very long chain fatty acids and their intestinal absorption in mice. J Biol Chem 2008; 283: 13108-13115.
10. Adolph TE, Grander C, Grabherr F, et al. Adipokines and non-alcoholic fatty liver disease: multiple interactions. Int J Mol Sci 2017; 18: 1649.
11. Lebensztejn DM, Flisiak-Jackiewicz M, Białokoz-Kalinowska I, et al. Hepatokines and non-alcoholic fatty liver disease. Acta Biochim Pol 2016; 63: 459-467.
12. Strowig T, Henao-Mejia J, Elinav E, et al. Inflammasomes in health and disease. Nature 2012; 481: 278-286.
13. Heid ME, Keyel PA, Kamga C, et al. Mitochondrial reactive oxygen species induces NLRP3-dependent lysosomal damage and inflammasome activation. J Immunol 2013; 191: 5230-5238.
14. Angulo P, Kleiner DE, Dam-Larsen S, et al. Liver fi brosis, but no other histologic features, is associated with long-term outcomes of patients with nonalcoholic fatty liver disease. Gastroenterology 2015; 149: 389-397.
15. Santos CX, Tanaka LY, Wosniak Jr J, et al. Mechanisms and implications of reactive oxygen species generation during the unfolded protein response: roles of endoplasmic reticulum oxidoreductases, mitochondrial electron transport, and NADPH oxidase. Antioxid Redox Signal 2009; 11: 2409-2427.
16. Wieckowska A, Zein NN, Yerian LM, et al. In vivo assessment of liver cell apoptosis as a novel biomarker of disease severity in nonalcoholic fatty liver disease. Hepatology 2006; 44: 27-33.
17. Romeo S, Kozlitina J, Xing C, et al. Genetic variation in PNPLA3 confers susceptibility to nonalcoholic fatty liver disease. Nat Genet 2008; 40: 1461-1465.
18. Sookoian S, Pirola CJ. Meta‐analysis of the influence of I148M variant of patatin‐like phospholipase domain containing 3 gene (PNPLA3) on the susceptibility and histological severity of nonalcoholic fatty liver disease. Hepatology 2011; 53: 1883-1894.
19. Sookoian S, Pirola CJ. Meta‐analysis of the influence of I148M variant of patatin‐like phospholipase domain containing 3 gene (PNPLA3) on the susceptibility and histological severity of nonalcoholic fatty liver disease. Hepatology 2011; 53: 1883-1894.
20. Coppola N, Rosa Z, Cirillo G, et al. TM6SF2 E167K variant is associated with severe steatosis in chronic hepatitis C, regardless of PNPLA3 polymorphism. Liver Int 2015; 35: 1959-1963.
21. Ding Y, Zhu MA, Wang ZX, et al. Associations of polymorphisms in the apolipoprotein APOA1-C3-A5 gene cluster with acute coronary syndrome. Biomed Res Biotechnol 2012; 2012: 509420.
22. Fabris C, Falleti E, Cussigh A, et al. IL-28B rs12979860 C/T allele distribution in patients with liver cirrhosis: role in the course of chronic viral hepatitis and the development of HCC. J Hepatol 2011; 54: 716-722.
23. Anstee, QM, Day CP. The genetics of nonalcoholic fatty liver disease: spotlight on PNPLA3 and TM6SF2. In Seminars in liver disease 2015; 35: 270-290.
24. Khatib MN, Gaidhane S, Gaidhane AM, et al. Ghrelin O acyl transferase (GOAT) as a novel metabolic regulatory enzyme. J Clin Diagn Res 2015; 9: LE01.
25. 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: 1212-1218.
26. 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: 1387-1391.
27. 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: 234-240.
28. Zois CD, Baltayiannis GH, Bekiari A, et al. Steatosis and steatohepatitis in postmortem material from Northwestern Greece. World J Gastroenterol 2010; 16: 3944-3949.
29. Hartleb M, Barański K, Zejda J, et al. Non‐alcoholic fatty liver (NAFL) and advanced fibrosis in the elderly: results from a community‐based Polish survey. Liver Int 2017; 37: 1706-1714.
30. Kargulewicz A, Stankowiak-Kulpa H, Grzymisławski M. Assessment of the prevalence of nonalcoholic fatty liver disease among obese polish people and the estimation of the knowledge of nutritional recommendations. Nowiny Lekarskie 2012; 81: 611-619.
31. Ludwig J, Viggiano TR, McGill DB, et al. Nonalcoholic steatohepatitis: Mayo Clinic experiences with a hitherto unnamed disease. Mayo Clin Proc 1980; 55: 434-438.
32. Grundy SM, Cleeman JI, Daniels SR, et al. Diagnosis and management of the metabolic syndrome: an American Heart Association/National Heart, Lung, and Blood Institute scientific statement. Circulation 2005; 112: 2735-2752.
33. Anania FA. Non-alcoholic fatty liver disease and fructose: bad for us, better for mice. J Hepatol 2011; 55: 218-220.
34. Lim JS, Mietus-Snyder M, Valente A, et al. The role of fructose in the pathogenesis of NAFLD and the metabolic syndrome. Nat Rev Gastroenterol Hepatol 2010; 7: 251-264.
35. Torres DM, Harrison SA. Diagnosis and therapy of nonalcoholic steatohepatitis. Gastroenterology 2008; 134: 1682-1698.
36. Polotsky VY, Patil SP, Savransky V, et al. Obstructive sleep apnea, insulin resistance, and steatohepatitis in severe obesity. Am J Respir Crit Care Med 2009; 179: 228-234.
37. Hallsworth K, Thoma C, Moore S, et al. Non-alcoholic fatty liver disease is associated with higher levels of objectively measured sedentary behaviour and lower levels of physical activity than matched healthy controls. Frontline Gastroenterol 2015; 6: 44-51.
38. Filozof C, Fernández Pinilla MC, Fernández‐Cruz A. Smoking cessation and weight gain. Obes Rev 2004; 5: 95-103.
39. Ballestri S, Nascimbeni F, Baldelli E, et al. 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: 1291-1326.
40. Anstee QM, Day CP. The genetics of NAFLD. Nat Rev Gastroenterol Hepatol 2013; 10: 645-655.
41. Adams LA, Angulo P. Treatment of non-alcoholic fatty liver disease. Postgrad Med J 2006; 82: 315-322.
42. Angulo P, Keach JC, Batts KP, et al. Independent predictors of liver fibrosis in patients with nonalcoholic steatohepatitis. Hepatology 1999; 30: 1356-1362.
43. Cotler SJ, Kanji K, Keshavarzian A, et al. Prevalence and significance of autoantibodies in patients with non-alcoholic steatohepatitis. J Clin Gastroenterol 2004; 38: 801-804.
44. Saadeh S, Younossi ZM, Remer EM, et al. The utility of radiological imaging in nonalcoholic fatty liver disease. Gastroenterology 2002; 123: 745-750.
45. Mueller S, Sandrin L. Liver stiffness: a novel parameter for the diagnosis of liver disease. Hepat Med 2010; 2: 49-67.
46. Kleiner DE, Brunt EM, Van Natta M, et al. Design and validation of a histological scoring system for nonalcoholic fatty liver disease. Hepatology 2005; 41: 1313-1321.
47. Valenti L, Fracanzani AL, Bugianesi E, et al. HFE genotype, parenchymal iron accumulation, and liver fibrosis in patients with nonalcoholic fatty liver disease. Gastroenterology 2010; 138: 905-912.
48. Angulo P. Nonalcoholic fatty liver disease. N Engl J Med 2002; 346: 1221-1231.
49. Bacon BR, Farahvash MJ, Janney CG, et al. Nonalcoholic steatohepatitis: an expanded clinical entity. Gastroenterology 1994; 107: 1103-1109.
50. Musso G, Cassader M, Rosina F, et al. 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: 885-904.
51. Chalasani N, Younossi Z, Lavine JE, et al. The diagnosis and management of non‐alcoholic fatty liver disease: Practice Guideline by the American Association for the Study of Liver Diseases, American College of Gastroenterology, and the American Gastroenterological Association. Hepatology 2012; 55: 2005-2023.
52. Wong VW, Chan RS, Wong GL, et al. Community-based lifestyle modification programme for non-alcoholic fatty liver disease: a randomized controlled trial. J Hepatol 2013; 59: 536-542.
53. Keating SE, Hackett DA, Parker HM, et al. Effect of aerobic exercise training dose on liver fat and visceral adiposity. J Hepatol 2015; 63: 174-182.
54. Ferolla SM, Silva LC, Ferrari MD, et al. Dietary approach in the treatment of nonalcoholic fatty liver disease. World J Hepatol 2015; 7: 2522-2534.
55. Kargulewicz A, Stankowiak-Kulpa H, Grzymisławski M. Dietary recommendations for patients with nonalcoholic fatty liver disease. Prz Gastroenterol 2014; 9: 18-23.
56. Lassailly G, Caiazzo R, Buob D, et al. Bariatric surgery reduces features of nonalcoholic steatohepatitis in morbidly obese patients. Gastroenterology 2015; 149: 379-388.
57. Francque SM, van der Graaff D, Kwanten WJ. Non-alcoholic fatty liver disease and cardiovascular risk: Pathophysiological mechanisms and implications. J Hepatol 2016; 65: 425-443.
58. Athyros VG, Katsiki N, Karagiannis A, et al. Statins and non-alcoholic steatohepatitis. J Hepatol 2016; 64: 241-242.
59. Abraldes JG, Albillos A, Banares R, et al. Simvastatin lowers portal pressure in patients with cirrhosis and portal hypertension: a randomized controlled trial. Gastroenterology 2009; 136: 1651-1658.
60. Singh S, Singh PP, Singh AG, et al. Statins are associated with a reduced risk of hepatocellular cancer: a systematic review and meta-analysis. Gastroenterology 2013; 144: 323-332.
61. Park H, Shima T, Yamaguchi K, et al. Efficacy of long-term ezetimibe therapy in patients with nonalcoholic fatty liver disease. J Gastroenterol 2011; 46: 101-107.
62. Yokoyama M, Origasa H, Matsuzaki M, et al. Effects of eicosa­pentaenoic acid on major coronary events in hypercholesterolaemic patients (JELIS): a randomised open-label, blinded endpoint analysis. Lancet 2007; 369: 1090-1098.
63. AIM-HIGH Investigators; Boden WE, Probstfield JL, Anderson T, et al. Niacin in patients with low HDL cholesterol levels receiving intensive statin therapy. N Engl J Med 2011; 365: 2255-2267.
64. Yee HF, Lidofsky SD. Acute liver failure. In: Feldman M, Friedman LS, Sleisenger MH (eds.). Sleisenger and Fordtran’s Gastrointestinal and Liver Disease: Pathophysiology, Diagnosis, Management. 7th ed. Saunders, Philadelphia 2002; 1567-1574.
65. Ratziu V, Giral P, Jacqueminet S, et al. Rosiglitazone for nonalcoholic steatohepatitis: one-year results of the randomized placebo-controlled Fatty Liver Improvement with Rosiglitazone Therapy (FLIRT) Trial. Gastroenterology 2008; 135: 100-110.
66. Ratziu V, Charlotte F, Bernhardt C, et al. Long‐term efficacy of rosiglitazone in nonalcoholic steatohepatitis: Results of the fatty liver improvement by rosiglitazone therapy (FLIRT 2) extension trial. Hepatology 2010; 51: 445-453.
67. Jain MR, Giri SR, Trivedi C, et al. Saroglitazar, a novel PPARα/γ agonist with predominant PPARα activity, shows lipid‐lowering and insulin‐sensitizing effects in preclinical models. Pharmacol Res Perspect 2015; 3: e00136.
68. Sathyanarayana P, Jogi M, Muthupillai R, et al. Eff ects of combined exenatide and pioglitazone therapy on hepatic fat content in type 2 diabetes. Obesity (Silver Spring) 2011; 19: 2310-2315.
69. Svegliati-Baroni G, Saccomanno S, Rychlicki C, et al. Glucagon-like peptide-1 receptor activation stimulates hepatic lipid oxidation and restores hepatic signalling alteration induced by a high-fat diet in nonalcoholic steatohepatitis. Liver Int 2011; 31: 1285-1297.
70. Armstrong MJ, Gaunt P, Aithal GP, et al. Liraglutide safety and efficacy in patients with non-alcoholic steatohepatitis (LEAN): a multicentre, double-blind, randomised, placebo-controlled phase 2 study. Lancet 2016; 387: 679-690. 71.
Liu J, Wang G, Jia Y, et al. GLP‐1 receptor agonists: effects on the progression of non‐alcoholic fatty liver disease. Diabetes Metab Res Rev 2015; 31: 329-335. 72.
Kanazawa I, Tanaka KI, SugimotoT. DPP-4 inhibitors improve liver dysfunction in type 2 diabetes mellitus. Med Sci Monit 2014; 20: 1662-1667.
73. Adams LA, Zein CO, Angulo P, Lindor KD. A pilot trial of pentoxifylline in nonalcoholic steatohepatitis. Am J Gastroenterol 2004; 99: 2365-2368.
74.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: 277-286.
75. Miele L, Valenza V, La Torre G, et al. Increased intestinal permeability and tight junction alterations in nonalcoholic fatty liver disease. Hepatology 2009; 49: 1877-1887.
76. Ma YY, Li L, Yu CH, et al. Effects of probiotics on nonalcoholic fatty liver disease: a meta-analysis. World J Gastroenterol 2013; 19: 6911-3618.
77. Van Bergen T, Marshall D, Van de Veire S, et al. The role of LOX and LOXL2 in scar formation after glaucoma surgery. Invest Ophthalmol Visual Sci 2013; 54: 5788-5796.
78. Sanyal A, Abdelmalek MF, Diehl AM, et al. Efficacy and safety of simtuzumab for the treatment of nonalcoholic steatohepatitis with bridging fibrosis or cirrhosis: results of two phase 2b, dose-ranging, randomized, placebo-controlled trials. J Hepatol 2017; 66: S54.
79. Traber PG, Zomer E. Therapy of experimental NASH and fibrosis with galectin inhibitors. PLoS One 2013; 8: e83481.
80. Shiffman M, Freilich B, Vuppalanchi R, et al. LP37: A placebo-controlled, multicenter, double-blind, randomised trial of emricasan in subjects with non-alcoholic fatty liver disease (NAFLD) and raised transaminases. J Hepatol 2015; 62: S282.
81. Agouridis AP, Kostapanos MS, Tsimihodimos V, et al. Effect of rosuvastatin monotherapy or in combination with fenofibrate or ω-3 fatty acids on lipoprotein subfraction profile in patients with mixed dyslipidaemia and metabolic syndrome. Int J Clin Pract 2012; 66: 843-853.
82. Neuschwander-Tetri BA, Loomba R, Sanyal AJ, et al. Farnesoid X nuclear receptor ligand obeticholic acid for non-cirrhotic, non-alcoholic steatohepatitis (FLINT): a multicentre, randomised, placebo-controlled trial. Lancet 2015; 385: 956-965.
Copyright: © 2018 Clinical and Experimental Hepatology. This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) License (http://creativecommons.org/licenses/by-nc-sa/4.0/), allowing third parties to copy and redistribute the material in any medium or format and to remix, transform, and build upon the material, provided the original work is properly cited and states its license.
Quick links
© 2024 Termedia Sp. z o.o.
Developed by Bentus.