ArticlePDF AvailableLiterature Review

Abstract and Figures

Non-alcoholic fatty liver disease (NAFLD), or metabolic (dysfunction)-associated fatty liver disease (MAFLD), is characterized by high global incidence and prevalence, a tight association with common metabolic comorbidities, and a substantial risk of progression and associated mortality. Despite the increasingly high medical and socioeconomic burden of NAFLD, the lack of approved pharmacotherapy regimens remains an unsolved issue. In this paper, we aimed to provide an update on the rapidly changing therapeutic landscape and highlight the major novel approaches to the treatment of this disease. In addition to describing the biomolecules and pathways identified as upcoming pharmacological targets for NAFLD, we reviewed the current status of drug discovery and development pipeline with a special focus on recent evidence from clinical trials. Prikhodko, V.A.; Bezborodkina, N.N.; Okovityi, S.V. Pharmacotherapy for Non-Alcoholic Fatty Liver Disease: Emerging Targets and Drug Candidates. Biomedicines 2022, 10, 274. https://doi.org/10.3390/biomedicines10020274.
Content may be subject to copyright.


Citation: Prikhodko, V.A.;
Bezborodkina, N.N.; Okovityi, S.V.
Pharmacotherapy for Non-Alcoholic
Fatty Liver Disease: Emerging
Targets and Drug Candidates.
Biomedicines 2022,10, 274.
https://doi.org/10.3390/
biomedicines10020274
Academic Editor: Henricus
A.M. Mutsaers
Received: 16 December 2021
Accepted: 24 January 2022
Published: 26 January 2022
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2022 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
biomedicines
Review
Pharmacotherapy for Non-Alcoholic Fatty Liver Disease:
Emerging Targets and Drug Candidates
Veronika A. Prikhodko 1,* , Natalia N. Bezborodkina 2and Sergey V. Okovityi 1,3
1Department of Pharmacology and Clinical Pharmacology, Saint Petersburg State Chemical and
Pharmaceutical University, 14A Prof. Popov Str., 197022 St. Petersburg, Russia;
sergey.okovity@pharminnotech.com
2Zoological Institute, Russian Academy of Sciences, 1 Universitetskaya emb., 199034 St. Petersburg, Russia;
natalia.bezborodkina@zin.ru
3Scientific, Clinical and Educational Center of Gastroenterology and Hepatology, Saint Petersburg State
University, 7/9 Universitetskaya emb., 199034 St. Petersburg, Russia
*Correspondence: veronika.prihodko@pharminnotech.com
Abstract:
Non-alcoholic fatty liver disease (NAFLD), or metabolic (dysfunction)-associated fatty liver
disease (MAFLD), is characterized by high global incidence and prevalence, a tight association with
common metabolic comorbidities, and a substantial risk of progression and associated mortality.
Despite the increasingly high medical and socioeconomic burden of NAFLD, the lack of approved
pharmacotherapy regimens remains an unsolved issue. In this paper, we aimed to provide an update
on the rapidly changing therapeutic landscape and highlight the major novel approaches to the
treatment of this disease. In addition to describing the biomolecules and pathways identified as
upcoming pharmacological targets for NAFLD, we reviewed the current status of drug discovery
and development pipeline with a special focus on recent evidence from clinical trials.
Keywords:
non-alcoholic fatty liver disease; non-alcoholic steatohepatitis; chronic liver disease;
hepatoprotection; metabolic disorders
1. Introduction
Non-alcoholic fatty liver disease (NAFLD) includes a range of chronic conditions
characterized by excessive hepatic lipid accumulation, defined by the presence of steatosis
in >5% of hepatocytes, in the absence of significant alcohol consumption or other causes
of liver injury [
1
]. The global prevalence of NAFLD is currently estimated at 25% [
2
],
and is projected to increase by 21% by 2030 [
3
]. About 20–25% of NAFLD cases are
classified as non-alcoholic steatohepatitis (NASH), which has a substantially higher risk
of progression to liver fibrosis, cirrhosis, and end-stage liver disease, and hepatocellular
carcinoma (HCC) [
3
]. Considering the highly heterogeneous pathogenesis of metabolic
liver diseases, the term ‘metabolic (dysfunction)-associated fatty liver disease’ (MAFLD)
has recently been proposed as a broader alternative to the conventional ‘NAFLD’ [4].
Due to the increasingly high medical and socioeconomic burden of this disease as
well as its strong association with obesity, metabolic and cardiovascular disorders, NAFLD
pharmacotherapy has been the focus of latest research [
3
,
5
]. However, to date, no drug has
been approved by neither the European Medicines Agency nor the United States Food and
Drug Administration (FDA) for NAFLD [2,6].
In the present paper, we aimed to review the most recent data on major pharmaco-
logical targets in this disease, summarize the clinical evidence for novel investigational
agents as well as currently marketed drugs, and highlight the latest advances in the drug
development pipeline for NAFLD.
Biomedicines 2022,10, 274. https://doi.org/10.3390/biomedicines10020274 https://www.mdpi.com/journal/biomedicines
Biomedicines 2022,10, 274 2 of 34
2. THRβAgonists
Thyroid hormone receptor beta (THR
β
) is a nuclear receptor and a transcription factor
that mediates the genomic effects of thyroid hormones. THR
β
1 is the predominant isoform
expressed in the human liver, while other THRs are also found in the heart, brain, kidney,
skeletal muscle, and other tissues. Upon activation by triiodothyronine or other agonist,
THR
β
displaces corepressor proteins from the deoxyribonucleic acid (DNA) and facilitates
coactivator binding, thus allowing for gene transcription. THR
β
can form heterodimers
with other THR or nuclear receptor superfamily proteins, such as liver X receptor or
peroxisome proliferator-activated receptors (PPAR) [7].
Hepatic THR
β
upregulate free fatty acid (FA) uptake and oxidation, lipophagy and
lipolysis, and promote mitochondrial biogenesis and respiration, leading to increased
adenosine triphosphate (ATP) consumption and energy expenditure. THR
β
have also been
shown to induce transcriptional activation of bile acid (BA) synthesis, biliary lipid secretion,
and cholesterol serum clearance, leading to a decrease in proatherogenic lipoprotein lev-
els [
8
,
9
]. Additionally, THR
β
stimulates normal hepatocyte proliferation, at the same time
downregulating nuclear signaling pathways and inhibiting tumour growth and metastasis
formation in HCC [10].
Selective THR
β
agonists that are currently being developed for the treatment of
NAFLD include resmetirom (MGL-3196), VK2809 (MB 07811), TERN501, ASC41, and
MGL-3745
. In a 36-week phase 2 randomized clinical trial (RCT) with an additional
36-week
open label extension, resmetirom was safe, well tolerated, and significantly im-
proved lipid profiles, liver steatosis (as indicated by magnetic resonance imaging/proton
density fat fraction (MRI-PDFF)), liver stiffness (assessed by transient elastography), and
N-terminal type III collagen pro-peptide (Pro-C3) levels in patients with biopsy-confirmed
non-alcoholic steatohepatitis (NASH) [
11
], supporting its further evaluation in three ongo-
ing phase 3 trials (NCT04197479, NCT04951219, NCT03900429). The lipid-lowering effect
of resmetirom was accompanied by a significant improvement in alanine aminotransferase
(ALT) and
γ
-glutamyl transpeptidase (GGT) levels as well as a reduction in NAFLD activity
(NAS) and enhanced liver fibrosis (ELF) scores [11].
VK2809 (MB07811) is a prodrug yielding an active metabolite via oxidation by hepatic
CYP3A4 enzyme. A 12-week phase 2a RCT found VK2809 to possess significant antisteatotic
activity in NAFLD patients [
12
]. ASC41, also a prodrug, improved serum lipids and liver
histology in rats [
13
], and subsequently demonstrated a good safety profile and substantial
hypolipidemic activity in healthy volunteers [
14
], advancing to phase 2 (NCT05118360).
A newer THRβagonist, TERN-501, reduced serum cholesterol levels and attenuated liver
steatosis and fibrosis in rodent models of hyperlipidaemia and NASH [
15
], and has been
recently approved for phase 1 clinical trials [
16
,
17
]. MGL-3745 is being evaluated in
preclinical studies, and no data regarding its therapeutic activity are available yet. Two
fixed-dose combinations (FDC), ASC43 and ASC45, containing ASC41 and either a FA
synthase (FASN) inhibitor (ASC40) or a farnesoid X receptor (FXR) agonist (ASC42), are
being evaluated in preclinical and phase 1 clinical trials, respectively [16].
3. Lipogenesis Inhibitors
Acetyl-CoA carboxylase (ACC) is a multi-subunit enzymatic complex that catalyzes
the irreversible carboxylation of acetyl-coenzyme A (CoA) to produce malonyl-CoA. Acetyl-
and malonyl-CoA are then utilized by FASN to synthesize palmitate, which is converted
to stearate by the elongation of long-chain fatty acids family member 6. Stearoyl-CoA
desaturase 1 (SCD1) converts stearoyl-CoA and palmitoyl-CoA into unsaturated fatty
acids, which are later esterified with glycerol by several transferases, including diglyceride
acyltransferase (DGAT), to produce triacylglycerides (TAG). Two DGAT isozymes have
been identified so far: DGAT1 and DGAT2, which appear to be primarily responsible for
esterifying exogenous and endogenous fatty acids, respectively [
18
]. ACC, FASN, SCD1,
and DGAT catalyze the rate-limiting steps in the de novo lipogenesis, and have therefore
been considered suitable targets for the treatment of NAFLD.
Biomedicines 2022,10, 274 3 of 34
However, some animal evidence suggests that inhibition of TAG synthesis may worsen
hepatic inflammation and fibrosis [
19
], and alter intestinal barrier function, leading to diar-
rhoea and steatorrhoea [
20
]. These findings might represent a challenge for the development
of lipogenesis inhibitors and potentially limit their clinical application to early stages of
liver steatosis not associated with inflammation and fibrogenesis [21].
3.1. ACC Inhibitors
Firsocostat (GS-0976) and clesacostat (PF-05221304) are selective ACC inhibitors cur-
rently under development for combination therapy of NASH. Since ACC is indirectly
inhibited by FXR [
22
], firsocostat has been combined with the FXR agonist (see 4.1 FXR
agonists) cilofexor in order to maximize the resulting effects. In the phase 2b ATLAS trial,
20 mg/d firsocostat + 30 mg/d cilofexor provided significant reductions in NAS scores,
liver steatosis, lobular inflammation and hepatocellular ballooning (HCB), and improved
liver biochemistry over 48 weeks in patients with septal fibrosis due to NASH [
23
]. Cle-
sacostat (2–50 mg/d) has demonstrated good efficacy in terms of reducing liver steatosis,
and is being developed in combination with the DGAT2 inhibitor ervogastat to address the
frequently observed elevation of serum TAG, a known effect of ACC inhibitors [
24
]. Two
phase 2 studies are ongoing to determine the optimal doses of both agents for NASH with
and without liver fibrosis (NCT04399538, NCT04321031).
3.2. FASN Inhibitors
Known investigational FASN inhibitors include ASC40 (TVB-2640), FT-8225, and
ASC44F, a fixed-dose combination containing ASC40 and ASC42 (a FXR agonist; see
FXR agonists). ASC40 (50–150 mg/d) provided near complete (~90%) inhibition of de
novo lipogenesis and reduced liver steatosis in obese subjects with substantial NAFLD
risk [
25
]. In the phase 2 FASCINATE-1 trial, ASC40 at lower doses (25 or 50 mg/d)
improved ALT and low-density lipoprotein (LDL) cholesterol levels, attenuated liver
steatosis, fibrosis, lipotoxicity, dyslipidemia, and hepatic insulin resistance in obese NASH
patients [
26
]. A subsequent phase 2 study, FASCINATE-2, has been initiated (NCT04906421).
The anti-inflammatory and antifibrotic properties of ASC40 have been linked to inhibited
proinflammatory cytokine production, T-cell differentiation, and repression of collagen
synthesis [
26
]. FT-8225 and ASC44F have entered preclinical development [
16
,
27
]; no data
regarding their efficacy is available yet.
3.3. SCD1 Inhibitors
Aramchol (C20-FABAC) is a synthetic conjugate of arachidic and cholic acids and the
first-in-class inhibitor of SCD1. In addition to inhibiting de novo lipogenesis and hepatic
stellate cell (HSC)-mediated fibrogenesis [
28
], aramchol is capable of reducing serum choles-
terol levels and promoting gallstone dissolution via stimulation of macrophage cholesterol
efflux and solubilization of cholesterol [
29
]. In the recently completed
52-week
phase 2b
ARREST trial, aramchol (600 mg/d) did not meet the primary endpoint (a significant
decrease in hepatic TAG), but was well tolerated and improved liver histology and ALT
levels [
30
]. The phase 3 ARMOR RCT has been designed to evaluate the safety and effi-
cacy of 300 mg/d aramchol in patients with biopsy-confirmed liver fibrosis due to NASH
(NCT04104321).
3.4. DGAT Inhibitors
Ervogastat (PF-06865571), a selective DGAT2 inhibitor, significantly reduced liver fat
fraction in patients with mild NAFLD [
31
]. Two phase 2 RCT are underway to assess
the safety and efficacy of ervogastat alone and in combination with clesacostat in NASH
patients with and without liver fibrosis (NCT04399538, NCT04321031). ION224 (IONIS-
DGAT2
Rx
) is a ligand-conjugated chimeric antisense oligonucleotide designed to suppress
the biosynthesis of DGAT2. ION224 reduced total liver fat content and several fibrosis
biomarker levels in a small-scale trial in patients with type 2 diabetes mellitus (T2DM) and
Biomedicines 2022,10, 274 4 of 34
NAFLD [
32
], and is planned to be evaluated in a longer phase 2 RCT in nondiabetic subjects
(NCT04932512). Newest DGAT2 inhibitors include PF-07202954, and DGAT1 inhibitors,
VK1430, SNP-610, and SNP-630; the latter two molecules are reported to have additional
CYP2E1-inhibiting properties [33].
3.5. ω-3 PUFAs
ω
-3 Polyunsaturated fatty acids (
ω
-3 PUFA) are long-chain FA characterized by the
presence of a double bond three atoms away from the terminal methyl group in their
chemical structure. The three most common biologically active
ω
-3 PUFA include
α
-
linolenic (ALA) acid and its metabolites eicosapentaenoic (EPA) and docosahexaenoic
(DHA) acids. ω-3 PUFAs cause transcriptional repression of the key enzymes involved in
hepatic glycolysis and de novo lipogenesis, such as ACC, FASN, and L-pyruvate kinase [
34
].
Increased
ω
-3 PUFA intake results in their increased incorporation into cell mem-
brane phospholipids, corresponding to a positive shift in the
ω
-3/
ω
-6 ratio. This leads
to decreased availability of the
ω
-6 arachidonic acid (AA), and the subsequent substrate-
dependent inhibition of eicosanoid inflammatory mediator production by the leukocytes.
Moreover, EPA and DHA have been found to suppress leukocyte chemotaxis and adhesion
molecule expression, altogether providing a multimodal anti-inflammatory effect [
35
]. In
addition,
ω
-3 PUFA are also known for their prominent antioxidant, regenerative, and
antitumour properties [36].
A 6-week treatment with
ω
-3 PUFA (64% ALA + 21% EPA + 16% DHA) was associ-
ated with significant improvement of hepatic proteomic and plasma lipidomic markers of
lipogenesis, lipotoxicity, oxidative stress, and mitochondrial respiration in patients with
biopsy-confirmed NASH [
37
]. The increase in plasma ALA and DHA levels, and the subse-
quent decrease in AA levels correlated with the percentage of patients with improvements
in NAS scores, lobular inflammation, and HCB. However, the overall liver histology and
body weight were not significantly altered by ω-3 PUFA treatment [38].
A systematic review and a meta-analysis confirmed that DHA and EPA attenuate liver
steatosis with no significant weight loss in adults [
39
], and a small-scale RCT reported
the same effects for dietary DHA supplementation in children [
40
]. More recently, three
meta-analyses, including up to 22 RCT and more than 1300 patients, found
ω
-3 PUFA to
significantly decrease ALT, aspartate aminotransferase (AST), GGT levels, liver fat content
and insulin resistance, having no significant effect on body weight in NAFLD [4143].
Epeleuton (15-hydroxyeicosapentaenoic acid ethyl ester, DS102) is a second-generation
synthetic EPA derivative. In a 16-week phase 2a RCT in obese subjects with NAFLD,
epeleuton (2 g/d) improved circulating inflammatory markers, lipid profiles, and insulin
resistance, but failed to reach either of the primary endpoints including reductions in ALT
levels and liver stiffness [
44
]. Since a slight dose-dependent attenuation of liver steatosis
was observed, further trials of longer duration are planned [45].
Icosabutate (NST-4016) is a structurally engineered
ω
-3 PUFA ether characterized to
increased liver exposure due to direct absorption into the portal vein [
46
]. The 62-week
phase 2b ICONA trial is underway to evaluate the efficacy of icosabutate (300 or 600 mg/d)
in patients with biopsy-confirmed NASH (NCT04052516). Interim analysis data indicated
significant dose-dependent decreases among both dosage groups in ALT, AST, GGT, and
alkaline phosphatase (ALP) levels, while patients dosed with 600 mg/d icosabutate also
had improvements in non-invasive fibrosis and inflammatory biomarker profiles [46].
4. Bile Acid Metabolism Modulators
4.1. FXR Agonists
Farnesoid X receptor (FXR), also known as bile acid receptor, is a nuclear receptor
expressed at high levels in the liver and ileum. FXR acts as a BA sensor and governs a
major negative feedback loop in the BA, glucose, cholesterol, and TAG metabolism [
22
].
Many of the effects of FXR are directly mediated by fibroblast growth factors (FGF) 19 and
21, downstream messengers whose functions are briefly described in Fibroblast growth
Biomedicines 2022,10, 274 5 of 34
factor analogues. Upon increased postprandial release of BA into the intestine, activated
FXR induces the transcriptional repression of the rate-limiting enzyme cholesterol 7
α
-
monooxygenase (CYP7A1) and several transporters involved in BA biosynthesis and liver
uptake. CYP7A1 inhibition by FXR enhances the excretion of excessive cholesterol via the
canalicular transporters into bile as well as directly into the intestinal lumen. FXR also
upregulates the expression of bile salt export pump, multidrug resistance protein-3, and
organic solute transporter
α
/
β
, which facilitate BA efflux from hepatocytes and maintain
the enterohepatic circulation of bile [
47
]. FXR activation also favors bile acid conjugation
and detoxification, and stimulates biliary phospholipid excretion.
FXR and its downstream targets repress sterol regulatory element-binding protein 1
SREBP-1c, the major transcription factor for lipogenic pathways, thereby inhibiting FASN,
ACC, and SCD1. In addition, prandial glucose can increase the intestinal FXR activity via
post-translational modification, shifting the equilibrium towards glycogen deposition and
reducing blood glucose levels [22]. The multifaceted nature of FXR has made it one of the
most attractive novel targets for NAFLD therapy. Current FXR-activating drug candidates
include the sterol derivatives obeticholic acid (OCA), EDP-305, INT-767, and INT-787, and
the non-steroidal compounds MET409, tropifexor, cilofexor, vonafexor, TERN-101, ASC42,
EDP-297, HPG1860, and HPG7233.
OCA (Ocaliva
®
), the first-in-class FXR agonist, is approved by the FDA for non-
cirrhotic primary biliary cholangitis (PBC) and is nearing approval for liver fibrosis due to
NASH [
48
]. In phase 2 studies, OCA increased insulin sensitivity and reduced markers of
liver damage in patients with T2DM and NAFLD [
49
], and reduced ALT levels, improved
NAS scores, prevented and partially reversed fibrogenesis in non-diabetic, pre-cirrhotic
NASH patients [
50
]. Currently, the efficacy and safety of OCA are being evaluated in two
phase 3 trials, the REGENERATE trial in NASH/fibrosis patients (NCT02548351), and the
REVERSE trial in patients with compensated cirrhosis due to NASH (NCT03439254).
OCA and especially its taurine and glycine conjugates are known to actively bind
Takeda G-protein receptor 5 (TGR5), a cell membrane G protein-coupled receptor (GPCR)
that has been largely implicated in BA-associated pruritus development [
51
] and gallstone
formation [
52
]. Hence, several newer molecules, e.g., EDP-305 and INT-787, have been
structurally optimized in an attempt to avoid TGR5 engagement and potential safety
concerns [
53
]. However, PBC trials have shown that highly selective FXR agonists retain
their adverse effects, at the same time losing in overall efficacy [
54
,
55
]. EDP-305 was later
repurposed for use in NASH, and, along with its close analogue EDP-297, is now intended
to be reserved for future combination regimens following mixed interim results of the phase
2b ARGON-2 study [
56
]. Since FXR and TGR5 seem to exert additive metabolic effects,
dual FXR/TGR5 agonists with balanced activity towards both targets, such as INT-767 and
BAR502, have been proposed for further development for NAFLD treatment [57,58].
MET409, a structurally optimized fexaramine-derived FXR agonist with greater affinity
towards intestinal rather than hepatic FXR, is characterized by improved efficacy and a dif-
ferentiated adverse effect (pruritus and increased LDL-cholesterol levels) profile compared
to OCA [
59
], and is currently being evaluated in a phase 2a RCT alone and in combination
with empagliflozin (NCT04702490).
Tropifexor (LJN452), cilofexor (GS-9674), and vonafexor (EYP001) represent a group
of non-steroidal small molecules with structures different from both OCA and MET409,
resulting in a differential pattern of FXR-related gene expression [
60
] and possibly improved
anticholestatic activity. The parent compound for this group, turofexorate (WAY-362450),
completed a phase 1 study, but its development was discontinued thereafter [
61
]. Tropifexor
(200, but not 140 mcg/d) reduced ALT, GGT levels, body weight, and liver fat content,
and attenuated liver fibrosis in patients with biopsy-confirmed NASH in the 48-week
phase 2 FLIGHT-FXR RCT [
62
,
63
]. Two additional phase 2 trials are currently underway to
investigate the efficacy of tropifexor combinations with licogliflozin (NCT04065841) and
LYS006 (NCT04147195) in NASH/fibrosis and NAFLD/NASH, respectively.
Biomedicines 2022,10, 274 6 of 34
Cilofexor reduced liver fat content (100 mg/d) and GGT, N-terminal type IV collagen
pro-peptide, and serum primary BA levels (30 or 100 mg/d) over 24 weeks, but did not
affect liver elasticity and ELF scores in NASH patients [
64
]. A phase 2 RCT has been initiated
to evaluate the efficacy of a fixed-dosed combination of cilofexor and firsocostat, alone or
in combination with semaglutide, for compensated cirrhosis due to NASH (NCT04971785).
Vonafexor (100 mg/d for 12 weeks), a 2nd generation, highly selective non-steroidal FXR
agonist, has recently been found to improve liver steatosis, the fibro-inflammation marker
cT1 levels, and estimated glomerular filtration rate (eGFR) in NASH patients with normal
or mildly decreased eGFR [
65
]. Nidufexor (LMB763), a non-steroidal partial FXR agonist,
was being developed for NASH, liver fibrosis, and cholestatic liver disease, but appears
to have been discontinued despite seemingly promising results from a phase 2 study in
NASH subjects [66].
TERN-101 demonstrated a favorable safety profile and improved ALT levels and liver
steatosis in the phase 2a LIFT RCT in patients with pre-cirrhotic NASH [
67
]. ASC42
treatment was associated with biochemical and histological improvements in animal
NASH models [
16
], while no data are available yet for novel compounds HPG1860 and
HPG7233 [68].
4.2. FGF Analogues
Fibroblast growth factors (FGF), including FGF19 and FGF21, are a family of hormone-
like peptides with broad metabolic, transcriptional, and mitogenic activity. FGF19 is
released by the ileal enterocytes into the enterohepatic circulation in response to postpran-
dial FXR activation by BA. FGF21 is highly expressed in the liver and is released in response
to PPAR
α
, carbohydrate-response element-binding protein (ChREBP), and general control
nonderepressible 2 kinase activation in the presence high serum glucose, high FFA and low
amino acid levels [69]. FGF19 and FGF21 exert their physiological effects via activation of
transmembrane complexes of the enzyme
β
-klotho (KLB) and its FGF co-receptors (FGFR)
FGFR1c/2c/3c/4. FGF21 has the highest affinity for the FGFR1c/KLB complex, which is
expressed in the adipose tissue and central nervous system, while FGF19 primarily targets
FGFR4/KLB, which is found in hepatocytes [69].
Thus, FGF19 and FGF21 act as major downstream messengers in the FXR and PPAR
α
signaling, and control the negative feedback loops to inhibit BA synthesis and lipolysis,
respectively [
69
]. FGF19 represses CYP7A1, controls postprandial BA release into the in-
testinal lumen, inhibits hepatic gluconeogenesis, promotes glycogen deposition, and plays
an important role in the regulation of hepatocyte proliferation and tumorigenesis. FGF21
lowers the preference for glucose intake, modulates energy expenditure, and promotes
glucose and lipid uptake in the adipose tissue, which prevents ectopic lipid accumulation
in liver and skeletal muscle [69].
The investigational FGF19 analogue aldafermin (NGM-282) (0.3–3 mg/d) failed to
meet the primary endpoint of the 24-week phase 2b ALPINE 2/3 trial in NASH patients
with stage 2 or 3 liver fibrosis, defined as a
1 NASH Clinical Research Network stage
improvement in fibrosis with no worsening of NASH. However, the drug was well tolerated
and was significantly superior to placebo in terms of NASH resolution, reduction of liver
steatosis and non-invasive markers of liver injury [
70
]. The ALPINE 4 trial is ongoing to
determine whether aldafermin could improve liver fibrosis and/or NASH in subjects with
compensated cirrhosis (NCT04210245).
Current FGF21 analogues include efruxifermin (AKR-001), BIO89-100, NN9500, BFKB8488A,
MK-3655, and GLP-1-Fc-FGF21 D1. Efruxifermin is a fusion protein of human immunoglob-
ulin G1 (IgG1) Fc domain linked to a modified human FGF21 with balanced affinity
towards FGFR1c/2c/3c. Efruxifermin (28–70 mg/d) improved lipoprotein profiles and
glycaemic control in T2DM patients, significantly attenuated liver steatosis in the 16-week
phase 2a BALANCED study [71], and is now being evaluated in three more phase 2 RCTs
(NCT05039450, NCT04767529, NCT03976401). BIO89-100 is a specifically engineered gly-
colpolyethylene glycol (PEG)-ylated FGF21 analogue that led to clinically meaningful
Biomedicines 2022,10, 274 7 of 34
reductions in liver fat content, markers of inflammation and fibrosis in a phase 1b/2a
proof-of-concept study in NASH patients [
72
]. Recently, the phase 2b ENLIVEN trial of
BIO89-100 for stage 2/3 fibrosis due to NASH has been initiated (NCT04929483), while two
phase 1 open-label studies are still underway (NCT05022693, NCT04048135). No preclinical
data is available yet for the investigational compound NN9500 [73].
BFKB8488A and MK-3655 are humanized bispecific anti-FGFR1c/KLB agonist mon-
oclonal antibodies (mAB). BFKB8488A was safe and adequately tolerated, reduced liver
steatosis in a dose-dependent fashion, and improved markers of cardiometabolic and liver
health in obese T2DM patients with NAFLD [
74
]. The phase 2b BANFF trial has been
initiated to explore the efficacy and safety of this compound in non-alcoholic liver fibrosis
(NCT04171765). MK-3655 (NMG313), an insulin-sensitizing anti-FGFR1c/KLB agonist
mAB, was effective against liver steatosis in obese and insulin-resistant NAFLD patients,
and has proceeded into a phase 2b study in pre-cirrhotic NASH with or without T2DM
(NCT04583423).
GLP-1-Fc-FGF21 D1 is a novel fusion protein incorporating a KLB-binding FGF21
variant and a glucagon-like peptide 1 receptor (GLP1R) agonist. In murine models of
T2DM and obesity, GLP-1-Fc-FGF21 D1 improved liver function, serum and hepatic lipid
profiles, and reduced body weight and NAS scores with an efficacy superior to either FGF21
or GLP1R agonists alone [
75
]. The PEGylated human recombinant FGF21 pegbelfermin
(ARX-618) has recently demonstrated suboptimal efficacy in compensated cirrhosis due
to NASH in the FALCON 2 trial, resulting in pending discontinuation and/or possible
repurposing for other indications [76].
5. Fibrogenesis Inhibitors
5.1. Galectin Antagonists
Galectins, formerly known as S-type lectins, are a family of carbohydrate-binding
proteins that are selective towards β-galactoside-containing glycans. To date, 15 subtypes
of galectins family have been identified in humans, of which galectin-1, -3, and -9 are
the most implicated in liver disease. Numerous experimental and clinical study results
suggest that galectins play important and diverse roles in fibrogenesis, cellular immunity
and inflammatory response, cell cycle regulation, apoptosis, regeneration, and tumorigene-
sis [77,78].
Galectin-1 promotes the proliferation, migration and activation of HSC, and the subse-
quent fibrogenesis via stimulation of transforming growth factor
β
(TGF
β
)/platelet-derived
growth factor signaling and disruption of cell adhesion. In HCC, galectin-1 promotes the
epithelial–mesenchymal transition, cell adhesion, metastasis, and immunosuppression.
However, elevated galectin-1 expression has also been found beneficial for liver regenera-
tion, liver allograft survival, and recovery after hepatic ischemia-reperfusion injury [77].
Galectin-3 has been identified as a pivotal regulator in the progression of hepatitis,
hepatic fibrosis, cirrhosis, and HCC. It was found to promote autocrine and paracrine HSC
activation and phagocytosis, mediate the TGF
β
-dependent fibrogenesis, and upregulate
the expression of certain profibrogenic cytokines such as interleukin (IL) 33. The exact role
of galectin-3 for liver cirrhosis is not as clear, but its increased expression has been linked
to accelerated cirrhosis development and deterioration of liver function [
77
]. However,
galectin-3 might be protective against adipose tissue inflammation, diabetes, and atheroscle-
rosis progression, most probably due to and its ability to scavenge the proinflammatory
and proapoptotic advanced glycation end products, the increased activation of the NF-
κ
B
signaling pathway, and the downregulation of NLRP3 inflammasome and IL1
β
expression
in immune cells [77].
A systematic review and meta-analysis by An et al. found that serum
galectin-3/9 levels
correlate with the risk of liver failure and cirrhosis, and high galectin-1/3 expression is associ-
ated with poorprognosis in HCC. However, the evidence for galectin involvement in chronic
liver disease remains controversial, and the impact of galectin-3 levels on NAFLD/NASH is
thought to be dependent upon the stage and severity of liver damage [
78
]. Hence, modern
Biomedicines 2022,10, 274 8 of 34
galectin-targeting drug candidates are intended for use in advanced NASH complicated by
liver fibrosis and/or cirrhosis.
GM-CT-01 and belapectin (GR-MD-02) are semi-synthetic polysaccharides (galac-
tomannan and galactoarabino-rhamnogalacturonan, respectively) having high affinity
towards extracellular galectin-3 and, to a lesser extent, galectin-1. Both compounds pro-
moted resolution of portal inflammation, HCB, liver fibrosis, and cirrhosis, and reduced
portal pressure in a toxin-induced liver injury model in rats [
79
]; however, only belapectin
seems to have advanced further into clinical development. In the 52-week phase 2b NASH-
CX study in subjects with NASH, liver cirrhosis, and portal hypertension, belapectin
(2 mg/kg/2 weeks) did not affect fibrosis or NAFLD activity, but reduced hepatic ve-
nous pressure gradient values and prevented the development of esophageal varices in a
sub-group of patients [
80
]. The phase 2/3 NAVIGATE trial has been initiated to further
evaluate belapectin in patients with liver cirrhosis due to NASH and clinical signs of portal
hypertension but without esophageal varices at baseline (NCT04365868).
GB1211, an oral small-molecule selective galectin-3 antagonist, was effective in several
preclinical fibrosis models, and well tolerated in human volunteers. A phase 1/2a trial
in NASH/fibrosis patients was approved but subsequently placed on hold due to an
undisclosed change in the clinical development strategy for this compound (NCT04607655).
5.2. TLR4 Antagonists
Toll-like receptor 4 (TLR4) belongs to the family of pattern recognition receptors that
activate the innate immune system by recognizing their major ligand, lipopolysaccharide
(LPS). In the liver, TLR4 are expressed in both parenchymal and non-parenchymal type
cells. TLR4 activation results in the activation of NF-
κ
B, mitogen-activated protein kinase
and interferon regulatory factor-mediated pathways, and the subsequent inflammatory
cytokine and interferon production [81].
TLR4 stimulate adhesion molecule expression and chemokine secretion by the HSC,
which induces Kupffer cell migration and the recruitment of extrahepatic monocytes into
the liver. TLR4 also downregulate HSC expression of the Bambi protein, an endogenous
TGF
β
receptor inhibitor, thus promoting profibrogenic TGF
β
signaling. Additionally, TLR4
can inhibit miR-29 expression and increase fibronectin production in the HSC, further
enhancing HSC activation and migration [82].
JKB-122 treatment (5 or 35 mg/d for 12 weeks) was well tolerated and significantly
improved ALT and AST levels, liver steatosis, and serum lipid profiles in a phase 2 study
in patients with NAFLD [
83
]. Another phase 2 RCT is planned to investigate whether
JKB-122 could ameliorate liver fibrosis due to NASH (NCT04255069). Recently, eritoran, a
synthetic bacterial lipid analogue, was found to significantly reduce ALT levels, lobular
inflammation, intrahepatic neutrophil infiltration, and liver fibrosis, but not liver steatosis,
in murine models of acute and chronic liver injury [81].
5.3. LOXL2 Inhibitors
Lysyl oxidase-like protein (LOXL) 2 is a histone modifier amine oxidase that has been
identified as the primary enzyme facilitating covalent crosslinking of collagen and elastin
fibers, thereby promoting collagen network formation and progression of liver fibrosis.
LOXL2 inhibition following the onset of fibrosis has been demonstrated to augment and
accelerate collagen degradation in rodent models [
84
]. Only a few studies have concerned
the role of LOXL3, a less common isozyme of LOXL, in fibrotic diseases, and its value as a
therapeutic target is somewhat controversial [85].
Simtuzumab (GS-6624), a humanized antagonist mAB, was one of the first LOXL2-
targeting drug candidates, that was discontinued after showing a lack of efficacy regarding
liver fibrosis and/or portal hypertension in two phase 2b trials in patients with bridging
fibrosis or compensated cirrhosis associated with NASH [86].
Novel LOXL2 antagonists are represented by orally available small molecules. PXS-
5153A, a dual LOXL2/LOXL3 inhibitor, reduced disease severity and improved liver
Biomedicines 2022,10, 274 9 of 34
function by diminishing collagen content and collagen crosslinks in two rodent models
of liver fibrosis [
87
]. Its cognate PXS-5382A, selective towards LOXL2, showed a satis-
factory pharmacokinetic profile in healthy volunteers (NCT04183517), and is expected to
advance into phase 2 trials. Another LOXL2 inhibitor, GB2064, was initially researched
for myelofibrosis, but appears to have had its potential indication list expanded to include
other fibrotic diseases [88].
5.4. ATX Inhibitors
Autotaxin (ATX) is a glycoprotein enzyme that converts membrane-derived lysophos-
pholipids into lysophosphatidic acid (LPA). LPA, in turn, acts as a multimodal signaling
molecule and causes cytoskeleton remodelling, alters cell proliferation and migration, and
promotes fibrogenesis as well as inflammatory reactions [
89
]. Non-competitive inhibition
of autotaxin/LPA signaling by the indole derivative PAT-505 resulted in a significant at-
tenuation of liver fibrosis in mouse models of NASH [
90
]. In addition, ATX induction has
been implicated in BA-mediated pruritus development, suggesting potentially favorable
safety profiles of ATX inhibitors. Two small-molecule ATX inhibitors, TJC0265 and TJC0316,
have been identified as lead compounds and are currently undergoing optimization and
preliminary in vivo testing [91].
6. Glucose Metabolism Modulators
6.1. PPAR Agonists
PPAR are a family of nuclear receptors that function as transcription factors and
play a regulatory role in glucose homeostasis, lipid metabolism, inflammatory response,
cell development and differentiation [
92
]. Up to now, three PPAR subtypes have been
identified: (1) PPAR
α
, expressed in the liver, adipose tissue, skeletal muscle, heart, and
kidney; (2) PPAR
γ
, expressed in the adipose tissue, colon, macrophages, pancreas, skeletal
muscle, etc.; and (3) PPAR
δ
(
β
), found ubiquitously. Upon ligand binding, PPAR induce
the release of corepressors and recruitment of coactivators, allowing for gene transcription.
Moreover, PPAR can regulate the mitogen-activated protein kinase pathways, and inhibit
inflammatory reactions via transrepression of several proinflammatory transcription factors
such as NF-κB [92].
Hepatic PPAR
α
stimulate mitochondrial FA uptake,
β
-oxidation, ATP production,
and ketogenesis. They upregulate glucose-sensing transcription factors ChREBP and sterol
regulatory element-binding transcription factor 1, and promote FGF21 expression, which
increases tissue insulin sensitivity and maintains glucose homeostasis. Recent experimental
studies strongly suggest that decreased hepatic PPAR
α
expression positively correlates
with insulin resistance and NASH severity, while NASH resolution is associated with an
upregulation of PPAR
α
as well as its target genes [
92
]. However, the evidence for the use
of selective PPARαagonists in NAFLD is limited.
Fenofibrate (200 mg/d) induced complete resolution of biochemical and ultrasono-
graphic evidence of NAFLD in almost half of the patients in an open-label RCT [
93
], but
had minimal efficacy regarding liver histology [
94
]. Gemfibrozil (600 mg/d for 4 weeks)
reduced liver enzyme levels but also did not produce any meaningful changes in liver
morphology [
95
], while clofibrate (2 g/d) failed to show any beneficial effect in NASH
patients [
96
]. Pemafibrate (0.2 mg/d) improved biochemical markers of liver damage and
steatohepatitis according to non-invasive measures in NAFLD subjects [
97
], and bezafibrate
reduced liver steatosis in obese mice with metabolic syndrome [98].
PPAR
γ
are critical positive regulators of adipocyte differentiation and lipogenesis,
insulin sensitivity, and glucose uptake by skeletal muscle. Additionally, PPAR
γ
may
reduce inflammation via inhibition of macrophage activation and tumour necrosis factor
α
production [
92
]. Latest EASL-EASD-EASO [
1
] and AASLD [
99
] clinical practice guidelines
support the use of pioglitazone, a selective PPAR
γ
agonist, in progressive and/or high-
risk, biopsy-proven NASH, due to its efficacy regarding liver histology in NASH patients
with or without T2DM. More recently, lobeglitazone has been demonstrated to reduce
Biomedicines 2022,10, 274 10 of 34
intrahepatic fat content and biochemical markers of liver damage in T2DM patients with
NAFLD [
100
]. An experimental study found nifedipine, an L-type calcium channel blocker,
to exert protective action against diet-induced NASH in rats, most probably due PPAR
γ
activation [101].
PPAR
δ
is the less studied PPAR isotype despite its ubiquitous expression. Experimen-
tal studies have reported PPAR
δ
to reduce very low-density lipoprotein cholesterol levels,
inhibit adipocyte growth and lipid uptake, prevent the formation of reactive oxygen species,
and modulate Kupffer cell activation [
92
]. A recent study found a correlation between
severe, but not moderate, hepatic steatosis and decreased hepatic PPAR
δ
expression [
102
].
Seladelpar (MBX-8025) is the only selective PPAR
δ
agonist currently in the pipeline
for the treatment of NAFLD. The interim analysis results of a 52-week phase 2b RCT of
seladelpar (10–50 mg/d) found it to reduce ALT, GGT, and ALP levels, but only minimally
influence liver steatosis at 12 weeks of treatment [
103
]. Further trials have so far focused
more on the anticholestatic properties of seladelpar, of potential value in primary biliary
cholangitis (NCT03301506).
Saroglitazar (Lipaglyn
®
), a dual PPAR
α
/
γ
agonist, is approved in India for use in
T2DM and precirrhotic NASH. In several RCT in NAFLD patients with and without T2DM,
saroglitazar (4 mg/d) was shown to significantly improve liver biochemistry as well as
hepatic steatosis by non-invasive measures over 16–24 weeks of treatment [
104
,
105
]. The
group of experimental triple PPAR
α
/
γ
/
δ
agonists is now represented by lanifibranor
(IVA337) alone, after its predecessor elafibranor (GFT-505) was discontinued due to lack
of efficacy in NAFLD in the phase 3 RESOLVE-IT trial (NCT02704403). Lanifibranor was
well tolerated and reduced liver enzyme levels and markers of fibrosis in patients with
precirrhotic, highly active NASH in the recently completed phase 2b study. The percentage
of patients with meaningful improvements in steatosis, activity, and fibrosis scores was
significantly higher in the lanifibranor-treated arms, indicating possible improvements
in hepatitis activity as well as HCB [
106
]. Two more trials to evaluate the efficacy of
lanifibranor in concomitant NAFLD and T2DM and in advanced fibrosis due to NASH are
ongoing (NCT03459079, NCT04849728).
6.2. MPC Inhibitors
In addition to their primary mechanism of action, pioglitazone and other thiazolidine-
diones are known to interact with the mitochondrial pyruvate carrier (MPC) to suppress
pyruvate transport into the mitochondrial matrix. Since this direct, non-genomic effect is
considered essential for the inhibition of hepatic gluconeogenesis by PPAR
γ
ligands, novel
MPC-inhibiting thiazolidinedione derivatives with minimal affinity towards PPAR
γ
have
been synthesized.
Azemiglitazone (MSDK-0602K) reduced liver enzyme levels in NASH/fibrosis pa-
tients regardless of T2DM presence [
107
], and was characterized by a markedly improved
safety profile [
108
]. However, the drug did not demonstrate significant effects on any of
the histological endpoints [
107
]. A phase 3 trial to evaluate azemiglitazone in diabetic or
prediabetic patients with NAFLD/NASH has been initiated (NCT03970031). An alternative
approach to minimizing the side effects of PPAR
γ
agonists is represented by PXL065, the
deuterium-stabilized R-isomer of pioglitazone [
109
], which is currently being assessed in a
phase 2 RCT (NCT04321343).
6.3. Incretin Mimetics
Incretins are a small family of intestinal L-cell-derived peptide hormones that includes
glucagon-like peptide 1 (GLP1), glucose-dependent insulinotropic polypeptide (GIP), and
oxyntomodulin. The incretin axis mediates the physiological response to hyperglycaemia
and couples glucose intake with pancreatic secretion [
110
]. GLP1 receptors (GLP1R) are
expressed in the
β
-cells, hepatocytes, white adipose tissue, brain, and skeletal muscle [
111
].
GLP1R activation induces insulin secretion, decreases insulin resistance, inhibits glucagon
release and lipogenesis, and suppresses appetite and gastrointestinal motility. Additionally,
Biomedicines 2022,10, 274 11 of 34
hepatic GLP1R stimulate FA
β
-oxidation, inhibit profibrogenic signaling pathways, and
exert a mild anti-inflammatory effect by indirectly reducing CRP, proinflammatory cytokine,
and chemokine production [112].
At the moment, semaglutide is the only GLP1R agonist being developed for the
treatment of NASH in nondiabetic subjects. In the recently completed 72-week phase 2 trial,
semaglutide treatment (0.4 mg/d) led to NASH resolution with no worsening of fibrosis
in 59% vs. 17% in the placebo group, which is considered the highest response rate that
a drug has ever achieved in a NASH trial up to now [
113
,
114
]. Currently, semaglutide is
being evaluated in several phase 2 trials as monotherapy (NCT03884075, NCT04216589)
as well as in combinations with empagliflozin (NCT04639414), and cilofexor/firsocostat
(NCT04971785). A 5-year-long phase 3 study designed to include 1200 patients with
precirrhotic NASH has been initiated (NCT04822181).
Exenatide, lixisenatide, liraglutide, and dulaglutide have all demonstrated signifi-
cant antisteatotic activity, and (with the exception of lixisenatide) improved intrahepatic
cholestasis in T2DM/NAFLD patients. Amelioration of cytolysis markers was reported
for exenatide [
115
,
116
] and dulaglutide [
117
,
118
], and lixisenatide was effective against
liver fibrosis and inflammation [
119
]. Liraglutide reduced hepatitis activity and liver fi-
brosis as well as attenuated HCB in NASH patients regardless of the presence of T2DM,
as determined by the phase 2 LEAN study [
120
]. A recent meta-analysis by Ghosal et al.,
including 8 RCT and over 600 patients, found that GLP1R agonists in general improve liver
function and histology by improving glycaemia, reducing body weight and hepatic fat
content, which in turn might be beneficial for hepatic inflammation in NAFLD concomitant
with T2DM [121].
While both GIP and GLP1 are potent insulin secretagogues, GIP has a more robust,
dose-dependent secretory pattern, and appears to make a greater contribution than GLP1 to
prandial insulin secretion in healthy subjects, while in T2M its activity is depleted. In addi-
tion, GIP, but not GLP1, stimulates glucagon secretion by the
α
-cells at low glycaemia under
physiological conditions, and in a glucose-independent fashion in T2M [
110
]. GIP receptors
(GIPR) upregulate lipogenesis, FA esterification and TAG accumulation in adipocytes, and
inhibits prandial lipid absorption. Supraphysiological levels of GIP have been associated
with increased systemic inflammatory response, and are considered a risk factor for the
development of NASH [122].
Tirzepatide (LY3298176) is a synthetic injectable dual GLP1/GIP peptide agonist
currently researched for NAFLD treatment. In T2DM subjects with NASH, tirzepatide
(
1–15 mg/week
for 26 weeks) effectively reduced ALT, AST, cytokeratin 18, Pro-C3, and
increased adiponectin levels [
123
]. Additionally, tirzepatide treatment led to greater im-
provements in liver fat content compared to titrated insulin degludec in T2DM, according
to the phase 3 SURPASS-3 RCT results [
124
]. A phase 2 study, designated SYNERGY-NASH,
has been initiated to evaluate the efficacy of tirzepatide in nondiabetic subjects with NASH
(NCT04166773).
Oxyntomodulin shares sequence similarity with both and GLP1 and glucagon, and
activates GLP1R and glucagon receptors (GCGR) under physiological conditions. Simulta-
neous GLP1R activation prevents hyperglycaemic response characteristic of glucagon, at
the same time potentiating its catabolic effects and greatly intensifying hepatic glycolysis,
glycogenolysis, and lipolysis [
125
]. Weight reduction, anorexigenic and hypoglycaemic
effects have been linked to GLP1 activation, while GCGR activation is thought to contribute
primarily to hepatic steatosis attenuation and improved mitochondrial respiration. The
clinical utility of oxyntomodulin itself is limited by a short circulatory half-life due to rapid
renal clearance and degradation by dipeptidyl peptidase 4 (DPP4) [126].
In contrast to the native hormone, synthetic oxyntomodulin mimetics are resistant to
proteolytic cleavage and have prolonged pharmacological action. Cotadutide (MEDI0382)
(100–300
µ
g/d) caused substantial improvements in liver enzyme levels and markers
of liver fibrosis in concomitant obesity, T2DM, and NASH in a phase 2b study that in-
cluded over 800 subjects [
127
]. Efinopegdutide (HM12525A, MK-6024), a PEGylated
Biomedicines 2022,10, 274 12 of 34
long-acting peptide agent, has demonstrated promising antihyperlipidaemic, antisteatotic,
and anti-inflammatory activity in mice and hamsters [
128
], and is going to be evaluated in
a phase 2 RCT in NASH with semaglutide as an active comparator (NCT04944992). Other
dual GLP1/GCCR agonists intended for use in NAFLD include pemvidutide (ALT-801)
(NCT05006885), danuglipron (PF-06882961) (in combination with ervogastat) [
129
], BI
456906 (NCT04771273), and HM14320 (a glucagon-containing combination) [
130
]. Finally,
a novel triple GLP1R/GCGR/GIPR agonist, HM15211, induced significant reductions in
liver steatosis, fibrosis, and inflammation in mice [
131
]; a phase 2 clinical trial is ongoing
(NCT04505436).
GLP2, usually not considered an incretin, is prevalent in the gastrointestinal tract,
where it promotes lipid absorption, regulates intestinal motility, mucosal morphology,
function and integrity of the intestine [
132
]. Teduglutide, a selective GLP2R agonist,
reduced liver steatosis and disease activity scores in rats, possibly by restoring normal
intestinal permeability [133].
DPP4 inhibitors represent a group of indirect incretin mimetics as they prevent the pro-
teolytic cleavage of GLP1, GIP, and oxyntomodulin. To the best of our knowledge, no DPP4
inhibitors are yet in the global pipeline for liver disease. However, a number of small-scale
clinical trials have evaluated their potential efficacy in NAFLD in the presence or absence
of concomitant T2DM. Among this group, only sitagliptin (100 mg/d) was found effective
against hepatic steatosis and HCB irrespective of T2DM in a 1-year open-label RCT [
134
].
Vildagliptin (100 mg/d) [
135
], saxagliptin (5 mg/d) [
136
], omarigliptin (25 mg/week) [
137
],
and teneligliptin (20 mg/d) [
138
] improved liver function and some non-invasive markers
of NAFLD, and alogliptin (25 mg/d) was only moderately effective against NASH over
12 months of treatment in T2DM/NAFLD patients [
139
]. A recent meta-analysis by dos
Santos et al. found the existing evidence for DPP4 inhibitors in NAFLD to be of poor
quality and altogether not supportive of their clinical effectiveness [140]. Evogliptin [141],
anagliptin [
142
,
143
], trelagliptin [
144
], gemigliptin [
145
], and linagliptin [
146
] have demon-
strated beneficial effects in experimental rodent models, but their clinical value remains to
be explored in future trials.
6.4. SGLT Inhibitors
Sodium/glucose cotransporter (SGLT) 2 inhibitors are a relatively novel class of oral
antidiabetic agents that increase urinary glucose excretion by inhibiting glucose reabsorp-
tion by SGLT2 in the proximal tubules. Several trials have demonstrated the improvement
of cardiovascular and renal outcomes by treatment with compounds of this class, namely,
empagliflozin, canagliflozin, and dapagliflozin [
147
]. SGLT2 inhibitors are known for their
multiple metabolic effects that are notably relevant to NAFLD pathophysiology, including
the general shift towards increased ketogenesis, gluconeogenesis, glycogenolysis, and FA
β
-oxidation. They inhibit leptin production by adipocytes, leading to decreased food intake,
increase adiponectin levels, provide mild insulin sensitization, suppress HSC activation
and fibrogenesis. Additionally, SGLT2 inhibitors may indirectly suppress sympathetic
innervation and increase the vagal tone, thereby preventing the activation of Kupffer cells
and the associated inflammatory processes [148,149].
Recent evidence mostly supports the efficacy of the majority of SGLT2 inhibitors for
improving liver dysfunction, steatosis and fibrosis in NAFLD concomitant with T2DM.
Among this group, only dapagliflozin, empagliflozin, and canagliflozin treatment was
associated with beneficial effects in nondiabetic NAFLD patients. Dapagliflozin (10 mg)
significantly reduced ALT, AST, and GGT levels, according to a retrospective study [
150
],
while empagliflozin (10 mg/d) also attenuated liver steatosis and liver stiffness, indicative
of potential antifibrotic activity, in a small-scale RCT [
151
]. The phase 3 DEAN trial to eval-
uate dapagliflozin in biopsy-confirmed NASH patients has been initiated (NCT03723252).
Canagliflozin (100 mg/d) improved liver enzyme levels and FIB-4 index values in an
open-label, uncontrolled pilot study [
152
]. Additionally, empagliflozin had a beneficial
effect on cognitive functions and reduced anxiety in an experimental NAFLD model [
153
].
Biomedicines 2022,10, 274 13 of 34
Ipragliflozin (50 mg/d for 72 weeks) ameliorated liver fibrosis and enhanced NASH
resolution [
154
], remogliflozin etabonate (50–1000 mg/d for 12 weeks) reduced FIB-4 and
NAFLD-fibrosis scores [
155
], and ertugliflozin (5 or 15 mg/d for 52 weeks) reduced liver
transaminase levels [
156
] in TD2M patients with different stages of NASH. Several pilot
studies in T2DM/NAFLD subjects confirmed the antisteatotic properties of luseogliflozin
and tofogliflozin [157159].
The SGLT1 subtype plays a relatively smaller (10–20% and up to 40% when SGLT2
are blocked) role in the renal glucose reabsorption, but is more abundant in the small
intestine along with the heart and lungs. Intestinal SGLT1 (iSGLT1) inhibition leads to
substantially reduced glucose and galactose absorption from the intestinal lumen, and
increased incretin (GLP1, peptide YY) release by enteroendocrine cells [
160
]. Currently,
licogliflozin (LIK066) is the only SGLT1/2 inhibitor being evaluated in NASH independent
of T2DM presence, alone and in combination with the FXR agonist tropifexor, in the ongoing
phase 2 ELIVATE study (NCT04065841). Previously, licogliflozin (150 mg/d) reduced
ALT, ALT, GGT levels and liver fat content over 12 weeks compared to placebo [
161
]. A
novel compound, SGL5213, has been identified as a selective iSGLT1 inhibitor, and has
demonstrated insulin-sensitizing, anti-inflammatory, and antifibrotic activity in a murine
model of NAFLD [162].
6.5. α-Glucosidase Inhibitors
α
-Glucosidase is a carbohydrate hydrolase located in the brush border of the small
intestine that catalyzes the breakdown of dietary starch and disaccharides to yield glucose.
α
-Glucosidase inhibitors slow down carbohydrate digestion and absorption, thereby reduc-
ing postprandial hyperglycaemia. However, they are characterized by only modest overall
antidiabetic activity, and are not too often used in clinical practice [
163
]. Despite some
scientific interest concerning the use of this class of drugs for the treatment of liver diseases,
data regarding their efficacy for NAFLD remain scarce. Acarbose (100 mg/d) improved
AST, ALT levels and lipid profiles, albeit to a lesser extent than ezetimibe, in a 10-week
small-scale RCT in non-diabetic NASH patients [
164
]. Miglitol treatment (
150 mg/d
for
12 months) was associated with significant improvements in steatosis, lobular and portal
inflammation, and NAS scores, while fibrosis and hepatocyte ballooning remained un-
changed [
165
]. Finally, voglibose prevented hepatic steatosis in obese rats, but was slightly
inferior to empagliflozin [166].
An overview of the current drug development pipeline for NAFLD is given in Figure 1.
Biomedicines 2022,10, 274 14 of 34
Biomedicines 2022, 10, x FOR PEER REVIEW 14 of 35
Figure 1. An overview of the current drug development pipeline for non-alcoholic fatty liver dis-
ease. FASN, fatty acid synthase; DGAT, diglyceride acyltransferase; FXR, farnesoid X receptor; FGF,
fibroblast growth factor; TLR4, toll-like receptor 4; LOXL2, lysyl oxidase-like protein 2; ATX, auto-
taxin; SGLT, sodium/glucose contransporter; MPC, mitochondrial pyruvate carrier; PPAR, peroxi-
some proliferator-activated receptor; THRβ, thyroid hormone receptor β; PUFA, polyunsaturated
fatty acid; ACC, acetyl-CoA carboxylase; BUDCA, berberine ursodeoxycholate; LMS, leucine-met-
formin-sildenafil; LMSC, liver-derived mesenchymal stromal cells.
7. Other Agents
7.1. Probiotics
According to recent evidence, NAFLD pathogenesis is tightly associated with intes-
tinal bacterial overgrowth and an upset balance between Bacteroidetes and Firmicutes
species, leading to alterations in the gut-derived metabolite and endotoxin production
[167]. A number of small-scale RCT in paediatric and adult NAFLD patients have demon-
strated the beneficial effects of probiotics containing various combinations of Bifidobacte-
rium spp. [168], Lactobacillus spp. [169,170], Streptococcus thermophilus [171], Pediococcus
pentosaceus [172], Lactococcus spp., Propionibacterium spp., and Acetobacter spp. [173] Two
meta-analyses of 134 and 535 NAFLD patients from 4 and 9 RCTs, respectively, have con-
firmed that probiotics can improve insulin sensitivity, ameliorate dyslipidaemia and
Figure 1.
An overview of the current drug development pipeline for non-alcoholic fatty liver disease.
FASN, fatty acid synthase; DGAT, diglyceride acyltransferase; FXR, farnesoid X receptor; FGF,
fibroblast growth factor; TLR4, toll-like receptor 4; LOXL2, lysyl oxidase-like protein 2; ATX, autotaxin;
SGLT, sodium/glucose contransporter; MPC, mitochondrial pyruvate carrier; PPAR, peroxisome
proliferator-activated receptor; THR
β
, thyroid hormone receptor
β
; PUFA, polyunsaturated fatty
acid; ACC, acetyl-CoA carboxylase; BUDCA, berberine ursodeoxycholate; LMS, leucine-metformin-
sildenafil; LMSC, liver-derived mesenchymal stromal cells.
7. Other Agents
7.1. Probiotics
According to recent evidence, NAFLD pathogenesis is tightly associated with intestinal
bacterial overgrowth and an upset balance between Bacteroidetes and Firmicutes species,
leading to alterations in the gut-derived metabolite and endotoxin production [
167
]. A
number of small-scale RCT in paediatric and adult NAFLD patients have demonstrated the
beneficial effects of probiotics containing various combinations of Bifidobacterium spp. [
168
],
Lactobacillus spp. [
169
,
170
], Streptococcus thermophilus [
171
], Pediococcus pentosaceus [
172
],
Lactococcus spp., Propionibacterium spp., and Acetobacter spp. [
173
] Two meta-analyses of 134
and 535 NAFLD patients from 4 and 9 RCTs, respectively, have confirmed that probiotics can
improve insulin sensitivity, ameliorate dyslipidaemia and systemic inflammation, reduce in-
Biomedicines 2022,10, 274 15 of 34
trahepatic fat content, and rescue impaired liver function, overall improving the clinical out-
comes in NAFLD [
174
,
175
]. Additionally, probiotics based on Saccharomyces boulardii [
176
]
and Clostridium butyricum [
177
] have demonstrated antisteatotic, anti-inflammatory, and
antifibrotic properties in animal models of NAFLD. Clinical evidence for probiotic use in
NAFLD is summarized in Table 1.
Table 1.
Probiotic formulations with evidence of hepatoprotective activity in non-alcoholic fatty liver disease.
Composition
Liver-Related Effects
Cytolysis Steatosis HCB Inflammation Fibrosis Cholestasis References
Bifidobacterium longum + + ± ± [168]
Lactobacillus acidophilus + [178]
L. acidophilus,B. lactis + [170]
L. rhamnosus + [169]
L. plantarum * + [179]
L. paracasei * + + [180]
L. johnsonii * + + [181]
L. reuteri + [182]
L. delbrueckii subsp.
bulgaricus,Streptococcus
thermophilus
+ [183]
L. acidophilus,L. rhamnosus,
B. bifidum,B. lactis + + [184]
L. acidophilus,L. rhamnosus,
L. plantarum,L. delbrueckii
subsp. bulgaricus,B. bifidum
+ + [185]
L. acidophilus,L. rhamnosus,
L. paracasei,B. lactis,B. breve,
Pediococcus pentosaceus
+ [172]
L. acidophilus,L. plantarum,
L. paracasei,L. delbrueckii
subsp. bulgaricus,B. breve,
B. longum, B. infantis
±[171]
Lactobacillus spp.,
Bifidobacterium spp.,
Lactococcus spp.,
Propionibacterium spp.,
Acetobacter spp.
+ + ±+ [173]
L. acidophilus,L. plantarum,
L. delbrueckii subsp.
bulgaricus,L. casei,B. breve,
B. longum,B. infantis,
S. thermophilus
+ [186]
Clostridium butyricum * + + [177]
Saccharomyces boulardii * + + [176]
* Preclinical data given. +, definite positive effect; ±, possible positive effect; HCB, hepatocellular ballooning.
Biomedicines 2022,10, 274 16 of 34
7.2. Mesenchymal Stromal Cells
Lately, cell-based therapy has emerged as a feasible alternative for the treatment of
different stages of NAFLD. In particular, experimental evidence supports the use of bone
marrow- [
187
], umbilical cord- [
188
], and compact bone-derived mesenchymal stromal
cells [
189
] as well as hepatocytes derived by differentiating induced pluripotent stem
cells [
190
]. The crosstalk between hepatic stem cells and their possible therapeutic applica-
tion for NAFLD are discussed in detail in a recent review by Overi et al. [191].
HepaStem
®
is a first-in-class allogeneic stem cell therapy product containing human
adult liver-derived progenitor cells with potential indications including cirrhotic and
precirrhotic NASH as well as acute-on-chronic liver failure (ACLF). HepaStem
®
cells,
obtained from healthy donors, are expected to modulate the inflammatory response and
inhibit HSC activation, thereby reducing liver fibrosis. A small-scale phase 2a RCT found
HepaStem to be safe and well tolerated, and indicated potential efficacy for ACLF and/or
decompensated liver cirrhosis [192].
7.3. Fraudulent Fatty Acids
Fraudulent, or abnormal fatty acids, represented by bempedoic acid (ETC-1002) and
gemcabene (CI-1027), are molecules with structures similar to those of oleic or linolenic
acid that regulate metabolic pathways in the liver, resulting in enhanced FA catabolism.
After conversion into its active form, bempedoic acid acts as false substrate and inhibits
hepatic adenosine triphosphate citrate (pro-S)-lyase, an enzyme upstream of 3-hydroxy-
3-methylglutaryl-CoA reductase in the cholesterol synthesis pathway. This links fraud-
ulent fatty acids to statins, whose possible beneficial effects for NAFLD are reviewed
elsewhere [193].
Bempedoic acid is approved in the USA and EU as monotherapy and as a fixed-
dose combination with ezetimibe for the treatment of hypercholesterolaemia [
194
]. In a
high-fat diet-induced murine model of NASH, it caused significant reductions in ALT and
AST levels, hepatic TAG accumulation, proinflammatory and profibrotic gene expression,
resulting in improved NAFLD activity and liver fibrosis by histological analysis [195].
Gemcabene (PD-72953), a structurally optimized derivative of bempedoic acid, forms
a CoA conjugate that inhibits ACC, and reduces apolipoprotein C-III expression [
194
].
In a mouse model of NASH/HCC, it diminished micro- and macrovesicular liver steato-
sis, HCB, inflammatory infiltration, and fibrosis, which corresponded to downregulated
proinflammatory, lipogenesis, and profibrogenic marker expression [
194
]. Gemcabene
was being developed for the treatment of paediatric NAFLD, but was discontinued and
repurposed for another indication after a lack of efficacy was demonstrated in a phase 2a
proof-of-concept study (NCT03436420).
7.4. Tesamorelin
Tesamorelin (TH9507) is a growth hormone (GH) releasing hormone analogue that
is thought to stimulate lipolysis via increasing endogenous GH levels while maintaining
feedback inhibition and limiting toxicity compared to native GH. Tesamorelin reduced liver
fat content and visceral fat in a preliminary study in antiretroviral-treated patients with
human immunodeficiency virus (HIV)-associated lipodystrophy [
196
]. A phase 2 trial to
evaluate the effects of tesamorelin on liver steatosis and cardiovascular risk in obese NASH
patients is recruiting (NCT03375788), and a phase 3 study in the general population with
NAFLD including a HIV cohort has been planned [197].
7.5. Berberine Ursodeoxycholate
Berberine ursodeoxycholate (BUDCA, HTD1801) is an ionic salt of the isoquino-
line alkaloid berberine and ursodeoxycholic acid (UDCA). According to a meta-analysis
by Wei et al., berberine can significantly improve liver function, lipid profiles, and gly-
caemic control in patients with NAFLD [198] due to adenosine monophosphate-activated
protein kinase (AMPK) activation, stimulation of glycolysis, and, possibly, inhibition of
Biomedicines 2022,10, 274 17 of 34
α
-glucosidase [
199
]. UDCA, in turn, is a bile acid long used for the treatment of NASH
and chronic cholestatic diseases, whose hepatoprotective effects are confirmed by several
systematic reviews and meta-analyses [
200
,
201
]. In a phase 2 proof-of-concept RCT in
T2DM patients with presumed NASH, BUDCA reduced liver enzyme levels and liver
steatosis by MRI-PDFF [202].
7.6. Miricorilant
Miricorilant (CORT 118335) is an investigational glucocorticoid receptor agonist/antagonist
and a mineralocorticoid receptor antagonist currently in development for NASH and antipsychotic-
induced weight gain. Results of a phase 2a study in NASH patients demonstrated that mirico-
rilant (600 mg/d) effectively ameliorated liver steatosis by radiological measures. However,
miricorilant treatment was associated by transient yet significant increases in serum
transaminases [
203
], and the trial was subsequently put on hold due to safety concerns
(NCT03823703).
7.7. Nitazoxanide
Nitazoxanide (Alinia
®
) is an FDA-approved broad-spectrum thiazolide antiprotozoal
and antiparasitic agent, lately reported to be a potent AMPK activator and inhibitor of HSC
activation. In experimental studies in mice, nitazoxanide (100 mg/kg/d) attenuated dyslip-
idaemia, liver steatosis [
204
], fibrosis, inflammation, and HCB, demonstrating synergistic
effects with the pan-PPAR agonist elafibranor [
205
]. Moreover, the anti-anaerobic activity of
nitazoxanide may determine its use in preventing the recurrence of hepatic encephalopathy
as a viable alternative to rifaximin (NCT04161053) [206].
7.8. Pirfenidone
Pirfenidone (Esbriet
®
) is a pyridine derivative with antifibrotic, anti-inflammatory,
and antioxidant properties, the precise mechanisms of which are still unclear. In the liver,
pirfenidone may decrease fibronectin, TGF
β
, collagen production and attenuate fibroge-
nesis, hepatocyte necrosis, and necroinflammation [
207
,
208
]. In the phase 2 PROMETEO
study, pirfenidone (1200 mg/d) markedly reduced transaminase levels and advanced liver
fibrosis of predominantly nonalcoholic aetiology [209].
7.9. Miscellaneous
Other investigational drugs with potential therapeutic value in NAFLD include an-
tileukotriene agents [
210
], GPCR modulators [
211
], anti-IL mABs [
212
], IL22 axis mod-
ulators [
213
], purinergic receptor agonists [
214
], antioxidants [
215
], antisense oligonu-
cleotides [
216
], multitarget epigenetic regulators [
217
], and many more. A comprehensive
list of drug candidates and experimental agents with evidence of hepatoprotective activity
in NAFLD is given in Table 2.
Biomedicines 2022,10, 274 18 of 34
Table 2. Drug candidates and experimental agents with evidence of hepatoprotective activity in non-alcoholic fatty liver disease.
Name Mechanism of Action Development
Phase
Liver-Related Effects
Cytolysis Steatosis HCB Inflammation Fibrosis Cholestasis References
Resmetirom THRβagonist 3 + + ± ± + + [11]
VK2809 THRβagonist 2 + [12]
ASC41 * THRβagonist 2 + + + + [13]
TERN-501 * THRβagonist 1 + + + [15]
Firsocostat ** ACC inhibitor 2 + + + + + ±[23]
Clesacostat ** ACC inhibitor 2 + [24,31]
ASC40 FASN inhibitor 2 + + + + [16]
Aramchol SCD1 inhibitor 3 + + + + + [30]
Ervogastat DGAT2 inhibitor 2 + [31]
ION224 DGAT2 inhibitor 1 + ±[216]
Docosahexaenoic acid ω-3 PUFA - + [40]
Epeleuton ω-3 PUFA 2 ±[44]
Icosabutate ω-3 PUFA 2 + + ±+ [46]
Eicosapentanoic acid * ω-3 PUFA - + + + [218]
Obeticholic acid FXR agonist 3 + + + + + + [49,50]
EDP-305 FXR agonist 2 + + + + [54,56]
Tropifexor FXR agonist 2 + + + + [63]
Cilofexor FXR agonist 2 + + + [64]
Vonafexor FXR agonist 2 + + + [65]
MET409 FXR agonist 2 + + [59]
TERN-101 FXR agonist 2 + + [67]
ASC42 * FXR agonist 2 + + + [16]
INT-767 * FXR agonist 2 + + + + [58]
EDP-297 * FXR agonist 1 + + + ±[219]
BAR502 * FXR agonist - + + + [57]
GW4064 * FXR agonist - + + [220]
Biomedicines 2022,10, 274 19 of 34
Table 2. Cont.
Name Mechanism of Action Development
Phase
Liver-Related Effects
Cytolysis Steatosis HCB Inflammation Fibrosis Cholestasis References
Aldafermin FGF19 analogue 2 + + + [70]
Efruxifermin FGF21 analogue 2 + + ±+ + [71]
BIO89-100 FGF21 analogue 2 + + [72]
BFKB8488A FGFR1c/KLB agonist 2 + [74]
MK-3655 FGFR1c/KLB agonist 2 + [221]
GLP-1-Fc-FGF21 D1 *
FGF21 analogue, GLP1R agonist
- + + [75]
GB1211 * galectin-3 antagonist 2 + [88]
GM-CT-01 * galectin-3/1 antagonist - + + + + [79]
JKB-122 TLR4 antagonist 2 + + [83]
Eritoran * TLR4 antagonist - + + + [81]
PXS-5153A * LOXL2/3 inhibitor - + + [87]
Bezafibrate * PPARαagonist - + + ±[98]
Pemafibrate PPARαagonist - + ±+ [97]
Fenofibrate PPARαagonist - + ±+ [93,94]
Gemfibrozil PPARαagonist + [95]
Nifedipine * PPARγagonist - + + [101]
Seladelpar PPARδagonist 2 + + [103]
Saroglitazar PPARα/γagonist 2 + + + + [104,105]
Lanifibranor PPARα/γ/δagonist 3 + + + + + [106]
Pioglitazone PPARγagonist, MPC inhibitor - + + + + [222]
Lobeglitazone PPARγagonist, MPC inhibitor - + + [100]
Azemiglitazone MPC inhibitor 3 + ± ± ± [107]
PXL065 * MPC inhibitor 2 + + + [109]
Biomedicines 2022,10, 274 20 of 34
Table 2. Cont.
Name Mechanism of Action Development
Phase
Liver-Related Effects
Cytolysis Steatosis HCB Inflammation Fibrosis Cholestasis References
Semaglutide GLP1R agonist 3 + + [113]
Exenatide GLP1R agonist - + + + [115,116]
Lixisenatide GLP1R agonist - + + + [119]
Liraglutide GLP1R agonist - + + + + + [120]
Dulaglutide GLP1R agonist - + + + [117,118]
Teduglutide * GLP2R agonist - + ± ± [113]
Tirzepatide GLP1R/GIPR agonist 2 + + [123,124]
Cotadutide GLP1R/GCGR agonist 2 + + ±[127]
Efinopegdutide * GLP1R/GCGR agonist 2 + ± ± [128]
Pemvidutide GLP1R/GCGR agonist 1 [223]
HM15211 * GLP1R/GCGR/GIPR agonist 2 + + [131]
Sitagliptin DPP4 inhibitor - + + ±[134]
Vildagliptin DPP4 inhibitor - + + + [135]
Saxagliptin DPP4 inhibitor - + + ±+ [136]
Alogliptin DPP4 inhibitor - + [139]
Omarigliptin DPP4 inhibitor - + + + + + [137]
Teneligliptin DPP4 inhibitor - + + [138]
Evogliptin * DPP4 inhibitor - + + + [141]
Anagliptin * DPP4 inhibitor - + + + [142,143]
Trelagliptin * DPP4 inhibitor - + + + ±[144]
Gemigliptin * DPP4 inhibitor - + + + [145]
Linagliptin * DPP4 inhibitor - + + [146]
Dapagliflozin SGLT2 inhibitor - + + ±+ [150,224]
Empagliflozin SGLT2 inhibitor - + + + + + [151,225]
Canagliflozin SGLT2 inhibitor - + + + + [152]
Biomedicines 2022,10, 274 21 of 34
Table 2. Cont.
Name Mechanism of Action Development
Phase
Liver-Related Effects
Cytolysis Steatosis HCB Inflammation Fibrosis Cholestasis References
Ipragliflozin SGLT2 inhibitor - ±+ + [154,226]
Ertugliflozin SGLT2 inhibitor - + [156]
Remogliflozin SGLT2 inhibitor - + + [155]
Luseogliflozin SGLT2 inhibitor - + + + [157,158]
Tofogliflozin SGLT2 inhibitor - + + + + [159]
Licogliflozin ** SGLT1/2 inhibitor 2 + + + [160,161]
SGL5213 * iSGLT1 inhibitor - + + ±+ [162]
Miglitol α-glucosidase inhibitor - + + ±+ + [165]
Acarbose α-glucosidase inhibitor - + [164]
Voglibose * α-glucosidase inhibitor - + [166]
Liver-derived MSC MSC 2 + + [192]
Umbilical cord-derived MSC * MSC - + [188]
Compact bone-derived MSC * MSC - + + + + [189]
Tesamorelin GHRH analogue 3 + + [196]
Berberine ursodeoxycholate multimodal metabolic 2/1 + + + [202]
Miricorilant GR agonist/antagonist, MR
antagonist 2 + [203]
Nitazoxanide * AMPK activator 2 + + [204,205]
PXL770 AMPK activator 2 + [227]
Leucine + metformin +
sildenafil
AMPK activator, eNOS
activator 2 + [228]
Pirfenidone * multimodal antifibrotic - + + [209]
PBI-4547 * GPCR84 antagonist - + + ±[229]
CpdA * GPCR84 antagonist - + + + + [230]
CpdB * GPCR84 antagonist - + + + + [230]
GPR120 agonist III * GPCR120 agonist - + ±+ [231]
Metabolitin * GPRC6A agonist - + + + [232]
SCO-267 * GPR40 agonist - + + + [233]
Biomedicines 2022,10, 274 22 of 34
Table 2. Cont.
Name Mechanism of Action Development
Phase
Liver-Related Effects
Cytolysis Steatosis HCB Inflammation Fibrosis Cholestasis References
Evolocumab PCSK9 inhibitor - ±[234]
Alirocumab PCSK9 inhibitor - ±[234]
X203 * IL11 antagonist - + + + + [212]
X209 * IL11RA antagonist - [212]
Ezetimibe NPC1L1 inhibitor - + ±+ [235]
ORMD-0801 oral insulin 2 + [236]
GalNAc-Stk25 ASO * anti-STK25 ASO - + + + + [237]
Tipelukast * LTR antagonist, PDE3/4
inhibitor, 5-LO/LT inhibitor 2±+ + + [210]
CM-101 * CCL24 antagonist 2 + ± ± ± +±[238]
Namodenoson A3AR agonist 2 + [214,239]
PLN-1474 * αvβ1antagonist 1 + [240]
CB4211 MOTS-c analogue 1 + + [241]
CER-209 * P2Y13R agonist 1 + + [242]
DUR-928
multitarget epigenetic regulator
2 + + + + [217]
CRV431 cyclophilin A/B/D inhibitor 2 + [243]
LPCN 1144 androgen receptor agonist 2 + + + + [244]
Osteocalcin * N/A - + + + + [245]
* Preclinical data given; ** developed in combination; -, not in development. +, definite positive effect;
±
, possible positive effect. HCB, hepatocellular ballooning; THR
β
, thyroid
hormone receptor
β
; ACC, acetyl-CoA carboxylase; FASN, fatty acid synthase; SCD1, stearoyl-CoA desaturase 1; DGAT2, diglyceride acyltransferase 2; PUFA, polyunsaturated fatty acid;
FXR, farnesoid X receptor; FGF, fibroblast growth factor; FGFR1c/KLB, FGF receptor/
β
-klotho complex; TLR4, toll-like receptor 4; LOXL, lysyl oxidase-like protein; PPAR, peroxisome
proliferator-activated receptor; MPC, mitochondrial pyruvate carrier; GLP1R, glucagon-like peptide 1 receptor; GIPR, glucose-dependent insulinotropic polypeptide receptor; GCGR,
glucagon receptor; DPP4, dipeptidyl peptidase 4; SGLT, sodium/glucose cotransporter; iSGLT1, intestinal SGLT 1; MSC, mesenchymal stromal cells; GHRH, growth hormone releasing
hormone; GR, glucocorticoid receptor; MR, mineralocorticoid receptor; AMPK, adenosine monophosphate-activated protein kinase; eNOS, endothelial nitric oxide synthase; GPCR,
G protein-coupled receptor; GPRC6A, G protein-coupled receptor family C group 6 member A; PCSK9, proprotein convertase subtilisin/kexin type 9; IL11, interleukin 11; IL11RA,
IL11 receptor
α
subunit; NPC1L1, Niemann-Pick C1-like protein 1; STK25, serine/threonine kinase 25; ASO, antisense nucleotide; LTR, leukotriene receptor; PDE, phosphodiesterase;
5-LO/LT, 5-lipoxygenase/leukotriene pathway; CCL24, C-C motif chemokine ligand 24; A3AR, A3 adenosine receptor; MOTS-c, mitochondrial open reading frame of the twelve S
ribosomal ribonucleic acid-c; P2Y13R, P2Y purinergic receptor 13.
Biomedicines 2022,10, 274 23 of 34
8. Future Directions
Other biomolecules and pathways most recently identified as potential therapeutic
targets in NAFLD have been highlighted in many comprehensive review articles [
246
249
].
Of those, the sirtuins, a family of highly conserved histone and protein deacetylases, can be
considered of special interest regarding future therapeutic concepts for NAFLD. Sirtuins act
as NAD
+
-sensing signaling proteins to facilitate stress response [
250
], maintain homeostasis
during acute and chronic inflammatory response [
251
], and partake in the regulation of
energy metabolism, redox balance, cell cycle, and suppression of tumorigenesis [252].
Sirtuin 1 (SIRT1), the SIRT1/NF-
κ
B axis, SIRT3, and SIRT4 play a major role in regulat-
ing hepatic lipid metabolism, controlling oxidative stress, and mediating chronic inflamma-
tion in NAFLD and alcoholic fatty liver disease [
253
255
]. Accordingly, SIRT1 activation
by the polyphenol resveratrol and several small molecules have been shown to provide
protection against NAFLD and T2DM in rodent models [
253
,
256
], and may be beneficial
in human metabolic disease [
257
]. SIRT2, SIRT3, and SIRT4 upregulation induced by an
investigational molecule also prevented the progression of hepatic steatosis and fibrosis
in obese rats [
258
]. While still at early research and development stages, selective sirtuin
activators and inhibitors are considered a promising group of drug candidate molecules
with primarily anti-inflammatory mode of action [251,254,257].
9. Conclusions
Being the most prevalent cause of chronic liver disease worldwide, NAFLD is at the
same time one of the greatest areas of unmet medical need in terms of availability and
adequacy of pharmacotherapeutic options. The modern pipeline for NAFLD includes
a plethora of drug candidates with diverse and innovative mechanisms of action, al-
though reported evidence on their clinical effectiveness has so far been limited. Upcoming
treatment approaches to different stages of NAFLD include the modulation of nuclear
receptor activity in order to maintain lipid and glucose homeostasis, stimulation of bile
acid metabolism, and direct inhibition of fibrogenesis, along with a few less explored but
highly promising options.
Despite several investigational agents being seemingly close to regulatory approval
as monotherapies, novel therapeutic strategies for NAFLD will likely involve the use
of multitarget drugs or rational drug combinations. Given the exceptionally complex
pathophysiology and the multifaceted nature of this disease, NAFLD pharmacotherapy
can be expected to remain a priority for biomedical research in the nearest future.
Author Contributions:
Data analysis, writing—original draft preparation, visualization, V.A.P.;
Conceptualization, writing—review and editing, N.N.B. and S.V.O. All authors have read and agreed
to the published version of the manuscript.
Funding: This research was funded by Analytics and Devices Ltd. (St. Petersburg, Russia).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement:
Part of the data analyzed in this study are openly available at Clinical-
Trials.gov, https://www.clinicaltrials.gov/ct2/home (accessed on 10 December 2021).
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
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, 1388–1402. [CrossRef] [PubMed]
2.
Monelli, F.; Venturelli, F.; Bonilauri, L.; Manicardi, E.; Manicardi, V.; Giorgi Rossi, P.; Massari, M.; Ligabue, G.; Riva, N.;
Schianchi, S.; et al. Systematic review of existing guidelines for NAFLD assessment. Hepatoma Res. 2021,7, 25. [CrossRef]
3.
Estes, C.; Razavi, H.; Loomba, R.; Younossi, Z.; Sanyal, A.J. Modeling the epidemic of nonalcoholic fatty liver disease demonstrates
an exponential increase in burden of disease. Hepatology 2018,67, 123–133. [CrossRef] [PubMed]
Biomedicines 2022,10, 274 24 of 34
4.
Eslam, M.; Sanyal, A.J.; George, J.; International Consensus Panel. MAFLD: A Consensus-Driven Proposed Nomenclature for
Metabolic Associated Fatty Liver Disease. Gastroenterology 2020,158, 1999–2014.e1. [CrossRef]
5.
Attia, S.L.; Softic, S.; Mouzaki, M. Evolving Role for Pharmacotherapy in NAFLD/NASH. Clin. Transl. Sci.
2021
,14, 11–19.
[CrossRef] [PubMed]
6.
Yoneda, M.; Honda, Y.; Saito, S.; Nakajima, A. What considerations are there for the pharmacotherapeutic management of
nonalcoholic steatohepatitis? Expert Opin. Pharmacother. 2021,22, 1217–1220. [CrossRef]
7.
Ortiga-Carvalho, T.M.; Sidhaye, A.R.; Wondisford, F.E. Thyroid hormone receptors and resistance to thyroid hormone disorders.
Nat. Rev. Endocrinol. 2014,10, 582–591. [CrossRef]
8.
Pramfalk, C.; Pedrelli, M.; Parini, P. Role of thyroid receptor
β
in lipid metabolism. Biochim. Biophys. Acta
2011
,1812, 929–937.
[CrossRef]
9.
Dawson, P.A.; Parini, P. Hepatic thyroid hormone receptor
β
1 agonism: Good for lipids, good for bile? J. Lipid Res.
2018
,59,
1551–1553. [CrossRef]
10.
Gionfra, F.; De Vito, P.; Pallottini, V.; Lin, H.Y.; Davis, P.J.; Pedersen, J.Z.; Incerpi, S. The Role of Thyroid Hormones in Hepatocyte
Proliferation and Liver Cancer. Front. Endocrinol. 2019,10, 532. [CrossRef]
11.
Harrison, S.A.; Bashir, M.; Moussa, S.E.; McCarty, K.; Frias, J.P.; Taub, R.; Alkhouri, N. Effects of Resmetirom on Noninvasive
Endpoints in a 36-Week Phase 2 Active Treatment Extension Study in Patients with NASH. Hepatol. Commun.
2021
,5, 573–588.
[CrossRef] [PubMed]
12.
Loomba, R.; Neutel, J.; Mohseni, R.; Bernard, D.; Severance, R.; Dao, M.; Saini, S.; Margaritescu, C.; Homer, K.; Tran, B.; et al.
VK2809, a Novel Liver-Directed Thyroid Receptor Beta Agonist, Significantly Reduces Liver Fat with Both Low and High Doses in
Patients with Non-Alcoholic Fatty Liver Disease: A Phase 2 Randomized, Placebo-Controlled Trial. J. Hepatol.
2019
,70, e141–e382.
[CrossRef]
13.
Li, X.; He, H.; Sho, E.; Yang, B.; Wu, J.J. Significant Improvement of NAFLD Activity Scores and Liver Fibrosis by ASC41, a
Selective THR-
β
Agonist, in High Fat Diet Induced NASH SD Rats. In Proceedings of the Digital International Liver Congress
2021, Virtual, 23–26 June 2021.
14.
Wu, J.; He, H.; Palmer, M.; Yan, Y. A Phase Ib Study to Evaluate the Safety, Tolerability, Pharmacokinetics and Pharmacodynamics
of Asc41, A THR-B Agonist, for 28-Days in Overweight and Obese Subjects with Elevated LDL-C, a Population Characteristic of
NAFLD. In Proceedings of the Liver Meeting 2021, Virtual, 12–15 November 2021.
15.
Kirschberg, T.; Jones, C.; Xu, Y.; Wang, Y.; Fenaux, M.; Klucher, K. TERN-501, a Potent and Selective Agonist of Thyroid Hormone
Receptor Beta, Strongly Reduces Histological Features and Biomarkers of Non-Alcoholic Steatohepatitis Associated Pathology in
Rodent Models. In Proceedings of the Digital International Liver Congress 2020, Virtual, 27–29 August 2020.
16. Ascletis Pharma Inc. Available online: https://ascletis.com/ (accessed on 10 December 2021).
17. Terns Pharmaceuticals. Available online: https://www.ternspharma.com/ (accessed on 10 December 2021).
18.
Qi, J.; Lang, W.; Geisler, J.G.; Wang, P.; Petrounia, I.; Mai, S.; Smith, C.; Askari, H.; Struble, G.T.; Williams, R.; et al. The use
of stable isotope-labeled glycerol and oleic acid to differentiate the hepatic functions of DGAT1 and -2. J. Lipid Res.
2012
,53,
1106–1116. [CrossRef] [PubMed]
19.
Yamaguchi, K.; Yang, L.; McCall, S.; Huang, J.; Yu, X.X.; Pandey, S.K.; Bhanot, S.; Monia, B.P.; Li, Y.X.; Diehl, A.M. Inhibiting
triglyceride synthesis improves hepatic steatosis but exacerbates liver damage and fibrosis in obese mice with nonalcoholic
steatohepatitis. Hepatology 2007,45, 1366–1374. [CrossRef] [PubMed]
20.
Takemoto, K.; Fukasaka, Y.; Yoshimoto, R.; Nambu, H.; Yukioka, H. Diacylglycerol acyltransferase 1/2 inhibition induces
dysregulation of fatty acid metabolism and leads to intestinal barrier failure and diarrhea in mice. Physiol. Rep.
2020
,8, e14542.
[CrossRef]
21. Parlati, L.; Régnier, M.; Guillou, H.; Postic, C. New targets for NAFLD. JHEP Rep. 2021,3, 100346. [CrossRef]
22.
Panzitt, K.; Wagner, M. FXR in liver physiology: Multiple faces to regulate liver metabolism. Biochim. Biophys. Acta Mol. Basis Dis.
2021,1867, 166133. [CrossRef]
23.
Loomba, R.; Noureddin, M.; Kowdley, K.V.; Kohli, A.; Sheikh, A.; Neff, G.; Bhandari, B.R.; Gunn, N.; Caldwell, S.H.;
Goodman, Z.; et al.
Combination Therapies Including Cilofexor and Firsocostat for Bridging Fibrosis and Cirrhosis Attributable
to NASH. Hepatology 2021,73, 625–643. [CrossRef]
24.
Calle, R.A.; Amin, N.B.; Carvajal-Gonzalez, S.; Ross, T.T.; Bergman, A.; Aggarwal, S.; Crowley, C.; Rinaldi, A.; Mancuso, J.;
Aggarwal, N.; et al. ACC inhibitor alone or co-administered with a DGAT2 inhibitor in patients with non-alcoholic fatty liver
disease: Two parallel, placebo-controlled, randomized phase 2a trials. Nat. Med. 2021,27, 1836–1848. [CrossRef]
25.
Syed-Abdul, M.M.; Parks, E.J.; Gaballah, A.H.; Bingham, K.; Hammoud, G.M.; Kemble, G.; Buckley, D.; McCulloch, W.;
Manrique-Acevedo, C. Fatty Acid Synthase Inhibitor TVB-2640 Reduces Hepatic de Novo Lipogenesis in Males with Metabolic
Abnormalities. Hepatology 2020,72, 103–118. [CrossRef]
26.
Loomba, R.; Rinella, M.; Harrison, S.A.; Wong, V.W.-S.; Ratziu, V.; Mohseni, R.; Lucas, J.; Gutierrez, J.A.; Rahimi, R.;
Trotter, J.; et al.
Novel, first-in-class, fatty acid synthase inhibitor, TVB-2640 versus placebo demonstrates clinically significant reduction in liver
fat by MRI-PDFF in NASH. In Proceedings of the Digital International Liver Congress 2020, Virtual, 27–29 August 2020.
27. Forma Therapeutics. Available online: https://www.formatherapeutics.com/ (accessed on 10 December 2021).
Biomedicines 2022,10, 274 25 of 34
28.
Bhattacharya, D.; Basta, B.; Mato, J.M.; Craig, A.; Fernández-Ramos, D.; Lopitz-Otsoa, F.; Tsvirkun, D.; Hayardeny, L.; Chandar, V.;
Schwartz, R.E.; et al. Aramchol downregulates stearoyl CoA-desaturase 1 in hepatic stellate cells to attenuate cellular fibrogenesis.
JHEP Rep. 2021,3, 100237. [CrossRef] [PubMed]
29.
Leikin-Frenkel, A.; Gonen, A.; Shaish, A.; Goldiner, I.; Leikin-Gobbi, D.; Konikoff, F.M.; Harats, D.; Gilat, T. Fatty acid bile acid
conjugate inhibits hepatic stearoyl coenzyme A desaturase and is non-atherogenic. Arch. Med. Res.
2010
,41, 397–404. [CrossRef]
[PubMed]
30.
Ratziu, V.; de Guevara, L.; Safadi, R.; Poordad, F.; Fuster, F.; Flores-Figueroa, J.; Arrese, M.; Fracanzani, A.L.; Ben Bashat, D.;
Lackner, K.; et al. Aramchol in patients with nonalcoholic steatohepatitis: A randomized, double-blind, placebo-controlled phase
2b trial. Nat. Med. 2021,27, 1825–1835. [CrossRef] [PubMed]
31.
Calle, R.; Aggarwal, S.; Mancuso, J.; Bergman, A.; Somayaji, V.; Ross, T.; Esler, W. Co-administration of PF-05221304 and PF-
06865571 delivers robust whole liver fat reduction and mitigation of acetyl-coa carboxilase inhibitor induced hypertriglyceridemia
in patients with NAFLD. J. Hepatol. 2020,73, S401–S652. [CrossRef]
32.
Loomba, R.; Morgan, E.; Watts, L.; Xia, S.; Hannan, L.A.; Geary, R.S.; Baker, B.F.; Bhanot, S. Novel antisense inhibition of
diacylglycerol O-acyltransferase 2 for treatment of non-alcoholic fatty liver disease: A multicentre, double-blind, randomised,
placebo-controlled phase 2 trial. Lancet Gastroenterol. Hepatol. 2020,5, 829–838. [CrossRef]
33. Sinew Pharma Inc. Available online: https://www.sinewpharma.com/en/ (accessed on 10 December 2021).
34.
Dentin, R.; Benhamed, F.; Pégorier, J.P.; Foufelle, F.; Viollet, B.; Vaulont, S.; Girard, J.; Postic, C. Polyunsaturated fatty acids
suppress glycolytic and lipogenic genes through the inhibition of ChREBP nuclear protein translocation. J. Clin. Investig.
2005
,
115, 2843–2854. [CrossRef]
35. Calder, P.C. n-3 polyunsaturated fatty acids, inflammation, and inflammatory diseases. Am. J. Clin. Nutr. 2006,83, 1505S–1519S.
[CrossRef]
36.
Siriwardhana, N.; Kalupahana, N.S.; Moustaid-Moussa, N. Health benefits of n-3 polyunsaturated fatty acids: Eicosapentaenoic
acid and docosahexaenoic acid. Adv. Food Nutr. Res. 2012,65, 211–222. [CrossRef]
37.
Okada, L.S.D.R.R.; Oliveira, C.P.; Stefano, J.T.; Nogueira, M.A.; Silva, I.D.C.G.D.; Cordeiro, F.B.; Alves, V.A.F.;
Torrinhas, R.S.
;
Carrilho, F.J.; Puri, P.; et al. Omega-3 PUFA modulate lipogenesis, ER stress, and mitochondrial dysfunction markers in
NASH-Proteomic and lipidomic insight. Clin. Nutr. 2018,37, 1474–1484. [CrossRef]
38.
Nogueira, M.A.; Oliveira, C.P.; Ferreira Alves, V.A.; Stefano, J.T.; Rodrigues, L.S.; Torrinhas, R.S.; Cogliati, B.; Barbeiro, H.;
Carrilho, F.J.; Waitzberg, D.L. Omega-3 polyunsaturated fatty acids in treating non-alcoholic steatohepatitis: A randomized,
double-blind, placebo-controlled trial. Clin. Nutr. 2016,35, 578–586. [CrossRef]
39.
Parker, H.M.; Johnson, N.A.; Burdon, C.A.; Cohn, J.S.; O’Connor, H.T.; George, J. Omega-3 supplementation and non-alcoholic
fatty liver disease: A systematic review and meta-analysis. J. Hepatol. 2012,56, 944–951. [CrossRef] [PubMed]
40.
Nobili, V.; Bedogni, G.; Alisi, A.; Pietrobattista, A.; Risé, P.; Galli, C.; Agostoni, C. Docosahexaenoic acid supplementation
decreases liver fat content in children with non-alcoholic fatty liver disease: Double-blind randomised controlled clinical trial.
Arch. Dis. Child. 2011,96, 350–353. [CrossRef] [PubMed]
41.
He, X.X.; Wu, X.L.; Chen, R.P.; Chen, C.; Liu, X.G.; Wu, B.J.; Huang, Z.M. Effectiveness of Omega-3 Polyunsaturated Fatty Acids in
Non-Alcoholic Fatty Liver Disease: A Meta-Analysis of Randomized Controlled Trials. PLoS ONE
2016
,11, e0162368. [CrossRef]
[PubMed]
42.
Yan, J.H.; Guan, B.J.; Gao, H.Y.; Peng, X.E. Omega-3 polyunsaturated fatty acid supplementation and non-alcoholic fatty liver
disease: A meta-analysis of randomized controlled trials. Medicine 2018,97, e12271. [CrossRef]
43.
Lee, C.H.; Fu, Y.; Yang, S.J.; Chi, C.C. Effects of Omega-3 Polyunsaturated Fatty Acid Supplementation on Non-Alcoholic Fatty
Liver: A Systematic Review and Meta-Analysis. Nutrients 2020,12, 2769. [CrossRef]
44.
Climax, J.; Newsome, P.N.; Hamza, M.; Weissbach, M.; Coughlan, D.; Sattar, N.; McGuire, D.K.; Bhatt, D.L. Effects of Epeleuton, a
Novel Synthetic Second-Generation n-3 Fatty Acid, on Non-Alcoholic Fatty Liver Disease, Triglycerides, Glycemic Control, and
Cardiometabolic and Inflammatory Markers. J. Am. Heart Assoc. 2020,9, e016334. [CrossRef]
45. Affimune. Available online: https://www.afimmune.com/ (accessed on 10 December 2021).
46.
Harrison, S.; Gunn, N.T.; Sheikh, M.Y.; Rudraraju, M.; Kohli, A.; Neff, G.; Round, P.; Fraser, D.A.; Beysen, C.; Rossi, S.; et al.
Icosabutate, a novel structurally engineered fatty acid, significantly reduces relevant markers of NASH and fibrosis in 16 weeks:
Interim analysis results of the ICONA trial. J. Hepatol. 2021,75, S294–S803.
47.
Jiang, L.; Zhang, H.; Xiao, D.; Wei, H.; Chen, Y. Farnesoid X receptor (FXR): Structures and ligands. Comput. Struct. Biotechnol. J.
2021,19, 2148–2159. [CrossRef]
48.
Intercept Pharmaceuticals. Available online: https://www.interceptpharma.com/homepage-non-usa/ (accessed on 10 December 2021).
49.
Mudaliar, S.; Henry, R.R.; Sanyal, A.J.; Morrow, L.; Marschall, H.U.; Kipnes, M.; Adorini, L.; Sciacca, C.I.; Clopton, P.; Castelloe, E.;
et al. Efficacy and safety of the farnesoid X receptor agonist obeticholic acid in patients with type 2 diabetes and nonalcoholic
fatty liver disease. Gastroenterology 2013,145, 574–582.e1. [CrossRef]
50.
Neuschwander-Tetri, B.A.; Loomba, R.; Sanyal, A.J.; Lavine, J.E.; Van Natta, M.L.; Abdelmalek, M.F.; Chalasani, N.; Dasarathy, S.;
Diehl, A.M.; Hameed, B.; 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. [CrossRef]
51.
Alemi, F.; Kwon, E.; Poole, D.P.; Lieu, T.; Lyo, V.; Cattaruzza, F.; Cevikbas, F.; Steinhoff, M.; Nassini, R.; Materazzi, S.; et al. The
TGR5 receptor mediates bile acid-induced itch and analgesia. J. Clin. Investig. 2013,123, 1513–1530. [CrossRef] [PubMed]
Biomedicines 2022,10, 274 26 of 34
52.
An, P.; Wei, G.; Huang, P.; Li, W.; Qi, X.; Lin, Y.; Vaid, K.A.; Wang, J.; Zhang, S.; Li, Y.; et al. A novel non-bile acid FXR agonist
EDP-305 potently suppresses liver injury and fibrosis without worsening of ductular reaction. Liver Int.
2020
,40, 1655–1669.
[CrossRef] [PubMed]
53.
Chau, M.; Li, Y.; Roqueta-Rivera, M.; Garlick, K.; Shen, R.; Wang, G.; Or, Y.S.; Jiang, L.-J. Characterization of EDP-305, a Highly
Potent and Selective Farnesoid X Receptor Agonist, for the Treatment of Non-alcoholic Steatohepatitis. Int. J. Gastroenterol.
2019
,
3, 4–16. [CrossRef]
54.
Enanta Pharmaceuticals, Inc. Enanta Announces Results of INTREPID Study of EDP-305 for the Treatment of Primary Biliary
Cholangitis: News Release; 2020. Available online: https://www.enanta.com/investors/news-releases/press-release/2020/Enanta-
Announces-Results-of-INTREPID-Study-of- EDP-305-for-the- Treatment-of-Primary-Biliary-Cholangitis/default.aspx (accessed
on 10 December 2021).
55.
Biagioli, M.; Fiorucci, S. Bile acid activated receptors: Integrating immune and metabolic regulation in non-alcoholic fatty liver
disease. Liver Res. 2021,5, 119–141. [CrossRef]
56.
Enanta Pharmaceuticals, Inc. Enanta Pharmaceuticals Provides Update on NASH FXR Agonist Programs: News Release; 2021. Available
online: https://www.enanta.com/investors/news-releases/press-release/2021/Enanta-Pharmaceuticals-Provides-Update-
on-NASH-FXR-Agonist-Programs/ (accessed on 10 December 2021).
57.
Carino, A.; Cipriani, S.; Marchianò, S.; Biagioli, M.; Santorelli, C.; Donini, A.; Zampella, A.; Monti, M.C.; Fiorucci, S. BAR502, a
dual FXR and GPBAR1 agonist, promotes browning of white adipose tissue and reverses liver steatosis and fibrosis. Sci. Rep.
2017,7, 42801. [CrossRef]
58.
Roth, J.D.; Feigh, M.; Veidal, S.S.; Fensholdt, L.K.; Rigbolt, K.T.; Hansen, H.H.; Chen, L.C.; Petitjean, M.; Friley, W.;
Vrang, N.; et al.
INT-767 improves histopathological features in a diet-induced ob/ob mouse model of biopsy-confirmed non-alcoholic steatohep-
atitis. World J. Gastroenterol. 2018,24, 195–210. [CrossRef]
59.
Harrison, S.A.; Bashir, M.R.; Lee, K.J.; Shim-Lopez, J.; Lee, J.; Wagner, B.; Smith, N.D.; Chen, H.C.; Lawitz, E.J. A structurally
optimized FXR agonist, MET409, reduced liver fat content over 12 weeks in patients with non-alcoholic steatohepatitis. J. Hepatol.
2021,75, 25–33. [CrossRef]
60.
Kremoser, C. FXR agonists for NASH: How are they different and what difference do they make? J. Hepatol.
2021
,75, 12–15.
[CrossRef]
61.
Schumacher, J.D.; Guo, G.L. Pharmacologic Modulation of Bile Acid-FXR-FGF15/FGF19 Pathway for the Treatment of Nonalco-
holic Steatohepatitis. Handb. Exp. Pharmacol. 2019,256, 325–357. [CrossRef]
62.
Lucas, K.J.; Lopez, P.; Lawitz, E.; Sheikh, A.; Aizenberg, D.; Hsia, S.; Boon Bee, G.G.; Vierling, J.; Frias, J.; White, J.; et al. Tropifexor,
a highly potent FXR agonist, produces robust and dose-dependent reductions in hepatic fat and serum alanine aminotransferase
in patients with fibrotic NASH after 12 weeks of therapy: FLIGHT-FXR Part C interim results. Dig. Liver Dis.
2020
,52, e19–e45.
[CrossRef]
63.
Novartis. A Randomized, Double-Blind, Placebo Controlled, 3-Part, Adaptive Design, Multicenter Study to Assess Safety, Tolera-
bility and Efficacy of Tropifexor (LJN452) in Patients with Non-Alcoholic Steatohepatitis (NASH): FLIGHT-FXR, 2020. Available
online: https://www.novctrd.com/ctrdweb/trialresult/trialresults/pdf?trialResultId=17816 (accessed on 10 December 2021).
64.
Patel, K.; Harrison, S.A.; Elkhashab, M.; Trotter, J.F.; Herring, R.; Rojter, S.E.; Kayali, Z.; Wong, V.W.; Greenbloom, S.;
Jayakumar, S.; et al.
Cilofexor, a Nonsteroidal FXR Agonist, in Patients with Noncirrhotic NASH: A Phase 2 Randomized
Controlled Trial. Hepatology 2020,72, 58–71. [CrossRef] [PubMed]
65.
Harrison, S.A.; Ratziu, V.; White, A.; Reiss, G.M.; Bowman, W.K.; Gunn, N.; Loustaud Ratti, V.; Bureau, C.; Lawitz, E.J.;
Alkhouri, N.; et al.
Vonafexor, s FXR Agonist, Induced Hepatic snd Renal Improvement in the Randomized, Double-Blind,
Placebocontrolled LIVIFYNASH Trial. In Proceedings of the Liver Meeting 2021, Virtual, 12–15 November 2021.
66.
Novartis. A randomized, Patient and Investigator Blinded, Placebo-Controlled, Multicenter Study to Assess the Safety, Tolerability,
Pharmacokinetics and Efficacy of LMB763 in Patients with Non-Alcoholic Steatohepatitis (NASH), 2020. Available online:
https://www.novctrd.com/ctrdweb/trialresult/trialresults/pdf?trialResultId=17527 (accessed on 10 December 2021).
67.
Loomba, R.; Kowdley, K.V.; Vuppalanchi, R.; Hassanein, T.; Rojter, S.E.; Sheikh, M.Y.; Moussa, S.; Chung, D.; Eng, C.;
Marmon, T.; et al.
Liver-Distributed FXR Agonist TERN-101 Demonstrates Favorable Safety and Efficacy Profile in NASH
Phase 2a LIFT Study. Hepatology 2021,74, 97A–98A.
68. Hepagene Therapeutics, Inc. Available online: http://www.hepagene.com/#/m_about2 (accessed on 10 December 2021).
69.
Henriksson, E.; Andersen, B. FGF19 and FGF21 for the Treatment of NASH-Two Sides of the Same Coin? Differential and
Overlapping Effects of FGF19 and FGF21 from Mice to Human. Front. Endocrinol. 2020,11, 601349. [CrossRef] [PubMed]
70.
Harrison, S.A.; Abdelmalek, M.F.; Neff, G.W.; Gunn, N.; Guy, C.D.; Alkhouri, N.; Bashir, M.; Freilich, B.; Almeda, J.;
Knapple, W.; et al.
Topline Results from the ALPINE 2/3 Study: A Randomized, Double-Blind, Placebo-Controlled, Multicenter,
Phase 2b Trial Evaluating 3 Doses of the FGF19 Analogue Aldafermin on Liver Histology in Patients with Nonalcoholic
Steatohepatitis and Stage 2 or 3 Fibrosis. Hepatology 2021,74, 5A.
71.
Harrison, S.A.; Ruane, P.J.; Freilich, B.L.; Neff, G.; Patil, R.; Behling, C.A.; Hu, C.; Fong, E.; de Temple, B.; Tillman, E.J.; et al.
Efruxifermin in non-alcoholic steatohepatitis: A randomized, double-blind, placebo-controlled, phase 2a trial. Nat. Med.
2021
,27,
1262–1271. [CrossRef]
72.
Frias, J.P.; Lawitz, E.J.; Ortiz-LaSanta, G.; Franey, B.; Morrow, L.; Chen, C.-Y.; Tseng, L.; Charlton, R.W.; Mansbach, H.;
Margalit, M.; et al.
BIO89-100 Demonstrated Robust Reductions in Liver Fat and Liver Fat Volume (LFV) by MRI-PDFF, Favorable
Biomedicines 2022,10, 274 27 of 34
Tolerability and Potential for Weekly (QW) or Every 2 Weeks (Q2W) Dosing in a Phase 1b/2a Placebo-Controlled, Double-Blind,
Multiple Ascending Dose Study in NASH. J. Endocr. Soc. 2021,5, A5–A6. [CrossRef]
73. Novo Nordisk. Available online: https://www.novonordisk.com/ (accessed on 10 December 2021).
74.
Baruch, A.; Wong, C.; Chinn, L.W.; Vaze, A.; Sonoda, J.; Gelzleichter, T.; Chen, S.; Lewin-Koh, N.; Morrow, L.; Dheerendra, S.; et al.
Antibody-mediated activation of the FGFR1/Klotho
β
complex corrects metabolic dysfunction and alters food preference in obese
humans. Proc. Natl. Acad. Sci. USA 2020,117, 28992–29000. [CrossRef]
75.
Pan, Q.; Lin, S.; Li, Y.; Liu, L.; Li, X.; Gao, X.; Yan, J.; Gu, B.; Chen, X.; Li, W.; et al. A novel GLP-1 and FGF21 dual agonist has
therapeutic potential for diabetes and non-alcoholic steatohepatitis. EBioMedicine 2021,63, 103202. [CrossRef]
76.
Abdelmalek, M.F.; Sanyal, A.J.; Nakajima, A.; Neuschwander-Tetri, B.A.; Goodman, Z.; Lawitz, E.J.; Harrison, S.A.; Jacobson, I.M.;
Imajo, K.; Gunn, N.; et al. Efficacy and Safety of Pegbelfermin in Patients with Nonalcoholic Steatohepatitis and Compensated
Cirrhosis: Results from the Phase 2b, Randomized, Double-Blind, Placebo-Controlled Falcon 2 Study. In Proceedings of the Liver
Meeting 2021, Virtual, 12–15 November 2021.
77. Sun, M.J.; Cao, Z.Q.; Leng, P. The roles of galectins in hepatic diseases. J. Mol. Histol. 2020,51, 473–484. [CrossRef]
78.
An, Y.; Xu, S.; Liu, Y.; Xu, X.; Philips, C.A.; Chen, J.; Méndez-Sánchez, N.; Guo, X.; Qi, X. Role of Galectins in the Liver Diseases: A
Systematic Review and Meta-Analysis. Front. Med. 2021,8, 744518. [CrossRef]
79.
Traber, P.G.; Chou, H.; Zomer, E.; Hong, F.; Klyosov, A.; Fiel, M.I.; Friedman, S.L. Regression of fibrosis and reversal of cirrhosis in
rats by galectin inhibitors in thioacetamide-induced liver disease. PLoS ONE 2013,8, e75361. [CrossRef] [PubMed]
80.
Chalasani, N.; Abdelmalek, M.F.; Garcia-Tsao, G.; Vuppalanchi, R.; Alkhouri, N.; Rinella, M.; Noureddin, M.; Pyko, M.;
Shiffman, M.
; Sanyal, A.; et al. Effects of Belapectin, an Inhibitor of Galectin-3, in Patients with Nonalcoholic Steatohepatitis with
Cirrhosis and Portal Hypertension. Gastroenterology 2020,158, 1334–1345.e5. [CrossRef] [PubMed]
81.
Hsieh, Y.C.; Lee, K.C.; Wu, P.S.; Huo, T.I.; Huang, Y.H.; Hou, M.C.; Lin, H.C. Eritoran Attenuates Hepatic Inflammation and
Fibrosis in Mice with Chronic Liver Injury. Cells 2021,10, 1562. [CrossRef]
82.
Yang, L.; Seki, E. Toll-like receptors in liver fibrosis: Cellular crosstalk and mechanisms. Front. Physiol.
2012
,3, 138. [CrossRef]
[PubMed]
83.
Huang, K.C.; Chen, M.Y.; Lin, C.L.; Yang, S.S.; Hsu, C.W.; Wang, C.C.; Huang, Y.H.; Chang, C.C.; Shih, Y.C.; Liu, S.H. JKB-111 in
Patients with Non-Alcoholic Fatty Liver: A Phase 2 Randomized Double-Blind Placebo-Control Study. In Proceedings of the
Digital International Liver Congress 2020, Virtual, 27–29 August 2020.
84.
Klepfish, M.; Gross, T.; Vugman, M.; Afratis, N.A.; Havusha-Laufer, S.; Brazowski, E.; Solomonov, I.; Varol, C.; Sagi, I. LOXL2
Inhibition Paves the Way for Macrophage-Mediated Collagen Degradation in Liver Fibrosis. Front. Immunol.
2020
,11, 480.
[CrossRef]
85.
Chen, W.; Yang, A.; Jia, J.; Popov, Y.V.; Schuppan, D.; You, H. Lysyl Oxidase (LOX) Family Members: Rationale and Their Potential
as Therapeutic Targets for Liver Fibrosis. Hepatology 2020,72, 729–741. [CrossRef]
86.
Harrison, S.A.; Abdelmalek, M.F.; Caldwell, S.; Shiffman, M.L.; Diehl, A.M.; Ghalib, R.; Lawitz, E.J.; Rockey, D.C.; Schall, R.A.;
Jia, C.; et al. Simtuzumab Is Ineffective for Patients with Bridging Fibrosis or Compensated Cirrhosis Caused by Nonalcoholic
Steatohepatitis. Gastroenterology 2018,155, 1140–1153. [CrossRef]
87.
Schilter, H.; Findlay, A.D.; Perryman, L.; Yow, T.T.; Moses, J.; Zahoor, A.; Turner, C.I.; Deodhar, M.; Foot, J.S.; Zhou, W.; et al. The
lysyl oxidase like 2/3 enzymatic inhibitor, PXS-5153A, reduces crosslinks and ameliorates fibrosis. J. Cell. Mol. Med.
2019
,23,
1759–1770. [CrossRef]
88. Galecto, Inc. Available online: https://galecto.com/ (accessed on 10 December 2021).
89.
Geraldo, L.H.M.; Spohr, T.C.L.S.; Amaral, R.F.D.; Fonseca, A.C.C.D.; Garcia, C.; Mendes, F.A.; Freitas, C.; dos Santos, M.F.; Lima,
F.R.S. Role of lysophosphatidic acid and its receptors in health and disease: Novel therapeutic strategies. Signal Transduct. Target.
Ther. 2021,6, 45. [CrossRef]
90.
Bain, G.; Shannon, K.E.; Huang, F.; Darlington, J.; Goulet, L.; Prodanovich, P.; Ma, G.L.; Santini, A.M.; Stein, A.J.; Lonergan, D.;
et al. Selective Inhibition of Autotaxin Is Efficacious in Mouse Models of Liver Fibrosis. J. Pharmacol. Exp. Ther.
2017
,360, 1–13.
[CrossRef]
91. TaiwanJ Pharmaceuticals. Available online: https://www.taiwanj.com/pages/page_index_en (accessed on 10 December 2021).
92.
Boeckmans, J.; Natale, A.; Rombaut, M.; Buyl, K.; Rogiers, V.; De Kock, J.; Vanhaecke, T.; Rodrigues, R.M. Anti-NASH Drug
Development Hitches a Lift on PPAR Agonism. Cells 2019,9, 37. [CrossRef] [PubMed]
93.
Athyros, V.G.; Mikhailidis, D.P.; Didangelos, T.P.; Giouleme, O.I.; Liberopoulos, E.N.; Karagiannis, A.; Kakafika, A.I.; Tziomalos,
K.; Burroughs, A.K.; Elisaf, M.S. Effect of multifactorial treatment on non-alcoholic fatty liver disease in metabolic syndrome: A
randomised study. Curr. Med. Res. Opin. 2006,22, 873–883. [CrossRef] [PubMed]
94.
Fernández-Miranda, C.; Pérez-Carreras, M.; Colina, F.; López-Alonso, G.; Vargas, C.; Solís-Herruzo, J.A. A pilot trial of fenofibrate
for the treatment of non-alcoholic fatty liver disease. Dig. Liver Dis. 2008,40, 200–205. [CrossRef] [PubMed]
95.
Basaranoglu, M.; Acbay, O.; Sonsuz, A. A controlled trial of gemfibrozil in the treatment of patients with nonalcoholic steatohep-
atitis. J. Hepatol. 1999,31, 384. [CrossRef]
96.
Laurin, J.; Lindor, K.D.; Crippin, J.S.; Gossard, A.; Gores, G.J.; Ludwig, J.; Rakela, J.; McGill, D.B. Ursodeoxycholic acid or
clofibrate in the treatment of non-alcohol-induced steatohepatitis: A pilot study. Hepatology 1996,23, 1464–1467. [CrossRef]
Biomedicines 2022,10, 274 28 of 34
97.
Hatanaka, T.; Kosone, T.; Saito, N.; Takakusagi, S.; Tojima, H.; Naganuma, A.; Takagi, H.; Uraoka, T.; Kakizaki, S. Effect of
48-week pemafibrate on non-alcoholic fatty liver disease with hypertriglyceridemia, as evaluated by the FibroScan-aspartate
aminotransferase score. JGH Open 2021,5, 1183–1189. [CrossRef]
98.
Sasaki, Y.; Shimada, T.; Iizuka, S.; Suzuki, W.; Makihara, H.; Teraoka, R.; Tsuneyama, K.; Hokao, R.; Aburada, M. Effects of
bezafibrate in nonalcoholic steatohepatitis model mice with monosodium glutamate-induced metabolic syndrome. Eur. J.
Pharmacol. 2011,662, 1–8. [CrossRef]
99.
Chalasani, N.; Younossi, Z.; Lavine, J.E.; Charlton, M.; Cusi, K.; Rinella, M.; Harrison, S.A.; Brunt, E.M.; Sanyal, A.J. The diagnosis
and management of nonalcoholic fatty liver disease: Practice guidance from the American Association for the Study of Liver
Diseases. Hepatology 2018,67, 328–357. [CrossRef]
100.
Lee, Y.H.; Kim, J.H.; Kim, S.R.; Jin, H.Y.; Rhee, E.J.; Cho, Y.M.; Lee, B.W. Lobeglitazone, a Novel Thiazolidinedione, Improves
Non-Alcoholic Fatty Liver Disease in Type 2 Diabetes: Its Efficacy and Predictive Factors Related to Responsiveness. J. Korean
Med. Sci. 2017,32, 60–69. [CrossRef]
101.
Nakagami, H.; Shimamura, M.; Miyake, T.; Shimosato, T.; Minobe, N.; Moritani, T.; Kiomy Osako, M.; Nakagami, F.; Koriyama, H.;
Kyutoku, M.; et al. Nifedipine prevents hepatic fibrosis in a non-alcoholic steatohepatitis model induced by an L-methionine-and
choline-deficient diet. Mol. Med. Rep. 2012,5, 37–40. [CrossRef]
102.
Zarei, M.; Barroso, E.; Palomer, X.; Dai, J.; Rada, P.; Quesada-López, T.; Escolà-Gil, J.C.; Cedó, L.; Zali, M.R.; Molaei, M.; et al.
Hepatic regulation of VLDL receptor by PPAR
β
/
δ
and FGF21 modulates non-alcoholic fatty liver disease. Mol. Metab.
2018
,8,
117–131. [CrossRef] [PubMed]
103.
CymaBay Therapeutics. CymaBay Therapeutics Reports Topline 12-Week Data from an Ongoing Phase 2b Study of Seladelpar in Patients
with Nonalcoholic Steatohepatitis; 2019. Available online: https://www.globenewswire.com/news-release/2019/06/11/1866763
/0/en/CymaBay-Therapeutics-Reports-Topline-12-Week-Data-from-an-Ongoing-Phase-2b-Study-of-Seladelpar-in-Patients-
with-Nonalcoholic-Steatohepatitis.html (accessed on 10 December 2021).
104.
Goyal, O.; Nohria, S.; Goyal, P.; Kaur, J.; Sharma, S.; Sood, A.; Chhina, R.S. Saroglitazar in patients with non-alcoholic fatty liver
disease and diabetic dyslipidemia: A prospective, observational, real world study. Sci. Rep.
2020
,10, 21117. [CrossRef] [PubMed]
105.
Gawrieh, S.; Noureddin, M.; Loo, N.; Mohseni, R.; Awasty, V.; Cusi, K.; Kowdley, K.V.; Lai, M.; Schiff, E.; Parmar, D.; et al.
Saroglitazar, a PPAR-
α
/
γ
Agonist, for Treatment of NAFLD: A Randomized Controlled Double-Blind Phase 2 Trial. Hepatology
2021,74, 1809–1824. [CrossRef] [PubMed]
106.
Francque, S.M.; Bedossa, P.; Ratziu, V.; Anstee, Q.M.; Bugianesi, E.; Sanyal, A.J.; Loomba, R.; Harrison, S.A.; Balabanska, R.;
Mateva, L.; et al. A Randomized, Controlled Trial of the Pan-PPAR Agonist Lanifibranor in NASH. N. Engl. J. Med.
2021
,385,
1547–1558. [CrossRef]
107.
Harrison, S.A.; Alkhouri, N.; Davison, B.A.; Sanyal, A.; Edwards, C.; Colca, J.R.; Lee, B.H.; Loomba, R.; Cusi, K.;
Kolterman, O.; et al.
Insulin sensitizer MSDC-0602K in non-alcoholic steatohepatitis: A randomized, double-blind, placebo-
controlled phase IIb study. J Hepatol. 2020,72, 613–626. [CrossRef]
108.
Fukunaga, T.; Zou, W.; Rohatgi, N.; Colca, J.R.; Teitelbaum, S.L. An insulin-sensitizing thiazolidinedione, which minimally
activates PPARγ, does not cause bone loss. J. Bone Miner. Res. 2015,30, 481–488. [CrossRef]
109.
Jacques, V.; Bolze, S.; Hallakou-Bozec, S.; Czarnik, A.W.; Divakaruni, A.S.; Fouqueray, P.; Murphy, A.N.; Van der Ploeg, L.H.T.;
DeWitt, S. Deuterium-Stabilized (R)-Pioglitazone (PXL065) Is Responsible for Pioglitazone Efficacy in NASH yet Exhibits Little to
No PPARγActivity. Hepatol. Commun. 2021,5, 1412–1425. [CrossRef]
110.
El, K.; Gray, S.M.; Capozzi, M.E.; Knuth, E.R.; Jin, E.; Svendsen, B.; Clifford, A.; Brown, J.L.; Encisco, S.E.; Chazotte, B.M.; et al.
GIP mediates the incretin effect and glucose tolerance by dual actions on
α
cells and
β
cells. Sci. Adv.
2021
,7, eabf1948. [CrossRef]
111.
Wang, X.C.; Gusdon, A.M.; Liu, H.; Qu, S. Effects of glucagon-like peptide-1 receptor agonists on non-alcoholic fatty liver disease
and inflammation. World J. Gastroenterol. 2014,20, 14821–14830. [CrossRef]
112.
Samson, S.L.; Sathyanarayana, P.; Jogi, M.; Gonzalez, E.V.; Gutierrez, A.; Krishnamurthy, R.; Muthupillai, R.; Chan, L.; Bajaj, M.
Exenatide decreases hepatic fibroblast growth factor 21 resistance in non-alcoholic fatty liver disease in a mouse model of obesity
and in a randomised controlled trial. Diabetologia 2011,54, 3093–3100. [CrossRef]
113.
Newsome, P.N.; Buchholtz, K.; Cusi, K.; Linder, M.; Okanoue, T.; Ratziu, V.; Sanyal, A.J.; Sejling, A.S.; Harrison, S.A.; NN9931-4296
Investigators. A Placebo-Controlled Trial of Subcutaneous Semaglutide in Nonalcoholic Steatohepatitis. N. Engl. J. Med.
2021
,
384, 1113–1124. [CrossRef] [PubMed]
114.
Liava, C.; Sinakos, E. Semaglutide for nonalcoholic steatohepatitis: Closer to a solution? Hepatobiliary Surg. Nutr.
2021
,10,
541–544. [CrossRef] [PubMed]
115.
Dutour, A.; Abdesselam, I.; Ancel, P.; Kober, F.; Mrad, G.; Darmon, P.; Ronsin, O.; Pradel, V.; Lesavre, N.; Martin, J.C.; et al.
Exenatide decreases liver fat content and epicardial adipose tissue in patients with obesity and type 2 diabetes: A prospective
randomized clinical trial using magnetic resonance imaging and spectroscopy. Diabetes Obes. Metab.
2016
,18, 882–891. [CrossRef]
[PubMed]
116.
Sofogianni, A.; Filippidis, A.; Chrysavgis, L.; Tziomalos, K.; Cholongitas, E. Glucagon-like peptide-1 receptor agonists in
non-alcoholic fatty liver disease: An update. World J. Hepatol. 2020,12, 493–505. [CrossRef]
117.
Cusi, K.; Sattar, N.; García-Pérez, L.E.; Pavo, I.; Yu, M.; Robertson, K.E.; Karanikas, C.A.; Haupt, A. 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, 1434–1439. [CrossRef]
Biomedicines 2022,10, 274 29 of 34
118.
Kuchay, M.S.; Krishan, S.; Mishra, S.K.; Choudhary, N.S.; Singh, M.K.; Wasir, J.S.; Kaur, P.; Gill, H.K.; Bano, T.;
Farooqui, K.J.; et al.
Effect of dulaglutide on liver fat in patients with type 2 diabetes and NAFLD: Randomised controlled trial (D-LIFT trial).
Diabetologia 2020,63, 2434–2445. [CrossRef]
119.
Koutsovasilis, A.; Sotiropoulos, A.; Pappa, M.; Apostolou, O.; Binikos, I.; Kordinas, V.; Papadaki, D.; Tamvakos, I.; Bousboulas,
S. Effectiveness of lixisenatide in nonalcoholic fatty liver disease in patients with type 2 diabetes after an acute coronary
syndrome compared to sitagliptin and pioglitazone. In Proceedings of the EASD Virtual Meeting 2016, Munich, Germany, 14
September 2016.
120.
Armstrong, M.J.; Gaunt, P.; Aithal, G.P.; Barton, D.; Hull, D.; Parker, R.; Hazlehurst, J.M.; Guo, K.; LEAN Trial Team;
Abouda, G.; 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. [CrossRef]
121.
Ghosal, S.; Datta, D.; Sinha, B. A meta-analysis of the effects of glucagon-like-peptide 1 receptor agonist (GLP1-RA) in nonalcoholic
fatty liver disease (NAFLD) with type 2 diabetes (T2D). Sci. Rep. 2021,11, 22063. [CrossRef]
122.
Thondam, S.K.; Cuthbertson, D.J.; Wilding, J.P.H. The influence of Glucose-dependent Insulinotropic Polypeptide (GIP) on human
adipose tissue and fat metabolism: Implications for obesity, type 2 diabetes and Non-Alcoholic Fatty Liver Disease (NAFLD).
Peptides 2020,125, 170208. [CrossRef]
123.
Hartman, M.L.; Sanyal, A.J.; Loomba, R.; Wilson, J.M.; Nikooienejad, A.; Bray, R.; Karanikas, C.A.; Duffin, K.L.; Robins, D.A.;
Haupt, A. Effects of Novel Dual GIP and GLP-1 Receptor Agonist Tirzepatide on Biomarkers of Nonalcoholic Steatohepatitis in
Patients with Type 2 Diabetes. Diabetes Care 2020,43, 1352–1355. [CrossRef]
124.
Gastaldelli, A.; Cusi, K.; Landó, F.; Bray, R.; Brouwers, B.; Rodríguez, Á. Effect of tirzepatide versus insulin degludec on liver fat
content and abdominal adipose tissue in patients with type 2 diabetes (SURPASS-3 MRI). In Proceedings of the EASD Virtual
Meeting 2021, Virtual, 30 September 2021.
125. Pocai, A. Action and therapeutic potential of oxyntomodulin. Mol. Metab. 2013,3, 241–251. [CrossRef] [PubMed]
126.
Boland, M.L.; Laker, R.C.; Mather, K.; Nawrocki, A.; Oldham, S.; Boland, B.B.; Lewis, H.; Conway, J.; Naylor, J.; Guionaud, S.; et al.
Resolution of NASH and hepatic fibrosis by the GLP-1R/GcgR dual-agonist Cotadutide via modulating mitochondrial function
and lipogenesis. Nat. Metab. 2020,2, 413–431. [CrossRef] [PubMed]
127.
Nahra, R.; Wang, T.; Gadde, K.M.; Oscarsson, J.; Stumvoll, M.; Jermutus, L.; Hirshberg, B.; Ambery, P. Effects of Cotadutide on
Metabolic and Hepatic Parameters in Adults with Overweight or Obesity and Type 2 Diabetes: A 54-Week Randomized Phase 2b
Study. Diabetes Care 2021,44, 1433–1442. [CrossRef] [PubMed]
128.
Jung, S.Y.; Park, Y.J.; Lee, J.S.; Kim, E.J.; Lee, Y.M.; Kim, Y.H.; Trautmann, M.; Kwon, S.C. Potent Cholesterol Lowering Effect
by HM12525A, A Novel Long-acting GLP-1/Glucagon Dual Receptor Agonist. In Proceedings of the American Diabetes
Association’s (ADA) 76th Scientific Sessions, New Orleans, LA, USA, 10–14 June 2016.
129. Pfizer Inc. Available online: https://www.pfizer.com/ (accessed on 10 December 2021).
130.
Hanmi Pharmaceutical Co., Ltd. Available online: http://www.hanmipharm.com/ehanmi/handler/Home-Start (accessed on 10
December 2021).
131.
Kim, J.K.; Lee, J.S.; Park, E.; Lee, J.; Bae, S.; Kim, D.; Choi, I.C. HM15211, a Novel Long-Acting GLP-1/GIP/Glucagon Triple
Agonist, Exhibits Anti-inflammatory and Fibrotic Effects in AMLN/TAA–Induced Liver Inflammation and Fibrosis Mice. Diabetes
2020,69, 1804-P. [CrossRef]
132.
Janssen, P.; Rotondo, A.; Mulé, F.; Tack, J. Review article: A comparison of glucagon-like peptides 1 and 2. Aliment. Pharmacol.
Ther. 2013,37, 18–36. [CrossRef]
133.
Hu, Y.-X. Therapeutic effect of teduglutide on non-alcoholic fatty liver disease in rats. World Chin. J. Dig.
2016
,24, 1009. [CrossRef]
134.
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. [CrossRef]
135.
Akta¸s, A.; Ozan, Z.T. The efficacy and safety of vildagliptin treatment for nonalcoholic fatty liver disease in type 2 diabetes
mellitus. Cumhur. Med. J. 2020,42, 491–499. [CrossRef]
136.
Li, J.J.; Zhang, P.; Fan, B.; Guo, X.L.; Zheng, Z.S. The efficacy of saxagliptin in T2DM patients with non-alcoholic fatty liver disease:
Preliminary data. Rev. Assoc. Med. Bras. 2019,65, 33–37. [CrossRef]
137.
Hattori, S.; Nomoto, K.; Suzuki, T.; Hayashi, S. Beneficial effect of omarigliptin on diabetic patients with non-alcoholic fatty liver
disease/non-alcoholic steatohepatitis. Diabetol. Metab. Syndr. 2021,13, 28. [CrossRef]
138.
Gupta, V.K. Teneligliptin Significantly Reduces Liver Fat Content (LFC) and Delays Progression of NASH in Type 2 Diabetes
Mellitus Patients. Diabetes 2019,68, 1029-P. [CrossRef]
139.
Mashitani, T.; Noguchi, R.; Okura, Y.; Namisaki, T.; Mitoro, A.; Ishii, H.; Nakatani, T.; Kikuchi, E.; Moriyasu, H.;
Matsumoto, M.; et al.
Efficacy of alogliptin in preventing non-alcoholic fatty liver disease progression in patients with
type 2 diabetes. Biomed. Rep. 2016,4, 183–187. [CrossRef]
140.
Dos Santos, L.R.; Duarte, M.L.; Peccin, M.S.; Gagliardi, A.R.T.; Melnik, T. Dipeptidyl Peptidase IV Inhibitors for Nonalcoholic
Fatty Liver Disease—Systematic Review and Metanalysis. Curr. Diabetes Rev. 2021,17, e101120187811. [CrossRef] [PubMed]
141.
Wang, Z.; Park, H.; Bae, E.J. Efficacy of evogliptin and cenicriviroc against nonalcoholic steatohepatitis in mice: A comparative
study. Korean J. Physiol. Pharmacol. 2019,23, 459–466. [CrossRef] [PubMed]
Biomedicines 2022,10, 274 30 of 34
142.
Sakai, Y.; Nagashimada, M.; Nagata, N.; Ni, Y.; Kaneko, S.; Ota, T. Dipeptidyl peptidase-4 inhibitor anagliptin alleviates
lipotoxicity-induced hepatic insulin resistance and steatohepatitis in mice. In Proceedings of the EASD Virtual Meeting 2015,
Stockholm, Sweden, 15 September 2015.
143.
Kawakubo, M.; Tanaka, M.; Ochi, K.; Watanabe, A.; Saka-Tanaka, M.; Kanamori, Y.; Yoshioka, N.; Yamashita, S.; Goto, M.;
Itoh, M.; et al.
Dipeptidyl peptidase-4 inhibition prevents nonalcoholic steatohepatitis-associated liver fibrosis and tumor devel-
opment in mice independently of its anti-diabetic effects. Sci. Rep. 2020,10, 983. [CrossRef] [PubMed]
144.
Wang, G.; Wu, B.; Zhang, L.; Jin, X.; Wang, K.; Xu, W.; Zhang, B.; Wang, H. The protective effects of trelagliptin on high-fat
diet-induced nonalcoholic fatty liver disease in mice. J. Biochem. Mol. Toxicol. 2021,35, e22696. [CrossRef]
145.
Choi, S.H.; Leem, J.; Park, S.; Lee, C.K.; Park, K.G.; Lee, I.K. Gemigliptin ameliorates Western-diet-induced metabolic syndrome
in mice. Can. J. Physiol. Pharmacol. 2017,95, 129–139. [CrossRef]
146.
Klein, T.; Fujii, M.; Sandel, J.; Shibazaki, Y.; Wakamatsu, K.; Mark, M.; Yoneyama, H. Linagliptin alleviates hepatic steatosis and
inflammation in a mouse model of non-alcoholic steatohepatitis. Med. Mol. Morphol. 2014,47, 137–149. [CrossRef]
147. Saisho, Y. SGLT2 Inhibitors: The Star in the Treatment of Type 2 Diabetes? Diseases 2020,8, 14. [CrossRef]
148.
Wu, P.; Wen, W.; Li, J.; Xu, J.; Zhao, M.; Chen, H.; Sun, J. Systematic Review and Meta-Analysis of Randomized Controlled Trials
on the Effect of SGLT2 Inhibitor on Blood Leptin and Adiponectin Level in Patients with Type 2 Diabetes. Horm. Metab. Res.
2019
,
51, 487–494. [CrossRef]
149.
Sumida, Y.; Yoneda, M.; Tokushige, K.; Kawanaka, M.; Fujii, H.; Yoneda, M.; Imajo, K.; Takahashi, H.; Ono, M.; Nozaki, Y.;
et al. Hepatoprotective Effect of SGLT2 Inhibitor on Nonalcoholic Fatty Liver Disease. Diabetes Res. Open Access
2020
,2, 17–25.
[CrossRef]
150.
Ribeiro Dos Santos, L.; Baer Filho, R. Treatment of nonalcoholic fatty liver disease with dapagliflozin in non-diabetic patients.
Metabol. Open 2020,5, 100028. [CrossRef] [PubMed]
151.
Taheri, H.; Malek, M.; Ismail-Beigi, F.; Zamani, F.; Sohrabi, M.; Reza Babaei, M.; Khamseh, M.E. Effect of Empagliflozin on
Liver Steatosis and Fibrosis in Patients with Non-Alcoholic Fatty Liver Disease without Diabetes: A Randomized, Double-Blind,
Placebo-Controlled Trial. Adv. Ther. 2020,37, 4697–4708. [CrossRef] [PubMed]
152.
Itani, T.; Ishihara, T. Efficacy of canagliflozin against nonalcoholic fatty liver disease: A prospective cohort study. Obes. Sci. Pract.
2018,4, 477–482. [CrossRef] [PubMed]
153.
Prikhodko, V.A.; Sysoev, Y.I.; Poveryaeva, M.A.; Bunyat, A.V.; Karev, V.E.; Ivkin, D.Y.; Sukhanov, D.S.; Shustov, E.B.; Okovityi, S.V.
Effects of Empagliflozin and L-Ornithine L-Aspartate on Behavior, Cognitive Functions, and Physical Performance in Mice with
Experimentally Induced Steatohepatitis. Bull. RSMU 2020,3, 49–57. [CrossRef]
154.
Takahashi, H.; Kessoku, T.; Kawanaka, M.; Nonaka, M.; Hyogo, H.; Fujii, H.; Nakajima, T.; Imajo, K.; Tanaka, K.; Kubotsu, Y.; et al.
Ipragliflozin Improves the Hepatic Outcomes of Patients with Diabetes with NAFLD. Hepatol. Commun. 2021, online ahead of print.
[CrossRef]
155.
Wilkison, B.; Cheatham, B.; Walker, S. Remogliflozin Etabonate Reduces FIB-4 and NAFLD Fibrosis Scores in Type 2 Diabetic
Subjects. Hepatology 2016,64, 548A.
156.
Gallo, S.; Calle, R.A.; Terra, S.G.; Pong, A.; Tarasenko, L.; Raji, A. Effects of Ertugliflozin on Liver Enzymes in Patients with Type 2
Diabetes: A Post-Hoc Pooled Analysis of Phase 3 Trials. Diabetes Ther. 2020,11, 1849–1860. [CrossRef]
157.
Shibuya, T.; Fushimi, N.; Kawai, M.; Yoshida, Y.; Hachiya, H.; Ito, S.; Kawai, H.; Ohashi, N.; Mori, A. 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, 438–442. [CrossRef]
158.
Sumida, Y.; Murotani, K.; Saito, M.; Tamasawa, A.; Osonoi, Y.; Yoneda, M.; Osonoi, T. Effect of luseogliflozin on hepatic fat content
in type 2 diabetes patients with non-alcoholic fatty liver disease: A prospective, single-arm trial (LEAD trial). Hepatol. Res.
2019
,
49, 64–71. [CrossRef]
159.
Yoneda, M.; Honda, Y.; Ogawa, Y.; Kessoku, T.; Kobayashi, T.; Imajo, K.; Ozaki, A.; Nogami, A.; Taguri, M.; Yamanaka, T.; et al.
Comparing the effects of tofogliflozin and pioglitazone in non-alcoholic fatty liver disease patients with type 2 diabetes mellitus
(ToPiND study): A randomized prospective open-label controlled trial. BMJ Open Diabetes Res. Care
2021
,9, e001990. [CrossRef]
160.
He, Y.L.; Haynes, W.; Meyers, C.D.; Amer, A.; Zhang, Y.; Mahling, P.; Mendonza, A.E.; Ma, S.; Chutkow, W.; Bachman, E. The
effects of licogliflozin, a dual SGLT1/2 inhibitor, on body weight in obese patients with or without diabetes. Diabetes Obes. Metab.
2019,21, 1311–1321. [CrossRef] [PubMed]
161.
Harrison, S.A.; Manghi, F.P.; Smith, W.B.; Alpenidze, D.; Aizenberg, D.; Burggraaf, K.; Chen, C.-Y.; Zuckerman, E.; Ravussin, E.;
Charatcharoenwitthaya, P.; et al. LIK066 (Licogliflozin), an SGLT1/2 Inhibitor, Robustly Decreases ALT and Improves Markers of
Hepatic and Metabolic Health in Patients with Non-Alcoholic Fatty Liver Disease: Interim Analysis of a 12-Week, Randomized,
Placebo-Controlled, Phase 2a Study. In Proceedings of the Liver Meeting 2019, Boston, MA, USA, 8–12 November 2019.
162.
Honda, Y.; Ozaki, A.; Iwaki, M.; Kobayashi, T.; Nogami, A.; Kessoku, T.; Ogawa, Y.; Tomeno, W.; Imajo, K.; Yoneda, M.; et al.
Protective effect of SGL5213, a potent intestinal sodium-glucose cotransporter 1 inhibitor, in nonalcoholic fatty liver disease in
mice. J. Pharmacol. Sci. 2021,147, 176–183. [CrossRef] [PubMed]
163.
Iogna Prat, L.; Tsochatzis, E.A. The effect of antidiabetic medications on non-alcoholic fatty liver disease (NAFLD). Hormones
2018,17, 219–229. [CrossRef] [PubMed]
164.
Hajiaghamohammadi, A.A.; Miroliaee, A.; Samimi, R.; Alborzi, F.; Ziaee, A. A Comparison of Ezetimibe and Acarbose in
Decreasing Liver Transaminase in Nonalcoholic Fatty Liver Disease: A Randomized Clinical Trial. Govaresh 2013,18, 186–190.
Biomedicines 2022,10, 274 31 of 34
165. Komatsu, M.; Tanaka, N.; Kimura, T.; Fujimori, N.; Sano, K.; Horiuchi, A.; Sugiura, A.; Yamazaki, T.; Shibata, S.; Joshita, S.; et al.
Miglitol attenuates non-alcoholic steatohepatitis in diabetic patients. Hepatol. Res. 2018,48, 1092–1098. [CrossRef]
166.
Kim, J.W.; Lee, Y.J.; You, Y.H.; Moon, M.K.; Yoon, K.H.; Ahn, Y.B.; Ko, S.H. Effect of sodium-glucose cotransporter 2 inhibitor,
empagliflozin, and
α
-glucosidase inhibitor, voglibose, on hepatic steatosis in an animal model of type 2 diabetes. J. Cell. Biochem.
2018, online ahead of print. [CrossRef]
167.
Meroni, M.; Longo, M.; Dongiovanni, P. The Role of Probiotics in Nonalcoholic Fatty Liver Disease: A New Insight into
Therapeutic Strategies. Nutrients 2019,11, 2642. [CrossRef]
168.
Malaguarnera, M.; Vacante, M.; Antic, T.; Giordano, M.; Chisari, G.; Acquaviva, R.; Mastrojeni, S.; Malaguarnera, G.; Mistretta, A.;
Li Volti, G.; et al. Bifidobacterium longum with fructo-oligosaccharides in patients with non alcoholic steatohepatitis. Dig. Dis.
Sci. 2012,57, 545–553. [CrossRef]
169.
Vajro, P.; Mandato, C.; Licenziati, M.R.; Franzese, A.; Vitale, D.F.; Lenta, S.; Caropreso, M.; Vallone, G.; Meli, R. Effects of
Lactobacillus rhamnosus strain GG in pediatric obesity-related liver disease. J. Pediatr. Gastroenterol. Nutr.
2011
,52, 740–743.
[CrossRef]
170.
Nabavi, S.; Rafraf, M.; Somi, M.H.; Homayouni-Rad, A.; Asghari-Jafarabadi, M. Effects of probiotic yogurt consumption on
metabolic factors in individuals with nonalcoholic fatty liver disease. J. Dairy Sci. 2014,97, 7386–7393. [CrossRef]
171.
Román, E.; Nieto, J.C.; Gely, C.; Vidal, S.; Pozuelo, M.; Poca, M.; Juárez, C.; Guarner, C.; Manichanh, C.; Soriano, G. Effect of a
Multistrain Probiotic on Cognitive Function and Risk of Falls in Patients with Cirrhosis: A Randomized Trial. Hepatol. Commun.
2019,3, 632–645. [CrossRef]
172.
Ahn, S.B.; Jun, D.W.; Kang, B.K.; Lim, J.H.; Lim, S.; Chung, M.J. Randomized, Double-blind, Placebo-controlled Study of a
Multispecies Probiotic Mixture in Nonalcoholic Fatty Liver Disease. Sci. Rep. 2019,9, 5688. [CrossRef] [PubMed]
173.
Kobyliak, N.; Abenavoli, L.; Mykhalchyshyn, G.; Kononenko, L.; Boccuto, L.; Kyriienko, D.; Dynnyk, O. A Multi-strain Probiotic
Reduces the Fatty Liver Index, Cytokines and Aminotransferase levels in NAFLD Patients: Evidence from a Randomized Clinical
Trial. J. Gastrointestin Liver Dis. 2018,27, 41–49. [CrossRef] [PubMed]
174.
Ma, Y.Y.; Li, L.; Yu, C.H.; Shen, Z.; Chen, L.H.; Li, Y.M. Effects of probiotics on nonalcoholic fatty liver disease: A meta-analysis.
World J. Gastroenterol. 2013,19, 6911–6918. [CrossRef] [PubMed]
175.
Gao, X.; Zhu, Y.; Wen, Y.; Liu, G.; Wan, C. Efficacy of probiotics in non-alcoholic fatty liver disease in adult and children: A
meta-analysis of randomized controlled trials. Hepatol. Res. 2016,46, 1226–1233. [CrossRef]
176.
Liu, Y.T.; Li, Y.Q.; Wang, Y.Z. Protective effect of Saccharomyces boulardii against intestinal mucosal barrier injury in rats with
nonalcoholic fatty liver disease. Zhonghua Gan Zang Bing Za Zhi 2016,24, 921–926. (In Chinese) [CrossRef]
177.
Yamauchi, R.; Takedatsu, H.; Yokoyama, K.; Yamauchi, E.; Kawashima, M.; Tsuchiya, N.; Takata, K.; Tanaka, T.; Morihara, D.;
Takeyama, Y.; et al. Synergistic effect of clostridium butyricum miyairi on rifaximin in mice model of non-alcoholic steatohepatitis
by methionine choline-deficient diet. J. Hepatol. 2020,73, S401–S652. [CrossRef]
178.
Abdel Monem, S.M. Probiotic Therapy in Patients with Nonalcoholic Steatohepatitis in Zagazig University Hospitals. Euroasian J.
Hepatogastroenterol. 2017,7, 101–106. [CrossRef]
179.
Li, C.; Nie, S.P.; Zhu, K.X.; Ding, Q.; Li, C.; Xiong, T.; Xie, M.Y. Lactobacillus plantarum NCU116 improves liver function, oxidative
stress and lipid metabolism in rats with high fat diet induced non-alcoholic fatty liver disease. Food Funct.
2014
,5, 3216–3223.
[CrossRef]
180.
Yao, F.; Jia, R.; Huang, H.; Yu, Y.; Mei, L.; Bai, L.; Ding, Y.; Zheng, P. Effect of Lactobacillus paracasei N1115 and fructooligosaccha-
rides in nonalcoholic fatty liver disease. Arch. Med. Sci. 2019,15, 1336–1344. [CrossRef]
181.
Xin, J.; Zeng, D.; Wang, H.; Ni, X.; Yi, D.; Pan, K.; Jing, B. Preventing non-alcoholic fatty liver disease through Lactobacillus
johnsonii BS15 by attenuating inflammation and mitochondrial injury and improving gut environment in obese mice. Appl.
Microbiol. Biotechnol. 2014,98, 6817–6829. [CrossRef]
182.
Ting, W.J.; Kuo, W.W.; Hsieh, D.J.; Yeh, Y.L.; Day, C.H.; Chen, Y.H.; Chen, R.J.; Padma, V.V.; Chen, Y.H.; Huang, C.Y. Heat
Killed Lactobacillus reuteri GMNL-263 Reduces Fibrosis Effects on the Liver and Heart in High Fat Diet-Hamsters via TGF-
β
Suppression. Int. J. Mol. Sci. 2015,16, 25881–25896. [CrossRef] [PubMed]
183.
Aller, R.; De Luis, D.A.; Izaola, O.; Conde, R.; Gonzalez Sagrado, M.; Primo, D.; De La Fuente, B.; Gonzalez, J. Effect of a probiotic
on liver aminotransferases in nonalcoholic fatty liver disease patients: A double blind randomized clinical trial. Eur. Rev. Med.
Pharmacol. Sci. 2011,15, 1090–1095. [PubMed]
184.
Famouri, F.; Shariat, Z.; Hashemipour, M.; Keikha, M.; Kelishadi, R. Effects of Probiotics on Nonalcoholic Fatty Liver Disease in
Obese Children and Adolescents. J. Pediatr. Gastroenterol. Nutr. 2017,64, 413–417. [CrossRef] [PubMed]
185.
Wong, V.W.; Won, G.L.; Chim, A.M.; Chu, W.C.; Yeung, D.K.; Li, K.C.; Chan, H.L. Treatment of nonalcoholic steatohepatitis with
probiotics. A proof-of-concept study. Ann. Hepatol. 2013,12, 256–262. [CrossRef]
186.
Alisi, A.; Bedogni, G.; Baviera, G.; Giorgio, V.; Porro, E.; Paris, C.; Giammaria, P.; Reali, L.; Anania, F.; Nobili, V. Randomised
clinical trial: The beneficial effects of VSL#3 in obese children with non-alcoholic steatohepatitis. Aliment. Pharmacol. Ther.
2014
,
39, 1276–1285. [CrossRef]
187.
Ezquer, M.; Ezquer, F.; Ricca, M.; Allers, C.; Conget, P. Intravenous administration of multipotent stromal cells prevents the onset
of non-alcoholic steatohepatitis in obese mice with metabolic syndrome. J. Hepatol. 2011,55, 1112–1120. [CrossRef]
Biomedicines 2022,10, 274 32 of 34
188.
Hsu, M.J.; Karkossa, I.; Schäfer, I.; Christ, M.; Kühne, H.; Schubert, K.; Rolle-Kampczyk, U.E.; Kalkhof, S.; Nickel, S.;
Seibel, P.; et al.
Mitochondrial Transfer by Human Mesenchymal Stromal Cells Ameliorates Hepatocyte Lipid Load in a Mouse Model of NASH.
Biomedicines 2020,8, 350. [CrossRef]
189. Wang, H.; Wang, D.; Yang, L.; Wang, Y.; Jia, J.; Na, D.; Chen, H.; Luo, Y.; Liu, C. Compact bone-derived mesenchymal stem cells
attenuate nonalcoholic steatohepatitis in a mouse model by modulation of CD4 cells differentiation. Int. Immunopharmacol.
2017
,
42, 67–73. [CrossRef]
190.
Chien, Y.; Huang, C.S.; Lin, H.C.; Lu, K.H.; Tsai, P.H.; Lai, Y.H.; Chen, K.H.; Lee, S.D.; Huang, Y.H.; Wang, C.Y. Improvement
of non-alcoholic steatohepatitis by hepatocyte-like cells generated from iPSCs with Oct4/Sox2/Klf4/Parp1. Oncotarget
2018
,9,
18594–18606. [CrossRef]
191.
Overi, D.; Carpino, G.; Franchitto, A.; Onori, P.; Gaudio, E. Hepatocyte Injury and Hepatic Stem Cell Niche in the Progression of
Non-Alcoholic Steatohepatitis. Cells 2020,9, 590. [CrossRef]
192.
Nevens, F.; Gustot, T.; Laterre, P.F.; Lasser, L.L.; Haralampiev, L.E.; Vargas, V.; Lyubomirova, D.; Albillos, A.; Najimi, M.;
Michel, S.; et al.
A phase II study of human allogeneic liver-derived progenitor cell therapy for acute-on-chronic liver failure and
acute decompensation. JHEP Rep. 2021,3, 100291. [CrossRef] [PubMed]
193.
Iqbal, U.; Perumpail, B.J.; John, N.; Sallam, S.; Shah, N.D.; Kwong, W.; Cholankeril, G.; Kim, D.; Ahmed, A. Judicious Use of Lipid
Lowering Agents in the Management of NAFLD. Diseases 2018,6, 87. [CrossRef] [PubMed]
194.
Oniciu, D.C.; Hashiguchi, T.; Shibazaki, Y.; Bisgaier, C.L. Gemcabene downregulates inflammatory, lipid-altering and cell-signaling
genes in the STAM™ model of NASH. PLoS ONE 2018,13, e0194568. [CrossRef] [PubMed]
195.
Sanjay, K.V.; Vishwakarma, S.; Zope, B.R.; Mane, V.S.; Mohire, S.; Dhakshinamoorthy, S. ATP citrate lyase inhibitor Bempedoic
Acid alleviate long term HFD induced NASH through improvement in glycemic control, reduction of hepatic triglycerides &
total cholesterol, modulation of inflammatory & fibrotic genes and improvement in NAS score. Curr. Res. Pharmacol. Drug Discov.
2021,2, 100051. [CrossRef] [PubMed]
196.
Stanley, T.L.; Feldpausch, M.N.; Oh, J.; Branch, K.L.; Lee, H.; Torriani, M.; Grinspoon, S.K. Effect of tesamorelin on visceral fat
and liver fat in HIV-infected patients with abdominal fat accumulation: A randomized clinical trial. JAMA
2014
,312, 380–389.
[CrossRef]
197. Theratechnologies Inc. Available online: https://www.theratech.com/ (accessed on 10 December 2021).
198.
Wei, X.; Wang, C.; Hao, S.; Song, H.; Yang, L. The Therapeutic Effect of Berberine in the Treatment of Nonalcoholic Fatty Liver
Disease: A Meta-Analysis. Evid.-Based Complementary Altern. Med. 2016,2016, 3593951. [CrossRef]
199. Yin, J.; Ye, J.; Jia, W. Effects and mechanisms of berberine in diabetes treatment. Acta Pharm. Sin. B 2012,2, 327–334. [CrossRef]
200.
Simental-Mendía, M.; Sánchez-García, A.; Simental-Mendía, L.E. Effect of ursodeoxycholic acid on liver markers: A systematic
review and meta-analysis of randomized placebo-controlled clinical trials. Br. J. Clin. Pharmacol.
2020
,86, 1476–1488. [CrossRef]
201.
Zhang, W.; Tang, Y.; Huang, J.; Hu, H. Efficacy of ursodeoxycholic acid in nonalcoholic fatty liver disease: An updated
meta-analysis of randomized controlled trials. Asia Pac. J. Clin. Nutr. 2020,29, 696–705. [CrossRef]
202.
Harrison, S.A.; Gunn, N.; Neff, G.W.; Kohli, A.; Liu, L.; Flyer, A.; Goldkind, L.; Di Bisceglie, A.M. A phase 2, proof of concept,
randomised controlled trial of berberine ursodeoxycholate in patients with presumed non-alcoholic steatohepatitis and type 2
diabetes. Nat. Commun. 2021,12, 5503. [CrossRef]
203.
Kowdley, K.V.; Butler, P.; Cubberley, S.; Hand, A.L.; Jenders, R.A.; Kroon, J.; Leibowitz, M.; Moore, A.C.; Guyer, B. Miricorilant, a
Selective GR Modulator, Induced a Rapid and Significant Reduction in Liver Fat Content in a Randomized, Placebo-Controlled
Phase 2a Study in Patients with Non-Alcoholic Steatohepatitis. In Proceedings of the Liver Meeting 2021, Virtual, 12–15
November 2021.
204.
Li, F.; Jiang, M.; Ma, M.; Chen, X.; Zhang, Y.; Zhang, Y.; Yu, Y.; Cui, Y.; Chen, J.; Zhao, H.; et al. Anthelmintics nitazoxanide
protects against experimental hyperlipidemia and hepatic steatosis in hamsters and mice. Acta Pharm. Sin. B 2021, in press,
corrected proof. [CrossRef]
205.
Walczak, R.; Carole, B.; Benoit, N.; Descamps, E.; Nathalie, D.; Meghien, S.; Hum, D.; Staels, B.; Friedman, S.; Loomba, R.; et al.
Elafibranor and nitazoxanide synergize to reduce fibrosis in a NASH model. J. Hepatol. 2018,68, S105–S364. [CrossRef]
206.
Glal, K.A.M.; Abd-Elsalam, S.M.; Mostafa, T.M. Nitazoxanide versus rifaximin in preventing the recurrence of hepatic en-
cephalopathy: A randomized double-blind controlled trial. J. Hepato-Biliary-Pancreat. Sci.
2021
,28, 812–824. [CrossRef] [PubMed]
207.
Flores-Contreras, L.; Sandoval-Rodríguez, A.S.; Mena-Enriquez, M.G.; Lucano-Landeros, S.; Arellano-Olivera, I.;
Alvarez-Álvarez, A.
;
Sanchez-Parada, M.G.; Armendáriz-Borunda, J. Treatment with pirfenidone for two years decreases fibrosis, cytokine levels and
enhances CB2 gene expression in patients with chronic hepatitis C. BMC Gastroenterol. 2014,14, 131. [CrossRef] [PubMed]
208.
Cui, Y.; Zhang, M.; Leng, C.; Blokzijl, T.; Jansen, B.H.; Dijkstra, G.; Faber, K.N. Pirfenidone Inhibits Cell Proliferation and Collagen
I Production of Primary Human Intestinal Fibroblasts. Cells 2020,9, 775. [CrossRef] [PubMed]
209.
Poo, J.L.; Torre, A.; Aguilar-Ramírez, J.R.; Cruz, M.; Mejía-Cuán, L.; Cerda, E.; Velázquez, A.; Patiño, A.; Ramírez-Castillo, C.;
Cisneros, L.; et al. Benefits of prolonged-release pirfenidone plus standard of care treatment in patients with advanced liver
fibrosis: PROMETEO study. Hepatol. Int. 2020,14, 817–827. [CrossRef]
210.
Matsuda, K.; Iwaki, Y. MN-001 (tipelukast), a novel, orally bioavailable drug, reduces fibrosis and inflammation and down-
regulates TIMP-1, collagen Type 1 and LOXL2 mRNA overexpression in an advanced NASH (non-alcoholic steatohepatitis)
model. In Proceedings of the Liver Meeting 2014, Boston, MA, USA, 7–11 November 2014.
Biomedicines 2022,10, 274 33 of 34
211.
Yang, M.; Zhang, C.Y. G protein-coupled receptors as potential targets for nonalcoholic fatty liver disease treatment. World J.
Gastroenterol. 2021,27, 677–691. [CrossRef]
212.
Widjaja, A.A.; Singh, B.K.; Adami, E.; Viswanathan, S.; Dong, J.; D’Agostino, G.A.; Ng, B.; Lim, W.W.; Tan, J.; Paleja, B.S.; et al.
Inhibiting Interleukin 11 Signaling Reduces Hepatocyte Death and Liver Fibrosis, Inflammation, and Steatosis in Mouse Models
of Nonalcoholic Steatohepatitis. Gastroenterology 2019,157, 777–792.e14. [CrossRef]
213.
Zai, W.; Chen, W.; Liu, H.; Ju, D. Therapeutic Opportunities of IL-22 in Non-Alcoholic Fatty Liver Disease: From Molecular
Mechanisms to Clinical Applications. Biomedicines 2021,9, 1912. [CrossRef]
214.
Stemmer, S.M.; Manojlovic, N.S.; Marinca, M.V.; Petrov, P.; Cherciu, N.; Ganea, D.; Ciuleanu, T.E.; Pusca, I.A.; Beg, M.S.; Purcell,
W.T.; et al. Namodenoson in Advanced Hepatocellular Carcinoma and Child-Pugh B Cirrhosis: Randomized Placebo-Controlled
Clinical Trial. Cancers 2021,13, 187. [CrossRef]
215.
Ferramosca, A.; Di Giacomo, M.; Zara, V. Antioxidant dietary approach in treatment of fatty liver: New insights and updates.
World J. Gastroenterol. 2017,23, 4146–4157. [CrossRef]
216. Ionis Pharmaceuticals, Inc. Available online: https://www.ionispharma.com/ (accessed on 10 December 2021).
217.
Lawitz, E.; Hassanein, T.; Denham, D.; Waters, M.; Borg, B.; Mille, G.; Scott, D.; Miksztal, A.; Culwell, J.; Ellis, D.; et al. Efficacy
Signals of 4-Week Oral DUR-928 in NASH Subjects. In Proceedings of the Digital International Liver Congress 2021, Virtual,
23–26 June 2021.
218.
Konuma, K.; Itoh, M.; Suganami, T.; Kanai, S.; Nakagawa, N.; Sakai, T.; Kawano, H.; Hara, M.; Kojima, S.; Izumi, Y.; et al.
Eicosapentaenoic acid ameliorates non-alcoholic steatohepatitis in a novel mouse model using melanocortin 4 receptor-deficient
mice. PLoS ONE 2015,10, e0121528. [CrossRef] [PubMed]
219.
Sojoodi, M.; Wang, Y.; Erstad, D.J.; Caravan, P.; Lanuti, M.; Qadan, M.; Fuchs, B.C.; Hoang, K.; Or, Y.S.; Jiang, L.; et al. EDP-297,
a novel and potent fxr agonist, exhibit robust antifibrotic effects with significant liver function in a rat model of non-alcoholic
steatohepatitis. J. Hepatol. 2020,73, S401–S652. [CrossRef]
220.
Ma, Y.; Huang, Y.; Yan, L.; Gao, M.; Liu, D. Synthetic FXR agonist GW4064 prevents diet-induced hepatic steatosis and insulin
resistance. Pharm. Res. 2013,30, 1447–1457. [CrossRef] [PubMed]
221.
Fraile, J.M.; Palliyil, S.; Barelle, C.; Porter, A.J.; Kovaleva, M. Non-Alcoholic Steatohepatitis (NASH)—A Review of a Crowded
Clinical Landscape, Driven by a Complex Disease. Drug Des. Dev. Ther. 2021,15, 3997–4009. [CrossRef]
222.
Aithal, G.P.; Thomas, J.A.; Kaye, P.V.; Lawson, A.; Ryder, S.D.; Spendlove, I.; Austin, A.S.; Freeman, J.G.; Morgan, L.; Webber, J.
Randomized, placebo-controlled trial of pioglitazone in nondiabetic subjects with nonalcoholic steatohepatitis. Gastroenterology
2008,135, 1176–1184. [CrossRef]
223.
Altimmune, Inc. Pemvidutide (ALT-801): Phase 1 12-Week Results. Available online: https://ir.altimmune.com/static-files/d331
9658-9809-44ad-a729-05dd6a6bf4c5 (accessed on 10 December 2021).
224.
Tobita, H.; Sato, S.; Miyake, T.; Ishihara, S.; Kinoshita, Y. Effects of Dapagliflozin on Body Composition and Liver Tests in Patients
with Nonalcoholic Steatohepatitis Associated with Type 2 Diabetes Mellitus: A Prospective, Open-label, Uncontrolled Study.
Curr. Ther. Res. Clin. Exp. 2017,87, 13–19. [CrossRef]
225.
Lai, L.L.; Vethakkan, S.R.; Nik Mustapha, N.R.; Mahadeva, S.; Chan, W.K. Empagliflozin for the Treatment of Nonalcoholic
Steatohepatitis in Patients with Type 2 Diabetes Mellitus. Dig. Dis. Sci. 2020,65, 623–631. [CrossRef]
226.
Seko, Y.; Sumida, Y.; Tanaka, S.; Mori, K.; Taketani, H.; Ishiba, H.; Hara, T.; Okajima, A.; Umemura, A.; Nishikawa, T.; et al. Effect
of sodium glucose cotransporter 2 inhibitor on liver function tests in Japanese patients with non-alcoholic fatty liver disease and
type 2 diabetes mellitus. Hepatol. Res. 2017,47, 1072–1078. [CrossRef]
227.
Cusi, K.; Alkhouri, N.; Harrison, S.A.; Fouqueray, P.; Moller, D.E.; Hallakou-Bozec, S.; Bolze, S.; Grouin, J.M.; Jeannin
Megnien, S.
;
Dubourg, J.; et al. Efficacy and safety of PXL770, a direct AMP kinase activator, for the treatment of non-alcoholic fatty liver
disease (STAMP-NAFLD): A randomised, double-blind, placebo-controlled, phase 2a study. Lancet Gastroenterol. Hepatol.
2021
,6,
889–902. [CrossRef]
228.
Chalasani, N.; Vuppalanchi, R.; Rinella, M.; Middleton, M.S.; Siddiqui, M.S.; Barritt, A.S., IV; Kolterman, O.; Flores, O.; Alonso, C.;
Iruarrizaga-Lejarreta, M.; et al. Randomised clinical trial: A leucine-metformin-sildenafil combination (NS-0200) vs placebo in
patients with non-alcoholic fatty liver disease. Aliment. Pharmacol. Ther. 2018,47, 1639–1651. [CrossRef]
229.
Simard, J.C.; Thibodeau, J.F.; Leduc, M.; Tremblay, M.; Laverdure, A.; Sarra-Bournet, F.; Gagnon, W.; Ouboudinar, J.; Gervais, L.;
Felton, A.; et al. Fatty acid mimetic PBI-4547 restores metabolic homeostasis via GPR84 in mice with non-alcoholic fatty liver
disease. Sci Rep 2020,10, 12778. [CrossRef] [PubMed]
230.
Puengel, T.; De Vos, S.; Hundertmark, J.; Kohlhepp, M.; Guldiken, N.; Pujuguet, P.; Auberval, M.; Marsais, F.; Shoji, K.F.;
Saniere, L.; et al.
The Medium-Chain Fatty Acid Receptor GPR84 Mediates Myeloid Cell Infiltration Promoting Steatohepatitis
and Fibrosis. J. Clin. Med. 2020,9, 1140. [CrossRef] [PubMed]
231.
Chen, X.; Liu, C.; Ruan, L. G-Protein-Coupled Receptors 120 Agonist III Improves Hepatic Inflammation and ER Stress in
Steatohepatitis. Dig. Dis. Sci. 2021,66, 1090–1096. [CrossRef]
232.
Teng, B.; Huang, C.; Cheng, C.L.; Udduttula, A.; Yu, X.F.; Liu, C.; Li, J.; Yao, Z.Y.; Long, J.; Miao, L.F.; et al. Newly identified
peptide hormone inhibits intestinal fat absorption and improves NAFLD through its receptor GPRC6A. J. Hepatol.
2020
,73,
383–393. [CrossRef] [PubMed]
Biomedicines 2022,10, 274 34 of 34
233.
Ookawara, M.; Matsuda, K.; Watanabe, M.; Moritoh, Y. The GPR40 Full Agonist SCO-267 Improves Liver Parameters in a Mouse
Model of Nonalcoholic Fatty Liver Disease without Affecting Glucose or Body Weight. J. Pharmacol. Exp. Ther.
2020
,375, 21–27.
[CrossRef] [PubMed]
234.
Shafiq, M.; Walmann, T.; Nutalapati, V.; Gibson, C.; Zafar, Y. Effects of proprotein convertase subtilisin/kexin type-9 inhibitors on
fatty liver. World J. Hepatol. 2020,12, 1258–1266. [CrossRef]
235.
Jun, B.G.; Cheon, G.J. The utility of ezetimibe therapy in nonalcoholic fatty liver disease. Korean J. Intern. Med.
2019
,34, 284–285.
[CrossRef]
236.
Kidron, M.; Perles, S.; Kaloti, R.; Ghantous, R.; Sandouka, S.F.; Malaabi, Y.; Safadi, R. Oral Insulin–Induced Reduction in Liver Fat
Content in T2DM Patients with Nonalcoholic Steatohepatitis. Diabetes 2020,69, 115-LB. [CrossRef]
237.
Cansby, E.; Nuñez-Durán, E.; Magnusson, E.; Amrutkar, M.; Booten, S.L.; Kulkarni, N.M.; Svensson, L.T.; Borén, J.; Marschall, H.U.;
Aghajan, M.; et al. Targeted Delivery of Stk25 Antisense Oligonucleotides to Hepatocytes Protects Mice Against Nonalcoholic
Fatty Liver Disease. Cell. Mol. Gastroenterol. Hepatol. 2019,7, 597–618. [CrossRef]
238.
Segal-Salto, M.; Barashi, N.; Katav, A.; Edelshtein, V.; Aharon, A.; Hashmueli, S.; George, J.; Maor, Y.; Pinzani, M.;
Haberman, D.; et al.
A blocking monoclonal antibody to CCL24 alleviates liver fibrosis and inflammation in experimental models
of liver damage. JHEP Rep. 2020,2, 100064. [CrossRef]
239.
Safadi, R.; Braun, M.; Francis, A.; Milgrom, Y.; Massarwa, M.; Hakimian, D.; Hazou, W.; Issachar, A.; Harpaz, Z.;
Farbstein, M.; et al.
Randomised clinical trial: A phase 2 double-blind study of namodenoson in non-alcoholic fatty liver
disease and steatohepatitis. Aliment. Pharmacol. Ther. 2021,54, 1405–1415. [CrossRef] [PubMed]
240. Pliant Therapeutics. Available online: https://pliantrx.com/ (accessed on 10 December 2021).
241. CohBar, Inc. Available online: https://www.cohbar.com/ (accessed on 10 December 2021).
242.
Cerenis Therapeutics. Results of the Phase I Study of Repeated and Increasing Doses to Assess CER-209 in NASH/NAFLD: Press Release;
2018. Available online: https://www.abionyx.com/images/pdfs/images/pr_cerenis_CER_209_ENG_final_e06d4.pdf (accessed
on 10 December 2021).
243. Hepion Pharmaceuticals. Available online: https://hepionpharma.com/ (accessed on 10 December 2021).
244. Lipocine. Available online: https://www.lipocine.com/ (accessed on 10 December 2021).
245.
Gupte, A.A.; Sabek, O.M.; Fraga, D.; Minze, L.J.; Nishimoto, S.K.; Liu, J.Z.; Afshar, S.; Gaber, L.; Lyon, C.J.; Gaber, O.; et al.
Osteocalcin protects against non-alcoholic steatohepatitis in a mouse model of metabolic syndrome. Endocrinology
2020
,155,
4697–4705. [CrossRef]
246.
Sookoian, S.; Pirola, C.J.; Valenti, L.; Davidson, N.O. Genetic Pathways in Nonalcoholic Fatty Liver Disease: Insights from Systems
Biology. Hepatology 2020,72, 330–346. [CrossRef] [PubMed]
247.
Kumar, V.; Xin, X.; Ma, J.; Tan, C.; Osna, N.; Mahato, R.I. Therapeutic targets, novel drugs, and delivery systems for diabetes
associated NAFLD and liver fibrosis. Adv. Drug Deliv. Rev. 2021,176, 113888. [CrossRef] [PubMed]
248.
Fang, Z.; Dou, G.; Wang, L. MicroRNAs in the Pathogenesis of Nonalcoholic Fatty Liver Disease. Int. J. Biol. Sci.
2021
,17,
1851–1863. [CrossRef]
249.
Wu, P.; Zhang, M.; Webster, N.J.G. Alternative RNA Splicing in Fatty Liver Disease. Front. Endocrinol.
2021
,12, 613213. [CrossRef]
250. Bosch-Presegué, L.; Vaquero, A. Sirtuins in stress response: Guardians of the genome. Oncogene 2014,33, 3764–3775. [CrossRef]
251.
Vachharajani, V.; McCall, C.E. Sirtuins: Potential therapeutic targets for regulating acute inflammatory response? Expert Opin.
Ther. Targets 2020,24, 489–497. [CrossRef]
252.
Zhu, Y.; Yan, Y.; Gius, D.R.; Vassilopoulos, A. Metabolic regulation of Sirtuins upon fasting and the implication for cancer. Curr.
Opin. Oncol. 2013,25, 630–636. [CrossRef]
253.
Colak, Y.; Ozturk, O.; Senates, E.; Tuncer, I.; Yorulmaz, E.; Adali, G.; Doganay, L.; Enc, F.Y. SIRT1 as a potential therapeutic target
for treatment of nonalcoholic fatty liver disease. Med. Sci. Monit. 2011,17, HY5–HY9. [CrossRef]
254.
Luci, C.; Bourinet, M.; Leclère, P.S.; Anty, R.; Gual, P. Chronic Inflammation in Non-Alcoholic Steatohepatitis: Molecular
Mechanisms and Therapeutic Strategies. Front. Endocrinol. 2020,11, 597648. [CrossRef] [PubMed]
255.
de Gregorio, E.; Colell, A.; Morales, A.; Marí, M. Relevance of SIRT1-NF-
κ
B Axis as Therapeutic Target to Ameliorate Inflammation
in Liver Disease. Int. J. Mol. Sci. 2020,21, 3858. [CrossRef] [PubMed]
256.
Milne, J.C.; Lambert, P.D.; Schenk, S.; Carney, D.P.; Smith, J.J.; Gagne, D.J.; Jin, L.; Boss, O.; Perni, R.B.; Vu, C.B.; et al. Small
molecule activators of SIRT1 as therapeutics for the treatment of type 2 diabetes. Nature
2007
,450, 712–716. [CrossRef] [PubMed]
257. Nassir, F.; Ibdah, J.A. Sirtuins and nonalcoholic fatty liver disease. World J. Gastroenterol. 2016,22, 10084–10092. [CrossRef]
258.
Kundu, A.; Dey, P.; Park, J.H.; Kim, I.S.; Kwack, S.J.; Kim, H.S. EX-527 Prevents the Progression of High-Fat Diet-Induced Hepatic
Steatosis and Fibrosis by Upregulating SIRT4 in Zucker Rats. Cells 2020,9, 1101. [CrossRef]
... Currently, there is no approved drug therapy by either the European Medicines Agency or the United States Food and Drug Administration (FDA) for MASLD, highlighting the urgent requirement for the creation of effective treatments for this widespread health issue [140]. However, a broad category available for the treatment management of MASLD includes antioxidants like vitamins C and E, insulin-sensitizing agents to improve the indirect causes of MASLD like DM, and hepatoprotective agents such as thioglita- ...
... Currently, there is no approved drug therapy by either the European Medicines Agency or the United States Food and Drug Administration (FDA) for MASLD, highlighting the urgent requirement for the creation of effective treatments for this widespread health issue [140]. However, a broad category available for the treatment management of MASLD includes antioxidants like vitamins C and E, insulin-sensitizing agents to improve the indirect causes of MASLD like DM, and hepatoprotective agents such as thioglitazones, Ursodeoxycholic acid, statins, pentoxifylline, and Orlistat [58]. ...
... Many pharmacological targets exist and are marketed in the drug development pipeline for future approval to treat MASLD: THRβ, lipogenesis inhibitors, ACC inhibitors, bile acid metabolism modulators, fibrogenesis inhibitors, glucose metabolism modulators, mesenchymal stromal cells, fraudulent fatty acids like bempedoic acid and gemcabene, tesamorelin for growth hormone modulation, berberine ursodeoxycholate for AMPK activation, miricorilant as a glucocorticoid receptor modulator, nitazoxanide with AMPK activation and HSC inhibition, pirfenidone with antifibrotic and anti-inflammatory properties, and other agents such a gut microbiome. These diverse strategies showcase the evolving landscape of potential treatments for different stages of MASLD [140]. ...
Article
Full-text available
Non-alcoholic fatty liver disease (NAFLD) is a widespread contributor to chronic liver disease globally. A recent consensus on renaming liver disease was established, and metabolic dysfunction-associated steatotic liver disease, MASLD, was chosen as the replacement for NAFLD. The disease’s range extends from the less severe MASLD, previously known as non-alcoholic fatty liver (NAFL), to the more intense metabolic dysfunction-associated steatohepatitis (MASH), previously known as non-alcoholic steatohepatitis (NASH), characterized by inflammation and apoptosis. This research project endeavors to comprehensively synthesize the most recent studies on MASLD, encompassing a wide spectrum of topics such as pathophysiology, risk factors, dietary influences, lifestyle management, genetics, epigenetics, therapeutic approaches, and the prospective trajectory of MASLD, particularly exploring its connection with organoids.
... 20 FXR, also known as a bile acid receptor, is a key nuclear receptor and activated by bile acids and expressed at high levels in the liver and the terminal ileum. 21 There are two generations of FXR agonists. The first generation is obeticholic acid (OCA) and the second generation is cilofexor and tropifexor. ...
... 30 4. Thyroid hormone receptor (THR) THR-β receptor is a nuclear receptor and a transcription factor that mediates the genomic effects of thyroid hormones. 21 Thyroid hormone receptor β (THR-β) had been shown to have effect in reduction of triglycerides and cholesterol, improving insulin sensitivity, reducing apoptosis and promoting live regeneration. 31 Resmetirom is an oral, liver-targeted, selective thyroid hormone receptor β (THR-β) agonists. ...
Article
Full-text available
Metabolic dysfunction associated steatotic liver disease (MASLD) or previously known Non-alcoholic fatty liver disease (NAFLD) is a common condition with an estimated global prevalence of around 30%. It is becoming a public health concern due to its close association with type 2 diabetes mellitus and obesity. It is important to screen for those with inflammation and fibrosis to halt the progression to cirrhosis. Cirrhosis is associated with liver related complications and liver cancer. Currently, there are no targeted treatments for MASLD at this stage and most treatments are currently in clinical trials. The focus of treatment had been on managing underlying risk metabolic risk factors. The purpose of this review to inform the readers of the change in the nomenclature from NAFLD to MASLD. This review will also focus on the background of MASLD, the pathogenesis as well as assessment and treatment of patients with MASLD.
... Несмотря на чрезвычайно высокую распространенность и большое медико-социальное значение, эффективные методы фармакотерапии как самой НАЖБП, так и ее неврологических осложнений до сих пор не найдены. Современные подходы к лечению НАЖБП включают применение средств с прямым гепатопротекторным действием, а также некоторых противодиабетических препаратовметформина, пиоглитазона, инкретиномиметиков и ингибиторов натрий-глюкозных котранспортеров (sodium/glucose cotransporter, SGLT), или глифлозинов [12]. Ввиду мультимодальности фармакологической активности последней группы средств она представляет особенный интерес с точки зрения применимости для коррекции неврологических расстройств у больных НАЖБП. ...
Article
Неалкогольная жировая болезнь печени (НАЖБП) и неалкогольный стеатогепатит (НАСГ) имеют ряд общих с поражениями центральной нервной системы факторов риска, а также выступают самостоятельной причиной развития цереброваскулярных, нейродегенеративных, когнитивных и психических расстройств. Разработка лекарственных средств, применимых не только для лечения самой НАЖБП, но и коррекции ее психоневрологических осложнений - актуальная задача современной экспериментальной биомедицины и фармакологии. Перспективной группой соединений, показавшей высокий терапевтический потенциал у больных НАЖБП, а также обладающей широким спектром плейотропных эффектов, являются ингибиторы натрий-глюкозных котранспортеров (глифлозины). В обзоре рассмотрены механизмы нейропротекторного действия глифлозинов, представляющие наибольший интерес в свете возможности коррекции неврологических осложнений НАЖБП. Non-alcoholic fatty liver disease (NAFLD) and non-alcoholic steatohepatitis (NASH) have a number of risk factors which are common to those of central nervous system lesions, and also act as an independent cause of the development of cerebrovascular, neurodegenerative, cognitive and mental disorders. The development of drugs applicable not only for the treatment of NAFLD itself, but also for the correction of its psychoneurological complications is an urgent task of modern experimental biomedicine and pharmacology. A promising group of compounds that have shown high therapeutic potential in patients with NAFLD, as well as having a wide range of pleiotropic effects, are sodium-glucose cotransporter inhibitors (gliflozins). Current review represents data concerning the mechanisms of gliflozins' neuroprotective efficacy, which are of greatest interest in light of the possibility of correcting the neurological complications of NAFLD.
... Sirt1 promotes mitochondrial biogenesis through peroxisome proliferator-activated receptor coactivator 1-alpha (PGC1a) (30,31). NR supplementation activates Sirt1 signaling (32), and activators of Sirt1 are promissory candidates for NAFLD treatment (33). The goal of the present study is to determine if the administration of Nic and HFCS in the oral consumption of Coca-Cola ™ (Coke) with HFCS can cause hepatic steatosis and that can be protected by NR. ...
Article
Full-text available
Introduction Cigarettes containing nicotine (Nic) are a risk factor for the development of cardiovascular and metabolic diseases. We reported that Nic delivered via injections or e-cigarette vapor led to hepatic steatosis in mice fed with a high-fat diet. High-fructose corn syrup (HFCS) is the main sweetener in sugar-sweetened beverages (SSBs) in the US. Increased consumption of SSBs with HFCS is associated with increased risks of non-alcoholic fatty liver disease (NAFLD). Nicotinamide riboside (NR) increases mitochondrial nicotinamide adenine dinucleotide (NAD⁺) and protects mice against hepatic steatosis. This study evaluated if Nic plus Coca-Cola™ (Coke) with HFCS can cause hepatic steatosis and that can be protected by NR. Methods C57BL/6J mice received twice daily intraperitoneal (IP) injections of Nic or saline and were given Coke (HFCS), or Coke with sugar, and NR supplementation for 10 weeks. Results Our results show that Nic+Coke caused increased caloric intake and induced hepatic steatosis, and the addition of NR prevented these changes. Western blot analysis showed lipogenesis markers were activated (increased cleavage of the sterol regulatory element-binding protein 1 [SREBP1c] and reduction of phospho-Acetyl-CoA Carboxylase [p-ACC]) in the Nic+Coke compared to the Sal+Water group. The hepatic detrimental effects of Nic+Coke were mediated by decreased NAD⁺ signaling, increased oxidative stress, and mitochondrial damage. NR reduced oxidative stress and prevented mitochondrial damage by restoring protein levels of Sirtuin1 (Sirt1) and peroxisome proliferator-activated receptor coactivator 1-alpha (PGC1) signaling. Conclusion We conclude that Nic+Coke has an additive effect on producing hepatic steatosis, and NR is protective. This study suggests concern for the development of NAFLD in subjects who consume nicotine and drink SSBs with HFCS.
... On the other hand, multiple drugs have been tested for their potential anti-inflammatory, cytoprotective, and antifibrotic effects in NAFLD patients [3,173,174]. Unfortunately, a great number of them have failed to demonstrate histologic improvement in fibrosis in clinical trials, but a few have shown some promising benefits [174][175][176]. However, there is no drug approved so far for NASH-associated fibrosis in patients with or without T2DM. ...
Article
Full-text available
The bidirectional relationship between type 2 diabetes and (non-alcoholic fatty liver disease) NAFLD is indicated by the higher prevalence and worse disease course of one condition in the presence of the other, but also by apparent beneficial effects observed in one, when the other is improved. This is partly explained by their belonging to a multisystemic disease that includes components of the metabolic syndrome and shared pathogenetic mechanisms. Throughout the progression of NAFLD to more advanced stages, complex systemic and local metabolic derangements are involved. During fibrogenesis, a significant metabolic reprogramming occurs in the hepatic stellate cells, hepatocytes, and immune cells, engaging carbohydrate and lipid pathways to support the high-energy-requiring processes. The natural history of NAFLD evolves in a variable and dynamic manner, probably due to the interaction of a variable number of modifiable (diet, physical exercise, microbiota composition, etc.) and non-modifiable (genetics, age, ethnicity, etc.) risk factors that may intervene concomitantly, or subsequently/intermittently in time. This may influence the risk (and rate) of fibrosis progression/regression. The recognition and control of the factors that determine a rapid progression of fibrosis (or its regression) are critical, as the fibrosis stages are associated with the risk of liver-related and all-cause mortality.
... Clinical investigations have included a wide range of novel therapeutic targets and prospective medications for the treatment of obesity and MAFLD; however, the results and efficacy have not yet been validated. [207][208][209][210] Glucagon-like peptide-1 (GLP-1)is an incretin hormone that originates from the gut. Additionally, it reduces caloric intake and stomach emptying while increasing insulin release through beta cells. ...
Article
Full-text available
Obesity,and metabolic dysfunction-associated fatty liver disease (MAFLD) have reached epidemic proportions globally. Obesity and MAFLD frequently coexist and act synergistically to increase the risk of adverse clinical outcomes (both hepatic and extrahepatic). Type 2 diabetes mellitus (T2DM) is the most important risk factor for rapid progression of steatohepatitis and advanced fibrosis. Conversely, the later stages of MAFLD are associated with an increased risk of T2DM incident. According to the proposed criteria, MAFLD is diagnosed in patients with liver steatosis and in at least one in three: overweight or obese, T2DM, or signs of metabolic dysregulation if they are of normal weight. However, the clinical classification and correlation between obesity and MAFLD is more complex than expected. In addition, treatment for obesity and MAFLD are associated with a reduced risk of T2DM, suggesting that liver-based treatments could reduce the risk of developing T2DM. This review describes the clinical classification of obesity and MAFLD, discusses the clinical features of various types of obesity and MAFLD, emphasizes the role of visceral obesity and insulin resistance (IR) in the development of MAFLD,and summarizes the existing treatments for obesity and MAFLD that reduce the risk of developing T2DM.
... Комбинированное использование гепатотропных препаратов является одним из наиболее многообещающих направлений фармакотерапии различных заболеваний печени. Это касается как традиционно используемых гепатопротекторов, так и новых препаратов, только проходящих клинические исследования [24][25][26]. Перспективы и возможности комбинированного применения ГК обусловлены, с одной ...
Article
Full-text available
Глицирризиновая кислота — соединение тритерпеновой природы растительного происхождения, обладающее антистеатозной, антицитолитической, противовоспалительной, антифибротической, антихолестатической, а также антиа поптотической, антинеопластической и другими видами активности. Результаты недавних исследований показывают, что помимо гепатопротекторного действия, глицирризиновая кислота способна к образованию надмолекулярных самоассоциатов и мицелл, что придает ей свой ства фармакокинетического, а следовательно, и фармакодинамического энхансера. Таким образом, перспективы и возможности комбинированного применения глицирризиновой кислоты при заболеваниях печени и других органов обусловлены, с одной стороны, фармакотерапевтическими свойствами самой молекулы, а с другой — ее особенностями как формообразующего вещества. Настоящий обзор посвящен фармакодинамическим и фармакокинетическим аспектам применения глицирризиновой кислоты в комбинациях с эссенциальными фосфолипидами и урсодезоксихолевой кислотой. [Glycyrrhizinic acid is a triterpenoid plant- derived compound with potent antisteatotic, anticytolitic, anti-infl ammatory, antifibrotic, anticholestatic as well as antiapoptotic, antineoplastic and some other eff ects. Recent studies have demonstrated glycyrrhizinic acid to form supramolecular self-associates and micelles, which makes it a pharmacokinetic, and, hence, a pharmacodynamic enhancer. Thus, the prospects and possibilities of combined use of glycyrrhizinic acid in liver disease and other pathologies arise due to the pharmacological properties of the molecule itself as well as its function as drug carrier and delivery enhancer. The present review is focused on the pharmacodynamic and pharmacokinetic features of glycyrrhizinic acid combinations with essential phospholipids and ursodeoxycholic acid.]
Article
Full-text available
Background The human gut microbiota (GM) is a diverse ecosystem crucial for health, impacting physiological processes across the host's body. This review highlights the GM's involvement in Non-Alcoholic Fatty Liver Disease (NAFLD) and explores its diagnosis, treatment, and management. Main Text The GM influences gut functionality, digestion, immunity, and more. Short-chain fatty acids (SCFAs), produced by microbial fermentation, regulate metabolism, inflammation, and immune responses. Bile acids (BAs) modulate the microbiome and liver functions, affecting NAFLD progression. Dysbiosis and increased gut permeability contribute to NAFLD through bacterial components and metabolites reaching the liver, causing inflammation and oxidative stress. The microbiome's impact on immune cells further exacerbates liver damage. Symptoms of NAFLD can be subtle or absent, making diagnosis challenging. Imaging techniques assist in diagnosing and staging NAFLD, but liver biopsy remains vital for accurate assessment. Promising treatments include FXR agonists, GLP-1 agonists, and FGF19 and FGF21 mimetics, targeting various pathways associated with NAFLD pathogenesis. Fecal Microbiota Transplantation (FMT) emerges as a potential therapeutic avenue to restore gut microbiota diversity and alleviate NAFLD. Lifestyle interventions, such as dietary modifications, exercise, and probiotics, also play a pivotal role in managing NAFLD and restoring gut health. Conclusion Despite significant progress, the complex interplay between the gut microbiome, NAFLD, and potential treatments necessitates further research to unravel underlying mechanisms and develop effective therapeutic strategies.
Article
Full-text available
Metabolic-associated Fatty Liver Disease is one of the outstanding challenges in gastroenterology. The increasing incidence of the disease is undoubtedly connected with the ongoing obesity pandemic. The lack of specific symptoms in the early phases and the grave complications of the disease require an active approach to prompt diagnosis and treatment. Therapeutic lifestyle changes should be introduced in a great majority of patients; but, in many cases, the adherence is not satisfactory. There is a great need for an effective pharmacological therapy for Metabolic-Associated Fatty Liver Disease, especially before the onset of steatohepatitis. Currently, there are no specific recommendations on the selection of drugs to treat liver steatosis and prevent patients from progression toward more advanced stages (steatohepatitis, cirrhosis, and cancer). Therefore, in this Review, we provide data on the clinical efficacy of therapeutic interventions that might improve the course of Metabolic-Associated Fatty Liver Disease. These include the drugs used in the treatment of obesity and hyperlipidemias, as well as affecting the gut microbiota and endocrine system, and other experimental approaches, including functional foods. Finally, we provide advice on the selection of drugs for patients with concomitant Metabolic-Associated Fatty Liver Disease.
Article
Full-text available
Aim: In this systematic review, guidelines on non-alcoholic fatty liver disease (NAFLD) were evaluated, aiming at a guideline synthesis focusing on diagnosis and staging. Methods: A systematic literature search was conducted on any relevant database or institutional website to find guidelines on NAFLD assessment intended for clinical use on humans, in English, published from January 2010 to August 2020. Included guidelines were appraised using the AGREE II Instrument; those with higher scores and intended for use in adult patients were included in a comparative analysis. Results: Fourteen guidelines were included in the systematic review, eight of which reached an AGREE II score sufficiently high to be recommended for clinical use, of which one developed for pediatric patients only. British and North American guidelines received the highest scores. Most guidelines recommend a screening or case-finding approach in patients with metabolic risk factors who are at increased risk of steatohepatitis or fibrosis. Ultrasound is mostly recommended to confirm steatosis, while the presence of metabolic syndrome, liver function tests, fibrosis scores, and elastographic techniques may help in selecting high-risk patients to be referred to the hepatologist, who may consider liver biopsy, although referral criteria for liver biopsy are not clearly defined. Most guidelines identify the development of noninvasive tests to replace liver biopsy as a research priority. Conclusion: Several high-quality guidelines exist for NAFLD assessment, with no complete agreement on whether to screen high-risk patients and on the tests and biomarkers suggested to stratify patients and select those to be referred to liver biopsy.
Article
Full-text available
Nonalcoholic fatty liver disease (NAFLD) represents one of the most common liver disorders and can progress into a series of liver diseases, including nonalcoholic steatohepatitis (NASH), fibrosis, cirrhosis, and even liver cancer. Interleukin-22 (IL-22), a member of the IL-10 family of cytokines, is predominantly produced by lymphocytes but acts exclusively on epithelial cells. IL-22 was proven to favor tissue protection and regeneration in multiple diseases. Emerging evidence suggests that IL-22 plays important protective functions against NAFLD by improving insulin sensitivity, modulating lipid metabolism, relieving oxidative and endoplasmic reticulum (ER) stress, and inhibiting apoptosis. By directly interacting with the heterodimeric IL-10R2 and IL-22R1 receptor complex on hepatocytes, IL-22 activates the Janus kinase 1 (JAK1)/ signal transducer and activator of transcription 3 (STAT3), c-Jun N-terminal kinase (JNK) and extracellular-signal regulated kinase (ERK) pathways to regulate the subsequent expression of genes involved in inflammation, metabolism, tissue repair, and regeneration, thus alleviating hepatitis and steatosis. However, due to the wide biodistribution of the IL-22 receptor and its proinflammatory effects, modifications such as targeted delivery of IL-22 expression and recombinant IL-22 fusion proteins to improve its efficacy while reducing systemic side effects should be taken for further clinical application. In this review, we summarized recent progress in understanding the physiological and pathological importance of the IL-22-IL-22R axis in NAFLD and the mechanisms of IL-22 in the protection of NAFLD and discussed the potential strategies to maneuver this specific cytokine for therapeutic applications for NAFLD.
Article
Full-text available
Treatment options for nonalcoholic fatty liver disease (NAFLD) and type 2 diabetes (T2D), two conditions which coexist, are limited though weight loss is an important strategy to improve outcomes in either disease. Glucagon-like peptide 1 receptor agonist (GLP1-RA) present a novel option to treat this dual disease by their salutary effects on glycaemic control and weight reduction. Eight randomized controlled trials on T2D and NAFLD from the Cochrane Library, Embase, and PubMed were included in this meta-analysis. The Comprehensive Meta-Analysis Software version 3 was used to calculate the effect size. In a pooled population of 615 patients—297 on GLP1-RA and 318 in the control arm, GLP1-RA produced a significant improvement in alanine aminotransferase [standardised mean difference (SDM), − 0.56, 95% CI − 0.88 to − 0.25, P < 0.01], aspartate aminotransferase (SDM, − 0.44, SE, 95% CI − 0.64 to − 0.24, P < 0.01), gamma glutaryl transaminase (SDM, − 0.60, 95% CI − 0.86 to − 0.34, P < 0.01) and reduction in liver fat content (LFC) (SDM, − 0.43, 95% CI − 0.74 to − 0.12, P < 0.01), as well as glycosylated haemoglobin (SDM, − 0.40, 95% CI, − 0.61 to − 0.19, P < 0.01) and weight (SDM, − 0.66, 95% CI, − 0.88 to − 0.44, P < 0.01), in comparison to standard of care or placebo. Significant improvement in biopsy resolution was also seen in the GLP1-RA arm (Rate Ratio, 6.60, 95% CI 2.67 to 16.29, P < 0.01). This is possibly the first meta-analysis conducted exclusively in patients with T2D and NAFLD which presents a strong signal that GLP1-RA, improve liver function and histology by improving glycaemia, reducing body weight and hepatic fat, which in turn reduces hepatic inflammation. Trial Registration: PROSPERO (ID: CRD42021228824).
Article
Full-text available
Background: Galectins, a family of β-galactoside-binding proteins, are related to the development and progression of various human diseases such as cancer, heart failure, and chronic kidney disease. However, its role in liver diseases is unclear. Methods: The PubMed, Embase, and Cochrane Library databases were searched. Hazard ratios (HRs), odds ratios (ORs), and mean differences (MDs) with 95% CIs were pooled to evaluate the association of the galectins with the outcomes and risk of liver diseases by a random effects model. Results: Thirty three studies involving 43 cohorts and 4,168 patients with liver diseases were included. In the patients with hepatocellular carcinoma (HCC), high expression of galectin-1 and -3 in the tissues was significantly associated with worse overall survival (galectin-1: HR = 1.94, 95% CI = 1.61–2.34, p < 0.001; galectin-3: HR = 3.29, 95% CI = 1.62–6.68, p < 0.001) and positive vascular invasion (galectin-1: OR = 1.74, 95% CI = 1.18–2.58, p = 0.005; galectin-3: OR = 2.98, 95% CI = 1.58–5.60, p = 0.001); but, high expression of galectin-4 and −9 in the tissues was significantly associated with better overall survival (galectin-4: HR = 0.53, 95% CI = 0.36–0.79, p = 0.002; galectin-9: HR = 0.56, 95% CI = 0.44–0.71, p < 0.001) and negative vascular invasion (galectin-4: OR = 0.36, 95% CI = 0.19–0.72, p = 0.003; galectin-9: OR = 0.60, 95% CI = 0.37–0.97, p = 0.037). Serum galectin-3 level was significantly higher in HCC (MD = 3.06, 95% CI = 1.79–4.32, p < 0.001), liver failure (MD = 0.44, 95% CI = 0.23–0.66, p < 0.001), liver cirrhosis (MD = 1.83, 95% CI = 1.15–2.51, p < 0.001), and chronic active hepatitis B (MD = 18.95, 95% CI = 10.91–27.00, p < 0.001); serum galectin-9 level was significantly higher in HCC (MD = 3.74, 95% CI = 2.57–4.91, p < 0.001) and autoimmune hepatitis (MD = 8.80, 95% CI = 7.61–9.99, p < 0.001). Conclusion: High galectin-1 and -3 and low galectin-4 and -9 expression indicate worse outcomes of patients with HCC. Serum galectin-3 and -9 levels are positively associated with the risk of chronic liver diseases.
Article
Full-text available
Background: Namodenoson, an A3 adenosine receptor (A3AR) agonist, improved liver function/pathology in non-alcoholic steatohepatitis (NASH) preclinical models. Aim: To evaluate the efficacy and safety of namodenoson for the treatment of non-alcoholic fatty liver disease (NAFLD) with or without NASH METHODS: This phase 2 study included 60 patients with NAFLD (ALT ≥60 IU/L) who were randomised (1:1:1) to oral namodenoson 12.5 mg b.d. (n = 21), 25 mg b.d. (n = 19), or placebo (n = 20) for 12 weeks (total follow-up: 16 weeks). The main efficacy endpoint involved serum ALT after 12 weeks of treatment. Results: Serum ALT decreased over time with namodenoson in a dose-dependent manner. The difference between change from baseline (CFB) for ALT in the namodenoson 25 mg b.d. arm vs placebo trended towards significance at 12 weeks (P = 0.066). Serum AST levels also decreased with namodenoson in a dose-dependent manner; at 12 weeks, the CFB for 25 mg b.d. vs placebo was significant (P = 0.03). At Week 12, 31.6% in the namodenoson 25 mg b.d. arm and 20.0% in the placebo arm achieved ALT normalisation (P = 0.405). At week 16, the respective rates were 36.8% and 10.0% (P = 0.038). A3AR expression levels were stable over time across study arms. Both doses of namodenoson were well tolerated with no drug-emergent severe adverse events, drug-drug interactions, hepatotoxicity, or deaths. Three adverse events were considered possibly related to study treatment: myalgia (12.5 mg b.d. arm), muscular weakness (25 mg b.d. arm), and headache (25 mg b.d. arm). Conclusion: A3AR is a valid target; namodenoson 25 mg b.d. was safe and demonstrated efficacy signals (ClinicalTrials.gov #NCT02927314).
Article
Full-text available
Non-alcoholic fatty liver disease (NAFLD) is a growing cause of chronic liver disease worldwide. It is characterised by steatosis, liver inflammation, hepatocellular injury and progressive fibrosis. Several preclinical models (dietary and genetic animal models) of NAFLD have deepened our understanding of its aetiology and pathophysiology. Despite the progress made, there are currently no effective treatments for NAFLD. In this review, we will provide an update on the known molecular pathways involved in the pathophysiology of NAFLD and on ongoing studies of new therapeutic targets.
Article
Full-text available
Alterations in lipid metabolism might contribute to the pathogenesis of non-alcoholic fatty liver disease (NAFLD). However, no pharmacological agents are currently approved in the United States or the European Union for the treatment of NAFLD. Two parallel phase 2a studies investigated the effects of liver-directed ACC1/2 inhibition in adults with NAFLD. The first study (NCT03248882) examined the effects of monotherapy with a novel ACC1/2 inhibitor, PF-05221304 (2, 10, 25 and 50 mg once daily (QD)), versus placebo at 16 weeks of treatment; the second study (NCT03776175) investigated the effects of PF-05221304 (15 mg twice daily (BID)) co-administered with a DGAT2 inhibitor, PF-06865571 (300 mg BID), versus placebo after 6 weeks of treatment. The primary endpoint in both studies was percent change from baseline in liver fat assessed by magnetic resonance imaging–proton density fat fraction. Dose-dependent reductions in liver fat reached 50–65% with PF-05221304 monotherapy doses ≥10 mg QD; least squares mean (LSM) 80% confidence interval (CI) was −7.2 (−13.9, 0.0), −17.1 (−22.7, −11.1), −49.9 (−53.3, −46.2), −55.9 (−59.0, −52.4) and −64.8 (−67.5, −62.0) with 16 weeks placebo and PF-05221304 2, 10, 25 and 50 mg QD, respectively. The overall incidence of adverse events (AEs) did not increase with increasing PF-05221304 dose, except for a dose-dependent elevation in serum triglycerides (a known consequence of hepatic acetyl-coenzyme A carboxylase (ACC) inhibition) in 23/305 (8%) patients, leading to withdrawal in 13/305 (4%), and a dose-dependent elevation in other serum lipids. Co-administration of PF-05221304 and PF-06865571 lowered liver fat compared to placebo (placebo-adjusted LSM (90% CI) −44.6% (−54.8, −32.2)). Placebo-adjusted LSM (90% CI) reduction in liver fat was −44.5% (−55.0, −31.7) and −35.4% (−47.4, −20.7) after 6 weeks with PF-05221304 or PF-06865571 alone. AEs were reported for 10/28 (36%) patients after co-administered PF-05221304 and PF-06865571, with no discontinuations due to AEs, and the ACC inhibitor-mediated effect on serum triglycerides was mitigated, suggesting that PF-05221304 and PF-06865571 co-administration has the potential to address some of the limitations of ACC inhibition alone. Two phase 2a trials demonstrate the efficacy of a new ACC inhibitor (PF 05221304) for reducing liver fat in patients with NAFLD, with co-administration of a DGAT2 inhibitor (PF-06865571) mitigating ACC inhibitor-mediated increases in serum triglycerides.
Article
Full-text available
Nonalcoholic steatohepatitis (NASH), a chronic liver disease without an approved therapy, is associated with lipotoxicity and insulin resistance and is a major cause of cirrhosis and hepatocellular carcinoma. Aramchol, a partial inhibitor of hepatic stearoyl-CoA desaturase (SCD1) improved steatohepatitis and fibrosis in rodents and reduced steatosis in an early clinical trial. ARREST, a 52-week, double-blind, placebo-controlled, phase 2b trial randomized 247 patients with NASH (n = 101, n = 98 and n = 48 in the Aramchol 400 mg, 600 mg and placebo arms, respectively; NCT02279524). The primary end point was a decrease in hepatic triglycerides by magnetic resonance spectroscopy at 52 weeks with a dose of 600 mg of Aramchol. Key secondary end points included liver histology and alanine aminotransferase (ALT). Aramchol 600 mg produced a placebo-corrected decrease in liver triglycerides without meeting the prespecified significance (−3.1, 95% confidence interval (CI) −6.4 to 0.2, P = 0.066), precluding further formal statistical analysis. NASH resolution without worsening fibrosis was achieved in 16.7% (13 out of 78) of Aramchol 600 mg versus 5% (2 out of 40) of the placebo arm (odds ratio (OR) = 4.74, 95% CI = 0.99 to 22.7) and fibrosis improvement by ≥1 stage without worsening NASH in 29.5% versus 17.5% (OR = 1.88, 95% CI = 0.7 to 5.0), respectively. The placebo-corrected decrease in ALT for 600 mg was −29.1 IU l⁻¹ (95% CI = −41.6 to −16.5). Early termination due to adverse events (AEs) was <5%, and Aramchol 600 and 400 mg were safe, well tolerated and without imbalance in serious or severe AEs between arms. Although the primary end point of a reduction in liver fat did not meet the prespecified significance level with Aramchol 600 mg, the observed safety and changes in liver histology and enzymes provide a rationale for SCD1 modulation as a promising therapy for NASH and fibrosis and are being evaluated in an ongoing phase 3 program.
Article
Background Management of nonalcoholic steatohepatitis (NASH) is an unmet clinical need. Lanifibranor is a pan-PPAR (peroxisome proliferator–activated receptor) agonist that modulates key metabolic, inflammatory, and fibrogenic pathways in the pathogenesis of NASH. Methods Download a PDF of the Research Summary. In this phase 2b, double-blind, randomized, placebo-controlled trial, patients with noncirrhotic, highly active NASH were randomly assigned in a 1:1:1 ratio to receive 1200 mg or 800 mg of lanifibranor or placebo once daily for 24 weeks. The primary end point was a decrease of at least 2 points in the SAF-A score (the activity part of the Steatosis, Activity, Fibrosis [SAF] scoring system that incorporates scores for ballooning and inflammation) without worsening of fibrosis; SAF-A scores range from 0 to 4, with higher scores indicating more-severe disease activity. Secondary end points included resolution of NASH and regression of fibrosis. Results A total of 247 patients underwent randomization, of whom 103 (42%) had type 2 diabetes mellitus and 188 (76%) had significant (moderate) or advanced fibrosis. The percentage of patients who had a decrease of at least 2 points in the SAF-A score without worsening of fibrosis was significantly higher among those who received the 1200-mg dose, but not among those who received the 800-mg dose, of lanifibranor than among those who received placebo (1200-mg dose vs. placebo, 55% vs. 33%, P=0.007; 800-mg dose vs. placebo, 48% vs. 33%, P=0.07). The results favored both the 1200-mg and 800-mg doses of lanifibranor over placebo for resolution of NASH without worsening of fibrosis (49% and 39%, respectively, vs. 22%), improvement in fibrosis stage of at least 1 without worsening of NASH (48% and 34%, respectively, vs. 29%), and resolution of NASH plus improvement in fibrosis stage of at least 1 (35% and 25%, respectively, vs. 9%). Liver enzyme levels decreased and the levels of the majority of lipid, inflammatory, and fibrosis biomarkers improved in the lanifibranor groups. The dropout rate for adverse events was less than 5% and was similar across the trial groups. Diarrhea, nausea, peripheral edema, anemia, and weight gain occurred more frequently with lanifibranor than with placebo. Conclusions In this phase 2b trial involving patients with active NASH, the percentage of patients who had a decrease of at least 2 points in the SAF-A score without worsening of fibrosis was significantly higher with the 1200-mg dose of lanifibranor than with placebo. These findings support further assessment of lanifibranor in phase 3 trials. (Funded by Inventiva Pharma; NATIVE ClinicalTrials.gov number, NCT03008070.) QUICK TAKE VIDEO SUMMARY A Pan-PPAR Agonist in NASH 02:04