1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63

REVIEW ARTICLE Obesity and the liver: nonalcoholic fatty liver disease Q23

SEAN W. P. KOPPE CHICAGO, ILL

The increasing prevalence of nonalcoholic fatty liver disease (NAFLD) parallels the rise of obesity and its complications. NAFLD is a common cause of cirrhosis and a leading indication for liver transplant. Genetic susceptibility, dietary composition, and exercise habits influence the development of NAFLD, and insulin resistance results in widespread metabolic perturbations with a net effect of triglyceride accumulation in the liver. Some patients will develop hepatocyte cellular injury and fibrosis of the liver, which can progress to cirrhosis and require liver transplant. Treatments targeting the pathophysiological mechanisms of NAFLD exist, but carry some potential risk and are not universally effective. Weight loss and lifestyle changes remain the most effective and safest approach, but sustainable change is difficult for most patients to achieve. Future work will continue to focus on developing effective and safe interventions to prevent the development of advanced liver disease, whereas efforts in the public health domain continue to combat obesity. (Translational Research 2014;-:1–11) Abbreviations: ApoC3 ¼ apolipoprotein C III; ChREBP ¼ carbohydrate response–elementbinding protein; CK-18 ¼ cytokeratin-18; DAG ¼ diacylglycerol; ER ¼ endoplasmic reticulum; HCC ¼ hepatocellular carcinoma; IRS-2 ¼ insulin receptor substrate 2; LOXL2 ¼ lysyl oxidaselike 2; LPS ¼ lipopolysaccharide; LXR ¼ liver X receptor; MET ¼ metabolic equivalent of task; MnSOD ¼ manganese-superoxide dismutase 2; NAFLD ¼ nonalcoholic fatty liver disease; NASH ¼ nonalcoholic steatohepatitis; OCA ¼ obeticholic acid; PKCε ¼ protein kinase Cε; PNPLA3 ¼ patatin-like phospholipase domain containing 3; ROS ¼ reactive oxygen species; SREBP-1 ¼ sterol regulatory element-binding protein 1; TLR4 ¼ toll-like receptor 4

Q3

Q4

INTRODUCTION

T

he exploding rate of obesity in the United States over the past several decades has led to an increased prevalence of diabetes and cardiovascular disease as well as nonalcoholic fatty liver disease (NAFLD). NAFLD is present in 20%–30% of adults in the United

Q1

States and 10% of children.1-3 NAFLD represents a spectrum of disease with most patients having only steatosis; however, a sizable number of patients can also develop inflammation, cellular injury, and fibrosis, termed nonalcoholic steatohepatitis (NASH). The cellular injury in NASH is characterized by ballooned

From the Division of Gastroenterology and Hepatology, Northwestern Medicine, Chicago, Ill.

1931-5244/$ - see front matter

Submitted for publication March 31, 2014; revision submitted June 18, 2014; accepted for publication June 19, 2014.

http://dx.doi.org/10.1016/j.trsl.2014.06.008

Ó 2014 Mosby, Inc. All rights reserved.

Reprint requests: Sean W.P. Koppe, Kovler Organ Transplantation Center, 676 N. St. Clair, Suite 1900, Chicago, IL 60611; e-mail: [email protected].

1 REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127

2

128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191

Translational Research - 2014

Koppe

20%

% Liver Transplants for Alcohol and NASH/Cryptogenic Cirrhosis

18% 16% 14% 12% 10% 8% 6% 4% 2% 0% 2000 2001 2002 2003 2004 2005 2006 2007 2008 2009 2010 2011 2012 2013

Fig 1. Percentage of liver transplants for alcohol and NASH/cryptogenic cirrhosis. The number of patients transplanted with a primary diagnosis of NASH/cryptogenic cirrhosis (black) has steadily risen and is nearly equal to the number of patients transplanted with primary diagnosis of alcohol (gray). Hepatitis C remains the most common indication for liver transplant. Patients with a primary diagnosis of hepatocellular carcinoma are excluded. http://optn.transplant.hrsa.gov. NASH, nonalcoholic steatohepatitis.

Q5

Q6

hepatocytes, which frequently contain Mallory-Denk bodies.4 Patients with NAFLD found to have only steatosis are unlikely to progress to cirrhosis; however, steatosis can accelerate other forms of liver disease.5-7 Patients found to have NASH represent a subset of those with NAFLD at the greatest risk of progressing to cirrhosis, and it is estimated that 3%–5% of the general population has NASH.8 NASH now represents the third most common indication for liver transplantation, although at its current trend it is expected to soon eclipse alcohol as an indication for liver transplantation (Fig 1). In addition to increasing rates of transplantation for NASH, increased prevalence of obesity and diabetes in potential cadaveric donors has led to increased organ discard rate exacerbating the mismatch between supply and demand for liver transplants.9 The rising rate of childhood obesity has also led to the development of NAFLD/NASH in the pediatric population. Estimates based on elevated ALT and an autopsy series suggest a prevalence of pediatric NAFLD of 8%–10%.10 Risk factors for pediatric NAFLD mirror those for adult NAFLD and children with NASH can even progress to cirrhosis and require liver transplantation.11 Similar histologic changes are often noted in pediatric NASH compared with adult NASH; however, some cases of pediatric NASH present with less

hepatocyte ballooning and more portal-based inflammation and fibrosis.4,10 Although NAFLD leads to increased morbidity and mortality from liver disease, the most frequent cause of morbidity and mortality in the NAFLD population remains cardiovascular. Efforts continue to better understand the pathophysiology of NAFLD and develop treatments addressing those patients who develop NASH and have progressive liver disease. However, NAFLD should continue to be viewed as the hepatic manifestation of the metabolic syndrome and treatment of associated diabetes, hyperlipidemia, and hypertension are equally as important as the treatment of the actual liver disease. This review will discuss current understanding of the pathophysiology of NAFLD/ NASH and its management. PATHOGENESIS OF NAFLD/NASH

Fat accumulation in the liver is the culmination of dysregulation of fatty acid influx to the liver, fatty acid efflux, and hepatic de novo lipogenesis. Circulating adipokines and cytokines mediate and result from this process. Insulin resistance is near universally present in NAFLD and diabetes is a very frequent finding in those patients who develop NASH. Lipotoxicity,

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

Q22

192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255

Translational Research Volume -, Number 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319

Koppe

3

Fig 2. Pathways of liver injury in NAFLD. Insulin resistance is characterized by hyperinsulinemia and elevated levels of circulating FFAs. Elevated insulin activates the lipogenic transcription factor sterol regulatory binding protein-1c and coupled with increased FFA levels results in increased de novo lipogenesis. Elevated FFA leads to increased DAG levels, which further worsens insulin resistance. Elevated FFA also results in increased levels of ceramides, increased fatty acid oxidation leading to generation of ROS, and increased ER stress. Collectively, these processes result in generation of proinflammatory cytokines and reduced adiponectin levels. Increased inflammatory cytokines can also be generated by the interaction of LPS and TLR4. Saturated fatty acids and ceramides also serve as non-LPS agonists of TLR4 leading to generation of proinflammatory cytokines. Genetics and lifestyle factors also influence the development of NAFLD. DAG, diacylglycerol; ER, endoplasmic reticulum; FFA, free fatty acid; IL-6, interleukin 6; IL-10, interleukin 10; LPS, lipopolysaccharide; NAFLD, nonalcoholic fatty liver disease; PNPLA3, patatin-like phospholipase domain containing 3; ROS, reactive oxygen species; TLR4, toll-like receptor 4; TNF a, tumor necrosis factor a.

Q7

mitochondrial dysfunction, oxidative stress, and endoplasmic reticulum (ER) stress are involved in hepatic injury. Adipose tissue distribution and genetic factors influence this process, and gut microbiota may also play a role (Fig 2). Genetics. Certain groups tend to have a higher prevalence of NAFLD/NASH with Mexican-Americans having a particularly high burden (24%) compared with non-Hispanic whites (18%) and non-Hispanic blacks (14%).1 There may be some differences based on alternate dietary patterns; however, genetic factors are a significant contributor to these differences. Patatin-like phospholipase domain containing 3 (PNPLA3) is a gene strongly associated with susceptibility to develop steatosis and progressive liver disease and may explain some of the racial differences. The precise function of PNPLA3 is unclear, but it appears to be involved with acylglycerol synthesis and also hydrolysis. Liver X receptor (LXR) and sterol regulatory element-binding protein 1 (SREBP-1) appear to regulate its expression, and fatty acids synthesized in the liver then delay degradation of PNPLA3.12 A variant of the PNPLA3 gene (rs738409[G]—I148M) has been shown to be significantly associated with hepatic steatosis, independent of body mass index (BMI) and the presence of diabetes, and also shown to

be associated with increased risk of NASH and even hepatocellular carcinoma (HCC).13-15 The initial study that discovered the importance of this particular allele found that it was most frequently found in Hispanics (0.49), followed by European Americans (0.23) and African Americans (0.17) and is therefore likely a strong factor in the racial differences observed with NAFLD.16 A precise understanding of the mechanism by which this polymorphism leads to liver disease is lacking but likely relates to liver-specific PNPLA3 dysfunction rather than adipose tissue PNPLA3 dysfunction.17 However, the finding that PNPLA3 genotype of liver transplant recipients is more important than the PNPLA3 genotype of donors in the development of post-transplant hepatic steatosis suggests that there may be other factors involved.18 Apolipoprotein C III (ApoC3) is found on VLDL and inhibits lipoprotein lipase activity, thereby reducing peripheral fat uptake and increasing serum triglycerides. Two common gene variants resulting in Q8 increased ApoC3 levels were found to place individuals at a higher risk of hypertriglyceridemia and NAFLD in the setting of modest caloric excess.19,20 ApoC3 accentuates steatosis related to caloric excess and its effect can be countered by low calorie, low-fat intake.21 A polymorphism in the SOD2 gene encoding Q9

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383

4

384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447

Q10

Q11

Translational Research - 2014

Koppe

manganese-superoxide dismutase 2 may also contribute to the severity of NAFLD by altering sensitivity to oxidative stress. Other genetic polymorphisms enhancing susceptibility to NAFLD/NASH have also been reported.22,23 Insulin resistance and lipotoxicity. Increased adiposity and insulin resistance leads to elevated circulating levels of free fatty acids (FFAs) in patients with NAFLD. Insulin resistance leads to impaired suppression of lipolysis in all adipose tissues, and visceral adipose tissue is felt to be a particularly significant contributor to hepatic FFA exposure given its drainage via the splanchnic circulation.24 Elevated hepatic FFAs result in increased triglyceride synthesis and worsening insulin resistance mediated via lipid intermediates. Hyperinsulinemia also leads to activation of SREBP-1c, a key transcription factor regulating triglyceride synthesis. SREBP-1c activation coupled with increased availability of fatty acids promotes steatosis. Although triglyceride synthesis plays some pathologic role, particularly in insulin resistance, it is also viewed as a potential protective mechanism to lessen the damage produced by FFAs and other lipid intermediates.25,26 Lipid-derived metabolites are a significant contributor to the insulin resistance observed in patients with NAFLD. Elevated FFAs can result in increased diacylglycerol (DAG) formation, and elevated levels of DAG have been documented in patients with NAFLD.27-30 Accumulating DAG interferes with insulin signaling in the liver by activating protein kinase Cε leading to reduced insulin-stimulated tyrosine phosphorylation of the insulin receptor substrate 2.31 Ceramides, sphingolipids important in the lipid bilayer of cellular membranes, are also increased in the setting of increased FFAs in the liver.32-34 Ceramides may contribute to insulin resistance through inactivation of Akt; however, the evidence for ceramides contributing directly to insulin resistance in the liver is conflicting.30,34,35 Ceramides appear to play a role in NAFLD and progression to NASH, but this may be related to ceramideassociated generation of proinflammatory cytokines, apoptosis, and/or ER stress.36,37 FFAs in the liver can be esterified into triglycerides or enter the mitochondria to undergo b-oxidation. During the process of b-oxidation, reactive oxygen species (ROS) are generated via the mitochondrial respiratory chain. Increased fatty acid levels in NAFLD result in increased mitochondrial fatty acid b-oxidation.38-40 Increases in fatty acid b-oxidation in NAFLD are accompanied by mitochondrial structural changes and increased ROS generation compatible with mitochondrial dysfunction.38,41,42 Cytochrome P450 2E1, present in high concentrations in mitochondria, is

increased in NAFLD and also contributes to increased ROS generation.43,44 Increased ROS can result in increased lipid peroxidation, which is well described in patients with NAFLD and NASH.45-47 Elevated ROS also contributes to proinflammatory cytokine production and cellular apoptosis resulting in progression to NASH.48 ER stress is also implicated in the development of NAFLD and progression to NASH. ER stress can result in the unfolded protein response, an adaptive mechanism by cells to deal with threats to homeostasis; however, with prolonged stress stimuli the unfolded protein response can result in apoptosis and cell death.49,50 Elevated levels of FFAs can lead to ER stress in the liver, likely because of unfavorably altering the amount of saturated phospholipid in ER membranes resulting in decreased membrane fluidity with loss of functionality.51-53 Increased markers of ER stress have been noted in patients with NAFLD and NASH, supporting an important role for this pathway.54-56 Activation of the ER stress pathway promotes apoptosis and inflammation and also is noted to worsen insulin resistance.49 Adiponectin, innate immune system, endotoxin and microbiota. Certain cytokines produced by adipose

tissue or immune cells within the liver or circulating in the periphery appear to play an important role in the pathogenesis of NAFLD and progression of liver disease. Tumor necrosis factor a (TNF-a), interleukin 6, interleukin 10, and adiponectin are adipocytokines frequently implicated in NAFLD. Adiponectin is pro- Q12 duced by adipose tissue and has been found to improve insulin sensitivity and decrease inflammation. Adiponectin is generally protective and reduced levels are frequently found in patients with progressive NAFLD, although this association is not universally present.57,58 Adiponectin appears to reduce ceramide levels via induction of a ceramidase in various tissues with resulting improvement in insulin sensitivity and reduced inflammatory cytokines.59 Curiously, patients with NASH cirrhosis have been found to have paradoxically elevated adiponectin levels despite reduced adipose tissue and loss of hepatic steatosis (ie, ‘‘burnt out NASH’’ or cryptogenic cirrhosis). This has been explained by increased levels of bile acids mediating production of adiponectin via farnesoid X receptor (FXR).60 Toll-like receptor 4 (TLR4) is an innate pattern recognition receptor found in greatest abundance on innate immune cells such as macrophages and dendritic cells, but is also found in other cells that are participants in developing NAFLD and its progression to NASH, such as hepatocytes, adipocytes, and the hepatic stellate cell—a critical mediator of liver fibrosis.61-63 Endotoxin

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511

Translational Research Volume -, Number 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575

Koppe

(lipopolysaccharide) is a potent activator of TLR4 and gut endotoxemia has long been suspected as an important potentiator of various liver diseases, including NAFLD.64 TLR4 activation results in proinflammatory cytokine production and can also contribute to insulin resistance.61,64 Additionally, there is evidence that TLR4 can be activated by nonendotoxin products particularly relevant in NAFLD, such as saturated fatty acids and ceramide.37,65,66 The gut microbiome has become increasingly recognized as an important factor in human health and disease, and work is now beginning to focus on the potential role the microbiome may play in the pathogenesis of obesity and NAFLD. Dietary composition can quickly modify the microbiome and provides an additional pathway by which diet influences obesity and its complications, such as NAFLD.67 Alterations in the microbiome are believed to influence obesity and NAFLD through generation of microbial products, which interact locally with enterocytes and influence signaling involved in obesity and NAFLD, increased permeability of the gut barrier leading to increased exposure to lipopolysaccharide and other TLR activators involved in inflammation, and through changes leading to increased availability of calories or pathologically altered micronutrient composition resulting from microbial metabolism of nutrients.68 Germ-free mice have improved insulin sensitivity and reduced inflammatory cytokines in response to high-fat diet.69 Transfer of microbiota from mice prone to develop insulin resistance also leads to higher rates of insulin resistance and upregulation of lipogenesis resulting in increased fatty liver when placed on a high-fat diet.70 A small study in humans has also suggested influence of microbiota on insulin resistance when transfer of microbiota from lean subjects resulted in improved insulin sensitivity in recipients who were obese with the metabolic syndrome.71 Additionally, work in mouse models suggests that the microbiome may influence the risk of developing HCC in obesity-related liver disease.72,73 MANAGEMENT OF NAFLD/NASH

According to the joint practice guideline from the American Association for the Study of Liver Disease, the American College of Gastroenterology, and the American Gastroenterological Association, the diagnosis of NAFLD requires the presence of steatosis by imaging or histology, the lack of any significant alcohol consumption, and the exclusion of other forms of chronic liver disease as well as no plausible alternative explanation for the development of hepatic steatosis.8 Simple steatosis is generally not felt to represent a significant risk to develop progressive fibrosis; however, it

5

is difficult to determine if a patient has simple steatosis or steatohepatitis (NASH) without an invasive liver biopsy. Patient characteristics can be useful in predicting the likelihood of NASH and helping to better determine which patients may be more likely to warrant a liver biopsy. The presence of the metabolic syndrome increases the likelihood of a patient having NASH and might factor in the decision to perform a liver biopsy. Additionally, determination of the NAFLD fibrosis score (http://nafldscore.com), a noninvasive scoring system using readily available clinical characteristics (age, BMI, platelet count, hyperglycemia, albumin, and AST-to-ALT ratio), could also be useful for targeting liver biopsies toward those expected to most likely have advanced fibrosis.8 Findings on biopsy of ballooning degeneration, inflammatory infiltrates, mega mitochondria, and fibrosis/cirrhosis are consistent Q13 with a diagnosis of NASH. Noninvasive biomarkers such as cytokeratin-18 (CK-18) and terminal peptide of procollagen III (PIIINP) hold promise but are not Q14 yet widely accepted. Apoptosis in the liver activates caspase 3 and results in cleaving of CK-18, an intermediate filament protein in the hepatocyte. Elevations in CK-18 fragments have been suggested to be a predictor of NASH; however, a more recent study suggests that it lacks suitable sensitivity to differentiate between NASH and simple steatosis.74,75 PIIINP is a byproduct of type III collagen and was shown in a recent study to differentiate between simple steatosis and NASH or advanced fibrosis, but further studies will be needed to validate its use.76 Noninvasive imaging techniques (ultrasound transient elastography and magnetic resonance imaging elastography) also exist for assessing Q15 patients for fibrosis in NAFLD and emerging data suggest that these may prove to be a useful alternative to liver biopsy; however, at the present time they remain investigational.77,78 The least expensive and most effective treatment for NAFLD/NASH is weight loss. Modest weight loss of 5% can often improve steatosis and liver chemistries, and weight loss of 10% is frequently associated with histologic improvement and lower risk of disease progression.79,80 In a prospective study that followed patients up to 5 years after bariatric surgery, sustained weight loss was associated with a significant reduction in steatosis and a significant reduction in the amount of hepatocyte ballooning, the characteristic lesion of NASH.81 Additionally, although there was a slight overall increase in fibrosis of the cohort, 96% of patients remained at a fibrosis level #F1.81 Weight loss has been shown to reduce markers of ER stress in adipose and liver tissues in a cohort of patients who underwent gastric bypass surgery.54 Adiponectin levels have also been found to favorably increase with weight loss after

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639

6

640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703

Translational Research - 2014

Koppe

laparoscopic gastric banding surgery.82 Roux-en-Y gastric bypass and weight loss achieved by lifestyle changes have been shown to reduce plasma ceramide levels resulting in a more favorable milieu in the liver as well.83,84 Exercise, independent of weight loss, can also be associated with improved insulin sensitivity and improvement of NASH and should be recommended to all patients. Exercise in patients with NAFLD for just 1 week has been shown to favorably increase the relative amounts of polyunsaturated lipids in the liver, increase serum adiponectin levels, and decrease levels of circulating CK-18.85,86 The intensity of exercise may also play a role in reducing the risk of NASH and fibrosis. An analysis of patients enrolled in the NASH Clinical Research Network database found that patients who reported vigorous exercise ($6 metabolic equivalent of task [MET]) for at least 75 min/wk were less likely to have NASH and advanced fibrosis.87 Light jogging is the equivalent of 6 MET and running at 5 mph is the equivalent of 8.3 MET. A comprehensive list of activities ranging from home repair to exercise to religious activities to hunting have been assigned a ‘‘MET’’ value and a review of the list can be both enlightening and entertaining for patients and clinicians to quantify the intensity and duration of their activities.88 Diet composition can have an influence on NAFLD regardless of overall calorie consumption. Although excess caloric intake clearly plays a role in the development of obesity and NAFLD, diets high in polyunsaturated fat and low in saturated fats and trans-fats are likely beneficial in NAFLD. Trans-fatty acids may be particularly toxic to the liver, but fortunately their consumption has been significantly decreased in recent years, and there has become widespread acceptance regarding their toxicity.89-91 Fructose consumption has more than doubled over the past several decades with much of that consumption in the form of soda, which has been shown to confer an increased risk of NAFLD regardless of BMI.92,93 Fructose consumption stimulates carbohydrate response–element-binding protein and SREBP-1c, which serves to increase de novo lipogenesis and fructose has a more pronounced effect on de novo lipogenesis, dyslipidemia, insulin sensitivity, and visceral adiposity compared with an isocaloric glucose load.94-96 Coffee consumption has been linked to lower rates of obesity, diabetes, cardiovascular disease, and mortality.97-99 Coffee consumption appears to also have a favorable impact on NAFLD with an inverse relationship between coffee consumption and degree of fibrosis.100 Coffee consumption is associated with reduced fibrosis from a variety of liver diseases and also lowers the risk of developing HCC.101,102 The

benefit of coffee is limited to filtered coffee and there are likely several components of coffee contributing to its benefits, but caffeine is likely an important mediator. Caffeine has been shown to reduce levels of the profibrotic cytokine transforming growth factor beta and reduce hepatic fibrosis in animal models.103 Hepatic stellate cells, responsible for coordinating the fibrotic response in the liver, are activated via the A2A adenosine receptor and given that caffeine is an established antagonist of adenosine receptors, it is plausible that this is an important mechanism by which coffee/ caffeine reduces liver fibrosis.103,104 Q16 Pharmacologic treatment may be considered for patients with NASH when efforts at weight loss are unsuccessful. Evidence supports a modest benefit of high-dose vitamin E (800 U) and the peroxisome proliferator–activated receptor gamma (PPAR-g) agonist pioglitazone for reducing steatosis, inflammation, and liver enzymes.8,105,106 Pioglitazone and vitamin E have not been shown to definitively reduce liver fibrosis; however, it is also possible that trials of longer duration would be needed to demonstrate this benefit. Vitamin E functions as an antioxidant and PPAR-g is a member of the nuclear hormone receptor superfamily and has a central role in adipose tissue metabolism and function. PPAR-g influences adipogenesis, adipose differentiation, circulating FFA levels, and adipokine levels.107 PPAR-g activation results in reduced circulating FFA levels, improved insulin sensitivity, and increased adiponectin levels.107 Rare individuals who possess dominant-negative mutations in PPAR-g offer insight into the importance of the role of PPAR-g as these individuals exhibit the metabolic syndrome, central adiposity, low adiponectin levels, and nonalcoholic fatty liver.108 Pioglitazone and vitamin E have shown some benefit in adult patients with NASH; however, adverse effects and concerns about long-term use dampen the enthusiasm of many clinicians. Vitamin E has also been shown to be of potential benefit in children and adolescents with NASH. In the TONIC trial, vitamin E was shown to improve Q17 some histologic endpoints and more frequently led to resolution of NASH compared with placebo.109 PPAR-g agonists are associated with weight gain and pioglitazone was associated with 4.7 kg weight gain in the large trial showing benefit in NASH.105,110 For a population already grappling with obesity, this weight gain can be difficult to rationalize. Additionally, concerns about long-term use of PPAR-g agonists increasing the risk of bladder cancer, causing fluid retention, and reducing bone mineral density also must be factored into the decision to use pioglitazone.107 Vitamin E use is also associated with potential safety concerns, namely the possible increase in overall

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767

Translational Research Volume -, Number 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831

Q18

Koppe

mortality that has been reported in prior meta-analysis and potential increased risk of prostate cancer and hemorrhagic stroke.106 The safety of long-term vitamin E usage in the pediatric population is also unknown. Although pioglitazone and vitamin E offer some benefit for patients with NASH, the concerns mentioned previously raise concern about their widespread use and limit their usefulness for chronic management of NASH. Pentoxifylline may also provide some modest benefit to patients with NASH. Pentoxifylline may produce its effects through reducing oxidative stress, reducing TNF-a production, and possibly via antagonism of the A2A adenosine receptor on hepatic stellate cells, similar to caffeine. Small, randomized controlled trials have suggested a benefit of pentoxifylline in NASH and a recent meta-analysis supports a benefit as well.111-113 Pentoxifylline was shown to decrease oxidized lipid products and likely has some impact on TNF-a as well.111,114 Pentoxifylline is a methylxanthine, similar to caffeine, and can act as an A2A adenosine receptor antagonist.115,116 Blockade of the A2A adenosine receptor has been shown to decrease liver fibrosis and this may be an underappreciated beneficial mechanism of pentoxifylline.104,117 Although there is evidence suggesting a benefit of pentoxifylline for NASH, larger randomized prospective trials are needed to recommend its widespread use. The semisynthetic bile acid derivative obeticholic acid (OCA) functions as an FXR agonist and has shown promising results in early clinical trials in patients with NASH. FXR is a nuclear hormone receptor that is regulated by bile acids and influences a wide array of metabolic pathways, including bile acid metabolism, glucose homeostasis, lipid metabolism, inflammatory responses, and vascular remodeling.118,119 Activation of FXR can result in downregulation of SREBP-1c leading to reduced hepatic triglyceride synthesis and can lead to decreased circulating VLDL levels via FXRmediated increase in VLDL receptor expression.120,121 FXR has also been shown to antagonize nuclear factor kb–mediated inflammatory response in the liver and also to inhibit hepatic stellate cell activation resulting in decreased liver fibrosis.122,123 FXR has also been shown in animal models to favorably reduce gluconeogenesis and improve insulin sensitivity.124,125 Interestingly, FXR signaling has also recently been shown to be an important mediator of improved glucose tolerance and sustained weight loss after bariatric surgery (vertical sleeve gastrectomy).126 This appears to be related to elevated circulating levels of bile acids that are observed after bariatric surgery.127,128 OCA’s potential multifaceted beneficial effects as an FXR agonist have been observed in early clinical trials of patients with NASH. In 1 small trial it was shown to

7

improve insulin sensitivity, improve liver biochemistries, and reduce markers of liver fibrosis.129 A phase II clinical trial (‘‘FLINT’’-ClinicalTrials.gov NCT012 65498) was then recently stopped early because OCA showed a significant improvement in liver histology during an interim analysis when compared with placebo. These early findings are encouraging but enthusiasm for OCA is tempered by the unexpected finding of a modest increase in LDL in both trials by unclear Q19 mechanisms.129 Elevating LDL in a population with a high prevalence of CAD will be a major concern unless Q20 it is ultimately shown to not increase the development of atherosclerosis or increase the risk of cardiac events. Another promising therapeutic under investigation is a direct antifibrotic approach with the inhibition of collagen cross-linking. Lysyl oxidase-like 2 promotes cross-linking of type 1 collagen, which results in stabilization of the extracellular matrix. Pharmacologic lysyl oxidase-like 2 inhibitors have been developed and clinical trials in various liver disorders are underway, including a trial examining their use in patients with NASH.130 There remains an unmet need for an effective, safe medication for NASH to prevent progression of liver disease and reduce the need for liver transplant and the risk of developing liver cancer; however, clinicians must continue to address the increased risk of cardiovascular mortality in patients with NAFLD/NASH. Statin use in patients with chronic, stable liver disease is well tolerated, and when indicated, should not be withheld from patients with NAFLD/NASH.131 CONCLUSIONS

The obesity epidemic has led to an increased burden of liver disease with increasing numbers of patients requiring liver transplants for NASH cirrhosis and its complications. The pathogenesis of NAFLD involves insulin resistance, adipose tissue distribution, and dietary and genetic factors that place an individual at risk of NASH. Like many obesity-related complications, addressing NAFLD/NASH is best accomplished with preventive approaches to promote calorie restriction and favorable dietary composition before liver disease develops. Weight loss and exercise remain inexpensive and effective solutions but few are successful at achieving durable lifestyle changes. Promising therapeutic targets are being addressed with modest success and ongoing research to delineate the mechanisms of steatohepatitis and fibrosis will likely lead to even more effective treatments. ACKNOWLEDGMENTS

Conflict of interests: None.

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895

8

896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959

Translational Research - 2014

Koppe

REFERENCES

Q21

1. Lazo M, Hernaez R, Eberhardt MS, et al. Prevalence of nonalcoholic fatty liver disease in the United States: the Third National Health and Nutrition Examination Survey, 1988-1994. Am J Epidemiol 2013;178:38–45. 2. Welsh JA, Karpen S, Vos MB. Increasing prevalence of nonalcoholic fatty liver disease among United States adolescents, 1988-1994 to 2007-2010. J Pediatr 2013;162:496–500.e1. 3. Vernon G, Baranova A, Younossi ZM. Systematic review: the epidemiology and natural history of non-alcoholic fatty liver disease and non-alcoholic steatohepatitis in adults. Aliment Pharmacol Ther 2011;34:274–85. 4. Kleiner DE, Brunt EM. Nonalcoholic fatty liver disease: pathologic patterns and biopsy evaluation in clinical research. Semin Liver Dis 2012;32:3–13. 5. Matteoni CA, Younossi ZM, Gramlich T, Boparai N, Liu YC, McCullough AJ. Nonalcoholic fatty liver disease: a spectrum of clinical and pathological severity. Gastroenterology 1999; 116:1413–9. 6. Ekstedt M, Franzen LE, Mathiesen UL, et al. Long-term followup of patients with NAFLD and elevated liver enzymes. Hepatology 2006;44:865–73. 7. Powell EE, Jonsson JR, Clouston AD. Steatosis: co-factor in other liver diseases. Hepatology 2005;42:5–13. 8. Chalasani N, Younossi Z, Lavine JE, et al. The diagnosis and management of non-alcoholic fatty liver disease: practice Guideline by the American Association for the Study of Liver Diseases, American College of Gastroenterology, and the American Gastroenterological Association. Hepatology 2012;55: 2005–23. 9. Orman ES, Barritt AS 4th, Wheeler SB, Hayashi PH. Declining liver utilization for transplantation in the United States and the impact of donation after cardiac death. Liver Transpl 2013;19: 59–68. 10. Ovchinsky N, Lavine JE. A critical appraisal of advances in pediatric nonalcoholic fatty liver disease. Semin Liver Dis 2012;32:317–24. 11. Feldstein AE, Charatcharoenwitthaya P, Treeprasertsuk S, Benson JT, Enders FB, Angulo P. The natural history of nonalcoholic fatty liver disease in children: a follow-up study for up to 20 years. Gut 2009;58:1538–44. 12. Huang Y, He S, Li JZ, et al. A feed-forward loop amplifies nutritional regulation of PNPLA3. Proc Natl Acad Sci U S A 2010; 107:7892–7. 13. Dongiovanni P, Donati B, Fares R, et al. PNPLA3 I148M polymorphism and progressive liver disease. World J Gastroenterol 2013;19:6969–78. 14. Sookoian S, Pirola CJ. Meta-analysis of the influence of I148M variant of patatin-like phospholipase domain containing 3 gene (PNPLA3) on the susceptibility and histological severity of nonalcoholic fatty liver disease. Hepatology 2011; 53:1883–94. 15. Trepo E, Nahon P, Bontempi G, et al. Association between the PNPLA3 (rs738409 C.G) variant and hepatocellular carcinoma: evidence from a meta-analysis of individual participant data. Hepatology 2014;59:2170–7. 16. Romeo S, Kozlitina J, Xing C, et al. Genetic variation in PNPLA3 confers susceptibility to nonalcoholic fatty liver disease. Nat Genet 2008;40:1461–5. 17. Li JZ, Huang Y, Karaman R, et al. Chronic overexpression of PNPLA3I148M in mouse liver causes hepatic steatosis. J Clin Invest 2012;122:4130–44. 18. Finkenstedt A, Auer C, Glodny B, et al. Patatin-like phospholipase domain-containing protein 3 rs738409-G in recipients of

19.

20.

21.

22.

23. 24.

25. 26.

27.

28. 29.

30.

31.

32.

33.

34.

35.

36.

37.

liver transplants is a risk factor for graft steatosis. Clin Gastroenterol Hepatol 2013;11:1667–72. Peter A, Kantartzis K, Machicao F, et al. Visceral obesity modulates the impact of apolipoprotein C3 gene variants on liver fat content. Int J Obes 2012;36:774–82. Petersen KF, Dufour S, Hariri A, et al. Apolipoprotein C3 gene variants in nonalcoholic fatty liver disease. N Eng J Med 2010; 362:1082–9. Lee HY, Birkenfeld AL, Jornayvaz FR, et al. Apolipoprotein CIII overexpressing mice are predisposed to diet-induced hepatic steatosis and hepatic insulin resistance. Hepatology 2011;54: 1650–60. Al-Serri A, Anstee QM, Valenti L, et al. The SOD2 C47T polymorphism influences NAFLD fibrosis severity: evidence from case-control and intra-familial allele association studies. J Hepatol 2012;56:448–54. Naik A, Kosir R, Rozman D. Genomic aspects of NAFLD pathogenesis. Genomics 2013;102:84–95. Nielsen S, Guo Z, Johnson CM, Hensrud DD, Jensen MD. Splanchnic lipolysis in human obesity. J Clin Invest 2004;113: 1582–8. Choi SS, Diehl AM. Hepatic triglyceride synthesis and nonalcoholic fatty liver disease. Curr Opin Lipidol 2008;19:295–300. Yamaguchi K, Yang L, McCall S, et al. Inhibiting triglyceride synthesis improves hepatic steatosis but exacerbates liver damage and fibrosis in obese mice with nonalcoholic steatohepatitis. Hepatology 2007;45:1366–74. Jornayvaz FR, Shulman GI. Diacylglycerol activation of protein kinase Cepsilon and hepatic insulin resistance. Cell Metab 2012; 15:574–84. Puri P, Baillie RA, Wiest MM, et al. A lipidomic analysis of nonalcoholic fatty liver disease. Hepatology 2007;46:1081–90. Kumashiro N, Erion DM, Zhang D, et al. Cellular mechanism of insulin resistance in nonalcoholic fatty liver disease. Proc Natl Acad Sci U S A 2011;108:16381–5. Magkos F, Su X, Bradley D, et al. Intrahepatic diacylglycerol content is associated with hepatic insulin resistance in obese subjects. Gastroenterology 2012;142:1444–1446.e2. Jornayvaz FR, Birkenfeld AL, Jurczak MJ, et al. Hepatic insulin resistance in mice with hepatic overexpression of diacylglycerol acyltransferase 2. Proc Natl Acad Sci U S A 2011; 108:5748–52. Chocian G, Chabowski A, Zendzian-Piotrowska M, Harasim E, Lukaszuk B, Gorski J. High fat diet induces ceramide and sphingomyelin formation in rat’s liver nuclei. Mol Cell Biochem 2010;340:125–31. Yetukuri L, Katajamaa M, Medina-Gomez G, SeppanenLaakso T, Vidal-Puig A, Oresic M. Bioinformatics strategies for lipidomics analysis: characterization of obesity related hepatic steatosis. BMC Syst Biol 2007;1:12. Holland WL, Brozinick JT, Wang LP, et al. Inhibition of ceramide synthesis ameliorates glucocorticoid-, saturated-fat-, and obesity-induced insulin resistance. Cell Metab 2007;5:167–79. Yang G, Badeanlou L, Bielawski J, Roberts AJ, Hannun YA, Samad F. Central role of ceramide biosynthesis in body weight regulation, energy metabolism, and the metabolic syndrome. Am J Physiol Endocrinol Metab 2009;297:E211–24. Pagadala M, Kasumov T, McCullough AJ, Zein NN, Kirwan JP. Role of ceramides in nonalcoholic fatty liver disease. Trends Endocrinol Metab 2012;23:365–71. Holland WL, Bikman BT, Wang LP, et al. Lipid-induced insulin resistance mediated by the proinflammatory receptor TLR4 requires saturated fatty acid-induced ceramide biosynthesis in mice. J Clin Invest 2011;121:1858–70.

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023

Translational Research Volume -, Number 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087

Koppe

38. Sanyal AJ, Campbell-Sargent C, Mirshahi F, et al. Nonalcoholic steatohepatitis: association of insulin resistance and mitochondrial abnormalities. Gastroenterology 2001;120:1183–92. 39. Chalasani N, Gorski JC, Asghar MS, et al. Hepatic cytochrome P450 2E1 activity in nondiabetic patients with nonalcoholic steatohepatitis. Hepatology 2003;37:544–50. 40. Li Z, Yang S, Lin H, et al. Probiotics and antibodies to TNF inhibit inflammatory activity and improve nonalcoholic fatty liver disease. Hepatology 2003;37:343–50. 41. Hassanein T, Frederick T. Mitochondrial dysfunction in liver disease and organ transplantation. Mitochondrion 2004;4:609–20. 42. Pessayre D, Fromenty B. NASH: a mitochondrial disease. J Hepatol 2005;42:928–40. 43. Leung TM, Nieto N. CYP2E1 and oxidant stress in alcoholic and non-alcoholic fatty liver disease. J Hepatol 2013;58:395–8. 44. Aubert J, Begriche K, Knockaert L, Robin MA, Fromenty B. Increased expression of cytochrome P450 2E1 in nonalcoholic fatty liver disease: mechanisms and pathophysiological role. Clin Res Hepatol Gastroenterol 2011;35:630–7. 45. Puri P, Wiest MM, Cheung O, et al. The plasma lipidomic signature of nonalcoholic steatohepatitis. Hepatology 2009;50:1827–38. 46. Chalasani N, Deeg MA, Crabb DW. Systemic levels of lipid peroxidation and its metabolic and dietary correlates in patients with nonalcoholic steatohepatitis. Am J Gastroenterol 2004;99: 1497–502. 47. Feldstein AE, Lopez R, Tamimi TA, et al. Mass spectrometric profiling of oxidized lipid products in human nonalcoholic fatty liver disease and nonalcoholic steatohepatitis. J Lipid Res 2010; 51:3046–54. 48. Rolo AP, Teodoro JS, Palmeira CM. Role of oxidative stress in the pathogenesis of nonalcoholic steatohepatitis. Free Radic Biol Med 2012;52:59–69. 49. Ozcan U, Cao Q, Yilmaz E, et al. Endoplasmic reticulum stress links obesity, insulin action, and type 2 diabetes. Science 2004; 306:457–61. 50. Zhang K, Kaufman RJ. The unfolded protein response: a stress signaling pathway critical for health and disease. Neurology 2006;66(2 Suppl 1):S102–9. 51. Borradaile NM, Han X, Harp JD, Gale SE, Ory DS, Schaffer JE. Disruption of endoplasmic reticulum structure and integrity in lipotoxic cell death. J Lipid Res 2006;47:2726–37. 52. Spector AA, Yorek MA. Membrane lipid composition and cellular function. J Lipid Res 1985;26:1015–35. 53. Ariyama H, Kono N, Matsuda S, Inoue T, Arai H. Decrease in membrane phospholipid unsaturation induces unfolded protein response. J Biol Chem 2010;285:22027–35. 54. Gregor MF, Yang L, Fabbrini E, et al. Endoplasmic reticulum stress is reduced in tissues of obese subjects after weight loss. Diabetes 2009;58:693–700. 55. Puri P, Mirshahi F, Cheung O, et al. Activation and dysregulation of the unfolded protein response in nonalcoholic fatty liver disease. Gastroenterology 2008;134:568–76. 56. Sharma NK, Das SK, Mondal AK, et al. Endoplasmic reticulum stress markers are associated with obesity in nondiabetic subjects. J Clin Endocrinol Metab 2008;93:4532–41. 57. Hui JM, Hodge A, Farrell GC, Kench JG, Kriketos A, George J. Beyond insulin resistance in NASH: TNF-alpha or adiponectin? Hepatology 2004;40:46–54. 58. Polyzos SA, Toulis KA, Goulis DG, Zavos C, Kountouras J. Serum total adiponectin in nonalcoholic fatty liver disease: a systematic review and meta-analysis. Metabolism 2011;60:313–26. 59. Holland WL, Miller RA, Wang ZV, et al. Receptor-mediated activation of ceramidase activity initiates the pleiotropic actions of adiponectin. Nat Med 2011;17:55–63.

9

60. van der Poorten D, Samer CF, Ramezani-Moghadam M, et al. Hepatic fat loss in advanced nonalcoholic steatohepatitis: are alterations in serum adiponectin the cause? Hepatology 2013; 57:2180–8. 61. Kawai T, Akira S. The role of pattern-recognition receptors in innate immunity: update on toll-like receptors. Nat Immunol 2010;11:373–84. 62. Seki E, Brenner DA. Toll-like receptors and adaptor molecules in liver disease: update. Hepatology 2008;48:322–35. 63. Poulain-Godefroy O, Le Bacquer O, Plancq P, et al. Inflammatory role of toll-like receptors in human and murine adipose tissue. Mediators Inflamm 2010;2010:823486. 64. Soares JB, Pimentel-Nunes P, Roncon-Albuquerque R, LeiteMoreira A. The role of lipopolysaccharide/toll-like receptor 4 signaling in chronic liver diseases. Hepatol Int 2010;4: 659–72. 65. Schaeffler A, Gross P, Buettner R, et al. Fatty acid-induced induction of toll-like receptor-4/nuclear factor-kappaB pathway in adipocytes links nutritional signalling with innate immunity. Immunology 2009;126:233–45. 66. Shi H, Kokoeva MV, Inouye K, Tzameli I, Yin H, Flier JS. TLR4 links innate immunity and fatty acid-induced insulin resistance. J Clin Invest 2006;116:3015–25. 67. Wu GD, Chen J, Hoffmann C, et al. Linking long-term dietary patterns with gut microbial enterotypes. Science 2011;334: 105–8. 68. Mehal WZ. The Gordian Knot of dysbiosis, obesity and NAFLD. Nat Rev Gastroenterol Hepatol 2013;10:637–44. 69. Rabot S, Membrez M, Bruneau A, et al. Germ-free C57BL/6J mice are resistant to high-fat-diet-induced insulin resistance and have altered cholesterol metabolism. FASEB J 2010;24: 4948–59. 70. Le Roy T, Llopis M, Lepage P, et al. Intestinal microbiota determines development of non-alcoholic fatty liver disease in mice. Gut 2013;62:1787–94. 71. Vrieze A, Van Nood E, Holleman F, et al. Transfer of intestinal microbiota from lean donors increases insulin sensitivity in individuals with metabolic syndrome. Gastroenterology 2012;143: 913–916.e7. 72. Yoshimoto S, Loo TM, Atarashi K, et al. Obesity-induced gut microbial metabolite promotes liver cancer through senescence secretome. Nature 2013;499:97–101. 73. Dapito DH, Mencin A, Gwak GY, et al. Promotion of hepatocellular carcinoma by the intestinal microbiota and TLR4. Cancer Cell 2012;21:504–16. 74. Feldstein AE, Wieckowska A, Lopez AR, Liu YC, Zein NN, McCullough AJ. Cytokeratin-18 fragment levels as noninvasive biomarkers for nonalcoholic steatohepatitis: a multicenter validation study. Hepatology 2009;50:1072–8. 75. Cusi K, Chang Z, Harrison S, et al. Limited value of plasma cytokeratin-18 as a biomarker for NASH and fibrosis in patients with non-alcoholic fatty liver disease. J Hepatol 2014;60: 167–74. 76. Tanwar S, Trembling PM, Guha IN, et al. Validation of terminal peptide of procollagen III for the detection and assessment of nonalcoholic steatohepatitis in patients with nonalcoholic fatty liver disease. Hepatology 2013;57:103–11. 77. Kim D, Kim WR, Talwalkar JA, Kim HJ, Ehman RL. Advanced fibrosis in nonalcoholic fatty liver disease: noninvasive assessment with MR elastography. Radiology 2013;268:411–9. 78. Kumar R, Rastogi A, Sharma MK, et al. Liver stiffness measurements in patients with different stages of nonalcoholic fatty liver disease: diagnostic performance and clinicopathological correlation. Dig Dis Sci 2013;58:265–74.

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151

10

1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215

Translational Research - 2014

Koppe

79. Promrat K, Kleiner DE, Niemeier HM, et al. Randomized controlled trial testing the effects of weight loss on nonalcoholic steatohepatitis. Hepatology 2010;51:121–9. 80. Harrison SA, Fecht W, Brunt EM, Neuschwander-Tetri BA. Orlistat for overweight subjects with nonalcoholic steatohepatitis: a randomized, prospective trial. Hepatology 2009;49:80–6. 81. Mathurin P, Hollebecque A, Arnalsteen L, et al. Prospective study of the long-term effects of bariatric surgery on liver injury in patients without advanced disease. Gastroenterology 2009; 137:532–40. 82. Moschen AR, Molnar C, Wolf AM, et al. Effects of weight loss induced by bariatric surgery on hepatic adipocytokine expression. J Hepatol 2009;51:765–77. 83. Promrat K, Longato L, Wands JR, de la Monte SM. Weight loss amelioration of non-alcoholic steatohepatitis linked to shifts in hepatic ceramide expression and serum ceramide levels. Hepatol Res 2011;41:754–62. 84. Huang H, Kasumov T, Gatmaitan P, et al. Gastric bypass surgery reduces plasma ceramide subspecies and improves insulin sensitivity in severely obese patients. Obesity 2011;19:2235–40. 85. Haus JM, Solomon TP, Kelly KR, et al. Improved hepatic lipid composition following short-term exercise in nonalcoholic fatty liver disease. J Clin Endocrinol Metab 2013;98:E1181–8. 86. Fealy CE, Haus JM, Solomon TP, et al. Short-term exercise reduces markers of hepatocyte apoptosis in nonalcoholic fatty liver disease. J Appl Physiol 2012;113:1–6. 87. Kistler KD, Brunt EM, Clark JM, et al. Physical activity recommendations, exercise intensity, and histological severity of nonalcoholic fatty liver disease. Am J Gastroenterol 2011;106: 460–8. quiz 9. 88. Ainsworth BE, Haskell WL, Herrmann SD, et al. 2011 Compendium of Physical Activities: a second update of codes and MET values. Med Sci Sports Exerc 2011;43:1575–81. 89. Neuschwander-Tetri BA, Ford DA, Acharya S, et al. Dietary trans-fatty acid induced NASH is normalized following loss of trans-fatty acids from hepatic lipid pools. Lipids 2012;47: 941–50. 90. Koppe SW, Elias M, Moseley RH, Green RM. Trans fat feeding results in higher serum alanine aminotransferase and increased insulin resistance compared with a standard murine high-fat diet. Am J Physiol Gastrointest Liver Physiol 2009;297: G378–84. 91. McCarthy M. US moves to ban trans fats. Br Med J 2013;347: f6749. 92. Marriott BP, Cole N, Lee E. National estimates of dietary fructose intake increased from 1977 to 2004 in the United States. J Nutr 2009;139:1228S–35S. 93. Zelber-Sagi S, Nitzan-Kaluski D, Goldsmith R, et al. Long term nutritional intake and the risk for non-alcoholic fatty liver disease (NAFLD): a population based study. J Hepatol 2007;47: 711–7. 94. Stanhope KL, Schwarz JM, Keim NL, et al. Consuming fructosesweetened, not glucose-sweetened, beverages increases visceral adiposity and lipids and decreases insulin sensitivity in overweight/obese humans. J Clin Invest 2009;119:1322–34. 95. Lim JS, Mietus-Snyder M, Valente A, Schwarz JM, Lustig RH. The role of fructose in the pathogenesis of NAFLD and the metabolic syndrome. Nat Rev Gastroenterol Hepatol 2010;7: 251–64. 96. Samuel VT. Fructose induced lipogenesis: from sugar to fat to insulin resistance. Trends Endocrinol Metab 2011;22:60–5. 97. Freedman ND, Park Y, Abnet CC, Hollenbeck AR, Sinha R. Association of coffee drinking with total and cause-specific mortality. N Engl J Med 2012;366:1891–904.

98. O’Keefe JH, Bhatti SK, Patil HR, DiNicolantonio JJ, Lucan SC, Lavie CJ. Effects of habitual coffee consumption on cardiometabolic disease, cardiovascular health, and all-cause mortality. J Am Coll Cardiol 2013;62:1043–51. 99. van Dam RM, Hu FB. Coffee consumption and risk of type 2 diabetes: a systematic review. JAMA 2005;294:97–104. 100. Molloy JW, Calcagno CJ, Williams CD, Jones FJ, Torres DM, Harrison SA. Association of coffee and caffeine consumption with fatty liver disease, nonalcoholic steatohepatitis, and degree of hepatic fibrosis. Hepatology 2012;55:429–36. 101. Saab S, Mallam D, Cox Ii GA, Tong MJ. Impact of coffee on liver diseases: a systematic review. Liver Int 2014;34:495–504. 102. Bravi F, Bosetti C, Tavani A, Gallus S, La Vecchia C. Coffee reduces risk for hepatocellular carcinoma: an updated metaanalysis. Clin Gastroenterol Hepatol 2013;11:1413–1421.e1. 103. Dranoff JA, Feld JJ, Lavoie EG, Fausther M. How does coffee prevent liver fibrosis? Biological plausibility for recent epidemiological observations. Hepatology 2014. Q2 104. Chiang DJ, Roychowdhury S, Bush K, et al. Adenosine 2A receptor antagonist prevented and reversed liver fibrosis in a mouse model of ethanol-exacerbated liver fibrosis. PloS One 2013;8:e69114. 105. Sanyal AJ, Chalasani N, Kowdley KV, et al. Pioglitazone, vitamin E, or placebo for nonalcoholic steatohepatitis. N Engl J Med 2010;362:1675–85. 106. Pacana T, Sanyal AJ. Vitamin E and nonalcoholic fatty liver disease. Curr Opin Clin Nutr Metab Care 2012;15:641–8. 107. Ahmadian M, Suh JM, Hah N, et al. PPARgamma signaling and metabolism: the good, the bad and the future. Nat Med 2013;19: 557–66. 108. Savage DB, Tan GD, Acerini CL, et al. Human metabolic syndrome resulting from dominant-negative mutations in the nuclear receptor peroxisome proliferator-activated receptorgamma. Diabetes 2003;52:910–7. 109. Lavine JE, Schwimmer JB, Van Natta ML, et al. Effect of vitamin E or metformin for treatment of nonalcoholic fatty liver disease in children and adolescents: the TONIC randomized controlled trial. JAMA 2011;305:1659–68. 110. Balas B, Belfort R, Harrison SA, et al. Pioglitazone treatment increases whole body fat but not total body water in patients with non-alcoholic steatohepatitis. J Hepatol 2007;47:565–70. 111. Du J, Ma YY, Yu CH, Li YM. Effects of pentoxifylline on nonalcoholic fatty liver disease: a meta-analysis. World J Gastroenterol 2014;20:569–77. 112. Van Wagner LB, Koppe SW, Brunt EM, et al. Pentoxifylline for the treatment of non-alcoholic steatohepatitis: a randomized controlled trial. Ann Hepatol 2011;10:277–86. 113. Zein CO, Yerian LM, Gogate P, et al. Pentoxifylline improves nonalcoholic steatohepatitis: a randomized placebo-controlled trial. Hepatology 2011;54:1610–9. 114. Zein CO, Lopez R, Fu X, et al. Pentoxifylline decreases oxidized lipid products in nonalcoholic steatohepatitis: new evidence on the potential therapeutic mechanism. Hepatology 2012;56:1291–9. 115. Kreth S, Ledderose C, Luchting B, Weis F, Thiel M. Immunomodulatory properties of pentoxifylline are mediated via adenosine-dependent pathways. Shock 2010;34:10–6. 116. Konrad FM, Neudeck G, Vollmer I, Ngamsri KC, Thiel M, Reutershan J. Protective effects of pentoxifylline in pulmonary inflammation are adenosine receptor A2A dependent. FASEB J 2013;27:3524–35. 117. Cronstein BN. Adenosine receptors and fibrosis: a translational review. F1000 Biol Rep 2011;3:21. 118. Adorini L, Pruzanski M, Shapiro D. Farnesoid X receptor targeting to treat nonalcoholic steatohepatitis. Drug Discov Today 2012;17:988–97.

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241 1242 1243 1244 1245 1246 1247 1248 1249 1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279

Translational Research Volume -, Number 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332 1333 1334 1335 1336 1337 1338 1339 1340 1341 1342 1343

Koppe

119. Fuchs M. Non-alcoholic fatty liver disease: the bile acidactivated farnesoid x receptor as an emerging treatment target. J Lipids 2012;2012:934396. 120. Watanabe M, Houten SM, Wang L, et al. Bile acids lower triglyceride levels via a pathway involving FXR, SHP, and SREBP-1c. J Clin Invest 2004;113:1408–18. 121. Sirvent A, Claudel T, Martin G, et al. The farnesoid X receptor induces very low density lipoprotein receptor gene expression. FEBS Lett 2004;566:173–7. 122. Fiorucci S, Antonelli E, Rizzo G, et al. The nuclear receptor SHP mediates inhibition of hepatic stellate cells by FXR and protects against liver fibrosis. Gastroenterology 2004;127:1497–512. 123. Wang YD, Chen WD, Wang M, Yu D, Forman BM, Huang W. Farnesoid X receptor antagonizes nuclear factor kappaB in hepatic inflammatory response. Hepatology 2008;48:1632–43. 124. Cariou B, van Harmelen K, Duran-Sandoval D, et al. The farnesoid X receptor modulates adiposity and peripheral insulin sensitivity in mice. J Biol Chem 2006;281:11039–49. 125. Ma K, Saha PK, Chan L, Moore DD. Farnesoid X receptor is essential for normal glucose homeostasis. J Clin Invest 2006; 116:1102–9.

11

126. Ryan KK, Tremaroli V, Clemmensen C, et al. FXR is a molecular target for the effects of vertical sleeve gastrectomy. Nature 2014; 509:183–8. 127. Patti ME, Houten SM, Bianco AC, et al. Serum bile acids are higher in humans with prior gastric bypass: potential contribution to improved glucose and lipid metabolism. Obesity 2009; 17:1671–7. 128. Myronovych A, Kirby M, Ryan KK, et al. Vertical sleeve gastrectomy reduces hepatic steatosis while increasing serum bile acids in a weight-loss-independent manner. Obesity 2014;22: 390–400. 129. Mudaliar S, Henry RR, Sanyal AJ, 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. 130. Barry-Hamilton V, Spangler R, Marshall D, et al. Allosteric inhibition of lysyl oxidase-like-2 impedes the development of a pathologic microenvironment. Nat Med 2010;16:1009–17. 131. Nseir W, Mahamid M. Statins in nonalcoholic fatty liver disease and steatohepatitis: updated review. Curr Atheroscler Rep 2013; 15:305.

REV 5.2.0 DTD  TRSL803_proof  12 July 2014  7:04 pm  ce

1344 1345 1346 1347 1348 1349 1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364 1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 1379 1380 1381 1382 1383 1384 1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397 1398 1399 1400 1401 1402 1403 1404 1405 1406 1407

Obesity and the liver: nonalcoholic fatty liver disease.

The increasing prevalence of nonalcoholic fatty liver disease (NAFLD) parallels the rise of obesity and its complications. NAFLD is a common cause of ...
1MB Sizes 0 Downloads 13 Views