Gut Microbiota and Pathogenesis of Organ Injury (Advances in Experimental Medicine and Biology, 1238) 9811523843, 9789811523847

This book aims to introduce the latest research in gut microbiota by systematically summarizing how it modulates the pat

103 28 4MB

English Pages 211 [207] Year 2020

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Contents
Contributors
1 Introduction
Abstract
1.1 Overview of the Gut Microbiota
1.2 Overview of Organ Injury
1.3 The Gut Microbiota Influence the Pathogenesis of Organ Injury Development
References
2 Gut Microbiota and Alimentary Tract Injury
Abstract
2.1 Introduction
2.2 The Role of Helicobacter pylori Infection and Microbiome Dysbiosis in Chromic Gastritis and Gastric Cancer
2.3 Microbiome Dysbiosis and Inflammatory Bowel Disease
2.4 Irritable Bowel Syndrome
2.5 Colorectal Cancer
2.6 Future Aspects
References
3 Gut Microbiota and Liver Injury (I)—Acute Liver Injury
Abstract
3.1 Introduction
3.2 The Gut Microbiota and Drug-Induced Liver Injury
3.2.1 APAP-Induced Liver Injury
3.2.2 Tacrine-Induced Liver Injury
3.2.3 Diclofenac-Induced Liver Injury
3.3 The Gut Microbiota and D-Galactosamine-Induced Acute Liver Injury
3.4 The Gut Microbiota and Acute Alcohol-Induced Liver Injury
3.5 The Gut Microbiota and Viral Hepatitis-Related ALI
3.6 The Gut Microbiota and Transplantation Linked Liver Injury
3.7 Conclusion
References
4 Gut Microbiota and Liver Injury (II): Chronic Liver Injury
Abstract
4.1 Overview
4.2 Hepatitis B Virus (A DNA Virus)
4.3 Hepatitis C Virus (An RNA Virus)
4.4 Autoimmune Hepatitis
4.5 Nonalcoholic Fatty Liver Disease
4.6 Alcoholic Liver Disease
4.7 Cirrhosis
4.8 Hepatocellular Carcinoma
4.9 Conclusion
References
5 Gut Microbiota and Lung Injury
Abstract
5.1 Introduction
5.2 Structural and Functional Similarities and Differences Between the Gut and Lung
5.3 Coordinated Development of Microbiota in Gastrointestinal Tract (GIT) and Respiratory Tract
5.4 Factors of Intestinal Microbiota that Influence the Lung
5.4.1 Exogenous Factors
5.4.2 Endogenous Factors
5.4.2.1 Delivery Mode
5.5 The Mechanism of the Gut–Lung Axis
5.6 Intestinal Dysbiosis and Lung Disease
5.6.1 Asthma
5.6.2 COPD
5.6.3 Pneumonia
5.6.4 Cystic Fibrosis
5.6.5 Lung Cancer
5.7 Intestinal Microbiome-Associated Strategies for the Treatment and Prevention of Lung Disease
5.7.1 The Asthma and the Gut Microbiome
5.7.2 The Pneumonia and the Gut Microbiome
5.7.3 The Immunotherapy and the Gut Microbiome
5.8 Conclusion
References
6 Gut Microbiota and Neurologic Diseases and Injuries
Abstract
6.1 Introduction
6.2 Neurodevelopmental Disorder—Autism Spectrum Disorder
6.3 Neurodegenerative Disease—Parkinson’s Disease and Alzheimer’s Disease
6.4 Autoimmune Disease—Multiple Sclerosis
6.5 Neuropsychiatric Disorders—Anxiety and Depression
6.6 Acute Pathology—Traumatic Brain Injury and Stroke
6.7 Summary
References
7 Gut Microbiota and Renal Injury
Abstract
7.1 Introduction
7.2 Impact of CKD/ESRD on the Intestinal Microbiota
7.3 Impairment of the Gut Barrier in CKD
7.4 SCFA, a Beneficial Product of Intestinal Flora
7.5 Gut-Derived Uremic Toxins
7.6 Modulations of Intestinal Flora Disorder in CKD/ESRD
7.6.1 Dietary Interventions
7.6.2 Probiotics, Prebiotics and Synbiotics
7.7 Therapeutic Drugs for Microbiota Regulation
7.8 Summary and Perspectives
Acknowledgment
References
8 Gut Microbiota and Heart, Vascular Injury
Abstract
8.1 Introduction
8.2 The Interaction Between Host Metabolism and Gut Microbial Metabolites
8.2.1 TMAO
8.2.2 BAs
8.2.3 SCFAs
8.2.4 H2S
8.3 Gut Microbiota and Atherosclerosis
8.3.1 Gut Dysbiosis and Atherosclerosis
8.3.2 BAs and Atherosclerosis
8.3.3 TMA/TMAO and Atherosclerosis
8.4 Gut Microbiota and Hypertension
8.4.1 SCFAs and Hypertension
8.4.2 H2S and Hypertension
8.5 Gut Microbiota and Diabetes
8.5.1 BAs and Diabetes
8.5.2 SCFAs and Diabetes
8.5.3 H2S and Diabetes
8.5.4 TMAO and Diabetes
8.6 Gut Microbiota and Obesity
8.6.1 SCFAs and Obesity
8.6.2 BAs and Obesity
8.7 Gut Microbiota and Heart Failure
8.8 Therapeutic Potential of Gut Microbiota in CVD
8.8.1 Antibiotic, Probiotic, and Prebiotic Treatment in CVD
8.8.2 Fecal Microbiota Transplantation in CVD Treatment
8.8.3 Reshaping the Population of Gut Bacteria by Dietary Intervention and Other Therapies
8.9 Conclusion
8.10 Sources of Funding
References
9 Gut Microbiota and Endocrine Disorder
Abstract
9.1 Gut Microbiota
9.2 Endocrine Systems
9.2.1 Neuroendocrine System
9.2.2 Peripheral Endocrine System
9.2.3 Consequences of Endocrine Disorders
9.3 Gut Microbiota and the Endocrine System
9.3.1 Gut Microbiota and Neuroendocrine System Diseases
9.3.2 Gut Microbiota and Peripheral Endocrine System Diseases
9.4 Conclusions and Potential Therapies
References
10 Gut Microbiota and Immune Responses
Abstract
10.1 Introduction
10.2 Role of Gut Microbiota in Immune System Development and Differentiation
10.2.1 Gut Bacteria
10.2.2 Gut Virus
10.2.3 Gut Fungus
10.3 Role of Gut Microbiota in the Regulation of Immune Responses
10.4 Gut Microbiota and Immune Diseases
10.4.1 Autoimmune Disorders
10.4.2 Inflammatory Bowel Diseases
10.4.3 Type 1 Diabetes
10.4.4 Systemic Lupus Erythematosus
10.4.5 Rheumatoid Arthritis
10.4.6 Immunodeficiency Disorders
10.4.7 Cancer
10.5 Concluding Remarks
References
11 Gut Microbiota and Multiple Organ Dysfunction Syndrome (MODS)
Abstract
11.1 Sepsis
11.2 Gut Microbiota and Sepsis
11.3 Conclusion
References
Recommend Papers

Gut Microbiota and Pathogenesis of Organ Injury (Advances in Experimental Medicine and Biology, 1238)
 9811523843, 9789811523847

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Advances in Experimental Medicine and Biology 1238

Peng Chen   Editor

Gut Microbiota and Pathogenesis of Organ Injury

Advances in Experimental Medicine and Biology Volume 1238

Series Editors Wim E. Crusio, Institut de Neurosciences Cognitives et Intégratives d’Aquitaine, CNRS and University of Bordeaux UMR 5287, Pessac Cedex, France John D. Lambris, University of Pennsylvania, Philadelphia, PA, USA Heinfried H. Radeke, Institute of Pharmacology & Toxicology, Clinic of the Goethe University Frankfurt Main, Frankfurt am Main, Hessen, Germany Nima Rezaei, Research Center for Immunodeficiencies, Children’s Medical Center, Tehran University of Medical Sciences, Tehran, Iran

Advances in Experimental Medicine and Biology provides a platform for scientific contributions in the main disciplines of the biomedicine and the life sciences. This series publishes thematic volumes on contemporary research in the areas of microbiology, immunology, neurosciences, biochemistry, biomedical engineering, genetics, physiology, and cancer research. Covering emerging topics and techniques in basic and clinical science, it brings together clinicians and researchers from various fields. Advances in Experimental Medicine and Biology has been publishing exceptional works in the field for over 40 years, and is indexed in SCOPUS, Medline (PubMed), Journal Citation Reports/Science Edition, Science Citation Index Expanded (SciSearch, Web of Science), EMBASE, BIOSIS, Reaxys, EMBiology, the Chemical Abstracts Service (CAS), and Pathway Studio. 2018 Impact Factor: 2.126.

More information about this series at http://www.springer.com/series/5584

Peng Chen Editor

Gut Microbiota and Pathogenesis of Organ Injury

123

Editor Peng Chen Department of Pathophysiology Southern Medical University Guangzhou, Guangdong, China

ISSN 0065-2598 ISSN 2214-8019 (electronic) Advances in Experimental Medicine and Biology ISBN 978-981-15-2384-7 ISBN 978-981-15-2385-4 (eBook) https://doi.org/10.1007/978-981-15-2385-4 © Springer Nature Singapore Pte Ltd. 2020 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd. The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721, Singapore

Contents

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Peng Chen

1

2

Gut Microbiota and Alimentary Tract Injury . . . . . . . . . . . . . . . . . Ye Chen, Guangyan Wu, and Yongzhong Zhao

11

3

Gut Microbiota and Liver Injury (I)—Acute Liver Injury . . . . . . . Guangyan Wu, Sanda Win, Tin A. Than, Peng Chen, and Neil Kaplowitz

23

4

Gut Microbiota and Liver Injury (II): Chronic Liver Injury . . . . . Susan S. Baker and Robert D. Baker

39

5

Gut Microbiota and Lung Injury . . . . . . . . . . . . . . . . . . . . . . . . . . Ji-yang Tan, Yi-chun Tang, and Jie Huang

55

6

Gut Microbiota and Neurologic Diseases and Injuries . . . . . . . . . . T. Tyler Patterson and Ramesh Grandhi

73

7

Gut Microbiota and Renal Injury . . . . . . . . . . . . . . . . . . . . . . . . . . Lei Zhang, Wen Zhang, and Jing Nie

93

8

Gut Microbiota and Heart, Vascular Injury . . . . . . . . . . . . . . . . . . 107 Cheng Zeng and Hongmei Tan

9

Gut Microbiota and Endocrine Disorder . . . . . . . . . . . . . . . . . . . . 143 Rui Li, Yifan Li, Cui Li, Dongying Zheng, and Peng Chen

10 Gut Microbiota and Immune Responses . . . . . . . . . . . . . . . . . . . . . 165 Lijun Dong, Jingwen Xie, Youyi Wang, and Daming Zuo 11 Gut Microbiota and Multiple Organ Dysfunction Syndrome (MODS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195 Peng Chen and Timothy Billiar

v

Contributors

Susan S. Baker Department of Pediatrics, Jacobs School of Medicine and Biomedical Sciences, State University of New York at Buffalo, Buffalo, USA Robert D. Baker Department of Pediatrics, Jacobs School of Medicine and Biomedical Sciences, State University of New York at Buffalo, Buffalo, USA Timothy Billiar Department of Surgery, University of Pittsburgh, Pittsburgh, PA, USA Peng Chen Department of Pathophysiology, Guangdong Provincial Key Laboratory of Proteomics, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China Ye Chen Department of Gastroenterology, Nanfang Hospital, Southern Medical University, Guangzhou, Guangdong, China Lijun Dong The Fifth Affiliated Hospital, Southern Medical University, Guangzhou, China; Department of Immunology, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China Ramesh Grandhi Department of Neurosurgery, University of Utah School of Medicine, Salt Lake City, Utah, USA Jie Huang Guangdong Provincial Key Laboratory of Translational Medicine in Lung Cancer, Guangdong Lung Cancer Institute, Guangdong Provincial People’s Hospital and Guangdong Academy of Medical Sciences, Guangzhou, China Neil Kaplowitz USC Research Center for Liver Disease, Department of Medicine, Keck School of Medicine of USC, Los Angeles, CA, USA Cui Li Department of Pathophysiology, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China

vii

viii

Contributors

Rui Li Department of Pathophysiology, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China Yifan Li Department of Pathophysiology, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China Jing Nie State Key Laboratory of Organ Failure Research, National Clinical Research Center of Kidney Disease, Key Laboratory of Organ Failure Research (Ministry of Education), Division of Nephrology, Nanfang Hospital, Southern Medical University, Guangzhou, China Hongmei Tan Department of Pathophysiology, Zhongshan School of Medicine, Sun Yat-sen University, Guangzhou, China Ji-yang Tan Peking Union Medical College, Chinese Academy of Medical Sciences, Plastic Surgery Hospital, Graduate School of Peking Union Medical College, Beijing, China Yi-chun Tang The Second School of Clinical Medicine, Southern Medical University, Guangzhou, China Tin A. Than USC Research Center for Liver Disease, Department of Medicine, Keck School of Medicine of USC, Los Angeles, CA, USA T. Tyler Patterson Department of Neurosurgery, University of Texas Health Science Center School of Medicine, San Antonio, TX, USA Youyi Wang Department of Immunology, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China; School of Laboratory Medicine and Biotechnology, Institute of Molecular Immunology, Southern Medical University, Guangzhou, China Sanda Win USC Research Center for Liver Disease, Department of Medicine, Keck School of Medicine of USC, Los Angeles, CA, USA Guangyan Wu Department of Pathophysiology, Guangdong Provincial Key Laboratory of Proteomics, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China Jingwen Xie Department of Immunology, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China Cheng Zeng Department of Pathophysiology, Zhongshan School of Medicine, Sun Yat-sen University, Guangzhou, China Lei Zhang State Key Laboratory of Organ Failure Research, National Clinical Research Center of Kidney Disease, Key Laboratory of Organ Failure Research (Ministry of Education), Division of Nephrology, Nanfang Hospital, Southern Medical University, Guangzhou, China

Contributors

ix

Wen Zhang State Key Laboratory of Organ Failure Research, National Clinical Research Center of Kidney Disease, Key Laboratory of Organ Failure Research (Ministry of Education), Division of Nephrology, Nanfang Hospital, Southern Medical University, Guangzhou, China Yongzhong Zhao Department of Cellular and Molecular Medicine, Lerner Research Institute, Cleveland, OH, USA Dongying Zheng Department of Pathophysiology, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China Daming Zuo School of Laboratory Medicine and Biotechnology, Institute of Molecular Immunology, Southern Medical University, Guangzhou, China; Microbiome Medicine Center, Zhujiang Hospital, Southern Medical University, Guangzhou, China

Chapter 1

Introduction Peng Chen

Abstract In the past decade, research focusing on the gut microbiota has attracted a growing number of scientists. We increasingly realize that the gut microbiota plays a key role in both maintaining host homeostasis and affecting the progression of multiple diseases, which means that the microorganisms living in the intestines would not only influence the gastrointestinal tract but also impact other important organs such as the liver, brain, kidney, and lung. The underlying modulatory mechanism is complicated; however, we can expand the existing insight on the development of organ damage pathogenesis and discover novel therapeutic targets for organ injury-related diseases by investigating this “hot-topic”. In this chapter, we will present a broad overview of the gut microbiota and organ injury, as well as the latest research achievements regarding the linkage between them. Keywords Gut microbiota

1.1

 Organ injury  Pathogenesis

Overview of the Gut Microbiota

There are approximately 100 trillion microbes living in our body, mainly residing in the gastrointestinal tract, skin, oral cavity, and vagina [1]. Of these habitats, the microbes in gastrointestinal tract comprise approximately 80% of the total microbiota. To support their survival, the gut microbiota also encodes many more unique genes than the host genome, and they generate more than 1000 metabolites [2, 3]. The distribution of the gut microbiota varies throughout the gastrointestinal tract. In particular, the total microbial load is approximately 103–107/ml in the upper gastrointestinal tract (stomach and small intestine), while in the cecum and colon, the number of total microbes increases up to 1012/ml [4]. This variation could be P. Chen (&) Department of Pathophysiology, Guangdong Provincial Key Laboratory of Proteomics, School of Basic Medical Sciences, Southern Medical University, N.No 1838 Guangzhou Ave., Guangzhou 510515, China e-mail: [email protected] © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_1

1

2

P. Chen

explained by multiple potential mechanisms; for example, gastric juices could penetrate into the small intestine and inhibit bacterial growth [5]. Bile acids derived from the liver are also able to restrict bacterial overgrowth in the small intestine [6]. Besides bacteria, other microorganisms such as viruses and fungi are also present in the gut [7]; however, the majority of gut microbiota are bacteria. Over 99% of the bacteria in the gut are anaerobes, which is consistent with the low-oxygen environment in the intestines [8]. Gut microbiota is not only located in the lumen, but they also exist in the mucosal layer and intestinal epithelial cells [9]. To assess the diversity of species in the bacterial communities of certain habitats, such as the intestine, alpha-diversity, and beta-diversity measurements are employed. Alpha-diversity is used to evaluate the richness of each sample [10]. It normally includes four parameters: Chao1 index, Shannon index, Simpson index, and Observed species (Table 1.1). Beta-diversity is employed to evaluate the compositional difference between different samples. Principal component analysis (PCA) and principal coordinates analysis (PCoA) are widely used to display the outcomes of beta-diversity. The composition signature of the intestinal bacteria in humans was disclosed recently. Scientists found that the overall composition of intestinal bacteria is completely different from the bacteria located in other habitats of the body [11]. Specifically at the phylum level, Firmicutes and Bacteroidetes are the predominant bacteria, comprising approximately 90% of all the bacteria in the gut. Other phyla, including Proteobacteria, Actinobacteria, Verrucomicrobia, Cyanobacteria, Fusobacteria, and Spirochaetes, are also present in the intestines. Notably, the entire gut flora should be recognized as “commensal” microbes, even though a number of potential pathogens such as Helicobacter pylori and Escherichia coli could be detected in healthy individuals [12]. The composition of gut microbiota varies dynamically; namely, the composition is different at two different time points in the same individual. It is also shifted among different persons. However, the within-subject variation over time is more stable than the between-subject variation. The gut microbiota could be affected by many endogenous and exogenous factors. Host intestinal alpha-defense secretions, as well as immunologic activity,

Table 1.1 Introduction of alpha-diversity Significance Observed species Chao1 index Shannon index

Simpson index

Represents community operational taxonomic units (OTUs) numbers The larger the value, the richer is the species in the sample Represents community richness The higher the value, the higher is the richness of the communities Represents community richness and equiability The larger the value, the higher is the sample community equiability in the sample Represents community richness and equiability The larger the Simpson index, the higher is the community equiability in the sample

1 Introduction

3

could impact the microbial balance in the gut [13, 14]. Besides, reports have revealed that the mode of delivery determines the “inoculation” of our gut microbiota [15]. The gut microbiota of vaginally delivered infants showed significantly more resemblance to the mothers’ gut microbiota than that of infants delivered by C-section, indicating that the mother-to-infant transmission was compromised in the C-section. Specifically, C-section markedly delayed the richness of Bifidobacterium in the gut, compared to vaginal delivery [16]. Breastfeeding is also more helpful than formula feeding for the gut microbiota to mature into adult-like microbiota [17]. The composition of gut microbiota was also influenced by geography. People from the USA have completely different microbial communities than Malawians [18]. Diet is another key determining factor in the shaping of gut microbiota. For example, a short-term plant-based diet versus an animal-based diet could rapidly alter the gut microbial composition in humans [19]. Other processes such as aging and exercise are also able to change the bacterial communities in the gut [20, 21]. Collectively, the composition and function of the gut microbiota could be influenced by multiple factors. Owing to the development of sequencing technology, as well as other “omics” analyses, great advances have been made in the past decade on the study of gut microbiota. Overall, the methods used in microbiota research include: 1. The 16S rRNA gene sequencing and diversity analysis. The 16S ribosomal RNA is the key component of the ribosome in prokaryotes [22]. Around 30 years ago, scientists started to use the 16S rDNA gene sequencing for phylogenetic studies [23]. The 16S rRNA gene contains nine variable regions (V1–V9) and ten conservative regions. The conservative regions normally exhibit a similar structure among the different bacterial strains, while variable regions display different structures depending on the strain. Hence, we can amplify the variable regions using specific primers in the conservative regions and sequence the PCR products to analyze the bacterial composition. The V3 and V4 regions are widely used for sequencing nowadays. The outcome of 16S ribosomal RNA gene sequencing contains alpha-diversity, beta-diversity, and the relative compositions of specific strains from the phylum to the genus levels. 2. Metagenomics. Generally, metagenomics is defined as the study of the whole genetic material obtained from the intestine and other bacterial habitats. Unlike the 16S rRNA gene sequencing and diversity analysis, metagenomics analyzes the whole genome. At present, it is mainly executed by two sequencing methods: shotgun sequencing and next-generation high-throughput sequencing. Metagenomics would reveal much more information than the 16S rRNA gene sequencing and diversity analysis. It not only discloses information about the bacterial composition but also predicts the genes’ functions, and further provides a functional comparison between different samples based on public databases such as the Kyoto Encyclopedia of Genes and Genomes (KEGG). Thus, one main usage of metagenomics analysis in biomedical research is to explore the differences in the function of metabolism-associated genes among various pathophysiological states. Since the database is not fully completed, the

4

P. Chen

functions of many sequenced genes are still unknown; thus, future research would improve our knowledge of gene function predictions and amplify the application of metagenomics. 3. Metabolomics. Gut microbial metabolomics analysis is recognized as the measurement of all the microbial metabolites. It can directly reflect the microbial function. Briefly, it can be divided into two types: untargeted metabolomics and targeted metabolomics. The former searches all the compounds detected in the system, while targeted metabolomics focuses on the specific type of molecule. The commonly used methods for metabolomics analysis are mass spectrometry (MS, like GC-MS, LC-MS) and nuclear magnetic resonance (NMR) spectroscopy-based techniques. With the development of metabolomics, we are closer to understanding the pathogenesis of many important diseases, since gut microbial metabolites are believed to not only support bacterial survival but also be involved in the progression of host diseases, such as digestive and cardiovascular diseases. Similar to metagenomics, many unknown or undetectable molecules must be explored in future. 4. Metaproteomics. Bacterial-derived protein is an important indicator for microbial function and plays an important role in the host’s pathophysiological changes. Metaproteomics is recognized as the study of all the proteins generated from environmental sources such as the gut microbiota. Previously, metaproteomics analyses were performed using two-dimensional polyacrylamide gel electrophoresis (2D-PAGE), but recent studies have been performed using mass spectrometry-based techniques, yielding higher accuracy and throughput. This approach could reveal the abundance of all the proteins existing in the system and other important information such as post-translational modifications (PTM). It would also be able to identify novel peptides. Hence, it is a powerful method to help us investigate the gut microbial function and their impact on host homeostasis. Besides the techniques mentioned above, more attention is being given to bacterial cultures in vitro, although the majority intestinal bacteria are currently unculturable [24]. Obtaining a single strain is helpful for us to understand their detailed function and for the basic and translational biomedical research on gut microbiota.

1.2

Overview of Organ Injury

As a main topic in biomedical research, organ injury is widely studied and many mediators have been revealed. Generally, organ injury is the phenomenon that results from cell death or dysfunction, and it could negatively influence organ function. The spectrum of organ injury may be broad, ranging from mild reversible degeneration, such as steatosis, to end-stage organ failure, such as organ cirrhosis,

1 Introduction

5

and it also comprises acute phase and chronic phase processes. The effectiveness of the current therapeutic approach for organ injury treatment is still limited. The pathogenesis of organ injury is complex, but it can be classified into three general points: 1. Immune responses. The host immune system can be classified into two types: innate immune system and adaptive immune system. Both of these could elicit injury. For example, there are a large number of macrophages in the lungs called pulmonary macrophages. This cell population is responsible for the clearance of dust and other potential pathogens; however, upon infection, they are over-activated and exert phenotypic alterations, as characterized by cytokine over-expression and secretion [25]. Over-produced cytokines such as TNF-a would further bind their receptor located in the surface of the pneumonocyte and trigger apoptotic signaling transduction, followed by lung cell dysfunction and death. Such inflammatory response is also the main contributor to other organ injury processes. Besides the innate immune reaction mentioned above, the adaptive immune reaction is another key contributor to injury. For example, T cells are responsible for destroying the infected cells; during infection with the hepatitis virus, if the T cells are activated and start to clear the virus-infected hepatocytes, large amounts of liver cells become damaged, resulting in organ dysfunction [26]. The immune reaction is believed to be the most important mediator in the development of organ injury. 2. Oxidative stress. Redox balance is important in maintaining normal cellular function. Upon disruption, the imbalanced oxidative status could lead to cell damage and finally cause organ injury [27, 28]. For example, during renal ischemia-reperfusion injury, large amounts of hypoxanthine, metabolized from ATP and xanthine oxidase, are accumulated at the hypoxia phase; when oxygen is present after reperfusion, xanthine oxidase can catalyze hypoxanthine into xanthine and uric acid, which is accompanied by the generation of superoxide radicals. Superoxide would directly or indirectly destroy the function of biomacromolecules, resulting in disruption of the cells’ functions [29]. This radical would also be released from the cell and exert harmful effects onto other cells. Beside superoxide, other radicals like hydroxyl radicals, hydrogen peroxide, and reactive nitrogen species would also be harmful to the host cells and organs when they become abnormally accumulated. 3. Circulatory defect. Our circulation system enables all of our organs and cells to obtain enough oxygen and nutrients so that they can “work” smoothly to support our life. However, when the circulation is negatively influenced, the organ may be injured because of hypoxia. For example, sepsis-induced multiple organ dysfunction is the main cause of death [30]. Sepsis development is normally associated with a dysfunction in coagulation, which leads to microcirculation defects. Consequently, the hypoxia would augment the cell damage and promote the development of organ injury. Hence, it is widely recognized that improving the circulation is an important approach to alleviating septic organ injury.

6

P. Chen

Collectively, organ injury may be influenced by many endogenous or exogenous factors. Studying the pathogenesis and developing more effective treatment methods are leading topics in medical research.

1.3

The Gut Microbiota Influence the Pathogenesis of Organ Injury Development

An increasing number of scientists believe that the gut microbiota is involved in both intestinal and extraintestinal organ injury development based on recent strong evidence. Regardless of acute or chronic processes, the gut microbiota is able to mediate the progression or severity of organ damage in response to multiple etiologic factors. It is expected that the gut microbiota could regulate gastrointestinal injury. Helicobacter pylori is recognized as a key risk factor for gastric ulcers and cancer development [31]. Intestinal abnormality induced by infection with Clostridium difficile is a common syndrome in gastroenterology [32]. Inflammatory bowel disease (IBD) is closely associated with gut microbial dysbiosis [33]. It has been reported that the diversity of microbiota is lower in IBD patients than in healthy individuals. Firmicutes, Bacteroidetes, Actinobacteria, and Proteobacteria are all altered during IBD development. Notably, Enterobacteriaceae is markedly enriched in IBD patients. The dysbiotic status is believed to promote IBD; however, the detailed mechanism is complex. The levels of intestinal short-chain fatty acids (SCFAs), which serve as the beneficial nutrients for intestinal epithelial cells (IECs), are disrupted by dysbiosis [34]. Additionally, shifted microbial composition and function could directly disrupt host immune responses, causing harmful mucosal inflammation [35]. Besides the alimentary tract, the liver is the most connected organ that tightly communicates with intestinal bacteria. This results from (1) the primary bile acids generated by hepatocytes, which enter the intestine and regulate the bacterial composition, or the bacteria that could metabolize these compounds into secondary bile acids, which would be reabsorbed into liver and react with liver cells to influence hepatic function [36]; (2) the gut microbial products could penetrate into the circulatory system and reach the liver through the portal vein to modulate the pathophysiologic function of the liver. Hence, the gut microbiota has a deep crosstalk with the liver and can directly influence liver damage development, including both acute and chronic injury [37]. For examples, enteric dysbiosis is observed in both alcoholic and nonalcoholic fatty liver disease, and some important strains such as Akkermansia directly influence these chronic liver injury developments [38, 39]. Moreover, gut microbial metabolites could directly modulate drug-induced acute liver injury [40, 41]. Many scientists believe that some liver diseases are actually intestinal disease, especially because of the gut microbial abnormalities.

1 Introduction

7

Gut microbiota has also been demonstrated to be able to regulate the injury of other remote organs in response to multiple challenges, such as renal injury, cardiovascular injury, lung injury, and psychological disorder [42–45]. We previously discovered that gut microbial compositional disorder was linked to chronic high salt intake, and such microbial disruption could promote early renal injury [35]. Microbial metabolite indoxyl-sulfate is recognized as a key driver for chronic kidney disease (CKD) development [46]. Trimethylamine (TMA) that is generated from choline by intestinal bacteria could be further metabolized into trimethylamine oxide (TMAO), which causes atherosclerosis [47]. Hypertension is reported to be closely related to enteric dysbiosis, and similar to other diseases, the dysbiotic microbiota in hypertension patients could independently enhance their blood pressure [48]. Moreover, autism is associated with microbial disturbances, and treatment of Bacteroides fragilis could serve as an efficient approach to restore the gut barrier dysfunction, microbial changes, and behavioral abnormalities during the progression of autism spectrum disorders [49]. Moreover, the gut microbiota could regulate sepsis-induced multiple organ dysfunction [50]. We recently found that the susceptibility of mice to sepsis-induced multiple organ damage was modulated by the gut microbiota, specifically by the microbial metabolite granisetron, which could reduce inflammatory responses [50]. In addition, gut microbial-generated lipids may activate intestinal immune cells such as the natural killer T cells (NKT), and these important immune cells can further translocate to remote organs to exacerbate organ injury [51]. Hence, it is clear that gut microbiota plays a key role in the development of multiple organ damage. In this book, we will summarize the latest research regarding the modulatory effects of gut microbiota on organ injury and conclude with the underlying mechanisms. Also, targeting the gut microbiota is recognized as a novel effective approach to combat diseases involved in multiple organ damage.

References 1. Ley RE, Peterson DA, Gordon JI (2006) Ecological and evolutionary forces shaping microbial diversity in the human intestine. Cell 124(4):837–848 2. Medini D, Serruto D, Parkhill J, Relman DA, Donati C, Moxon R, Falkow S, Rappuoli R (2008) Microbiology in the post-genomic era. Nat Rev Microbiol 6(6):419–430 3. Sonnenburg ED, Sonnenburg JL (2014) Starving our microbial self: the deleterious consequences of a diet deficient in microbiota-accessible carbohydrates. Cell Metab 20 (5):779–786 4. Mowat AM, Agace WW (2014) Regional specialization within the intestinal immune system. Nat Rev Immunol 14(10):667–685 5. Martinsen TC, Bergh K, Waldum HL (2005) Gastric juice: a barrier against infectious diseases. Basic Clin Pharmacol Toxicol 96(2):94–102 6. Kurdi P, Kawanishi K, Mizutani K, Yokota A (2006) Mechanism of growth inhibition by free bile acids in lactobacilli and bifidobacteria. J Bacteriol 188(5):1979–1986 7. Clemente JC, Ursell LK, Parfrey LW, Knight R (2012) The impact of the gut microbiota on human health: an integrative view. Cell 148(6):1258–1270

8

P. Chen

8. Vedantam G, Hecht DW (2003) Antibiotics and anaerobes of gut origin. Curr Opin Microbiol 6(5):457–461 9. Atarashi K, Tanoue T, Ando M, Kamada N, Nagano Y, Narushima S, Suda W, Imaoka A, Setoyama H, Nagamori T, Ishikawa E, Shima T, Hara T, Kado S, Jinnohara T, Ohno H, Kondo T, Toyooka K, Watanabe E, Yokoyama S, Tokoro S, Mori H, Noguchi Y, Morita H, Ivanov II, Sugiyama T, Nuñez G, Camp JG, Hattori M, Umesaki Y, Honda K (2015) Th17 cell induction by adhesion of microbes to intestinal epithelial cells. Cell 163(2):367–380 10. Clarke A, Lidgard S (2000) Spatial patterns of diversity in the sea: bryozoan species richness in the North Atlantic. J Anim Ecol 69(5):799–814 11. Human Microbiome Project Consortium (2012) Structure, function and diversity of the healthy human microbiome. Nature 486(7402):207–214 12. Cover TL, Blaser MJ (2009) Helicobacter pylori in health and disease. Gastroenterology 136 (6):1863–1873 13. Cerf-Bensussan N, Gaboriau-Routhiau V (2010) The immune system and the gut microbiota: friends or foes? Nat Rev Immunol 10(10):735–744 14. Hooper LV, Littman DR, Macpherson AJ (2012) Interactions between the microbiota and the immune system. Science 336(6086):1268–1273 15. Salminen S, Gibson GR, McCartney AL, Isolauri E (2004) Influence of mode of delivery on gut microbiota composition in seven year old children. Gut 53(9):1388–1389 16. Grönlund MM, Lehtonen OP, Eerola E, Kero P (1999) Fecal microflora in healthy infants born by different methods of delivery: permanent changes in intestinal flora after cesarean delivery. J Pediatr Gastroenterol Nutr 28(1):19–25 17. Bäckhed F, Roswall J, Peng Y, Feng Q, Jia H, Kovatcheva-Datchary P, Li Y, Xia Y, Xie H, Zhong H, Khan MT, Zhang J, Li J, Xiao L, Al-Aama J, Zhang D, Lee YS, Kotowska D, Colding C, Tremaroli V, Yin Y, Bergman S, Xu X, Madsen L, Kristiansen K, Dahlgren J, Wang J (2015) Dynamics and stabilization of the human gut microbiome during the first year of life. Cell Host Microbe 17(5):690–703 18. Yatsunenko T, Rey FE, Manary MJ, Trehan I, Dominguez-Bello MG, Contreras M, Magris M, Hidalgo G, Baldassano RN, Anokhin AP, Heath AC, Warner B, Reeder J, Kuczynski J, Caporaso JG, Lozupone CA, Lauber C, Clemente JC, Knights D, Knight R, Gordon JI (2012) Human gut microbiome viewed across age and geography. Nature 486 (7402):222–227 19. David LA, Maurice CF, Carmody RN, Gootenberg DB, Button JE, Wolfe BE, Ling AV, Devlin AS, Varma Y, Fischbach MA, Biddinger SB, Dutton RJ, Turnbaugh PJ (2014) Diet rapidly and reproducibly alters the human gut microbiome. Nature 505(7484):559–563 20. Saraswati S, Sitaraman R (2015) Aging and the human gut microbiota-from correlation to causality. Front Microbiol 5:764 21. Lambert JE, Myslicki JP, Bomhof MR, Belke DD, Shearer J, Reimer RA (2015) Exercise training modifies gut microbiota in normal and diabetic mice. Appl Physiol Nutr Metab 40 (7):749–752 22. Weisburg WG, Barns SM, Pelletier DA, Lane DJ (1991) 16S ribosomal DNA amplification for phylogenetic study. J Bacteriol 173(2):697–703 23. Ward DM, Weller R, Bateson MM (1990) 16S rRNA sequences reveal numerous uncultured microorganisms in a natural community. Nature 345(6270):63–65 24. Flint HJ, O’Toole PW, Walker AW (2010) Special issue: the human intestinal microbiota. Microbiology 156(Pt 11):3203–3204 25. Wang H, Ma S (2008) The cytokine storm and factors determining the sequence and severity of organ dysfunction in multiple organ dysfunction syndrome. Am J Emerg Med 26(6):711– 715 26. Gehring AJ, Xue SA, Ho ZZ, Teoh D, Ruedl C, Chia A, Koh S, Lim SG, Maini MK, Stauss H, Bertoletti A (2011) Engineering virus-specific T cells that target HBV infected hepatocytes and hepatocellular carcinoma cell lines. J Hepatol 55(1):103–110 27. Chow CW, Herrera Abreu MT, Suzuki T, Downey GP (2003) Oxidative stress and acute lung injury. Am J Respir Cell Mol Biol 29(4):427–431

1 Introduction

9

28. Sener G, Toklu H, Ercan F, Erkanli G (2005) Protective effect of beta-glucan against oxidative organ injury in a rat model of sepsis. Int Immunopharmacol 5(9):1387–1396 29. Ihara N, Schmitz S, Kurisawa M, Chung JE, Uyama H, Kobayashi S (2004) Amplification of inhibitory activity of catechin against disease-related enzymes by conjugation on poly (epsilon-lysine). Biomacromolecules 5(5):1633–1636 30. Vincent JL, Moreno R, Takala J, Willatts S, De Mendonça A, Bruining H, Reinhart CK, Suter PM, Thijs LG (1996) The SOFA (Sepsis-related Organ Failure Assessment) score to describe organ dysfunction/failure. On behalf of the Working Group on Sepsis-Related Problems of the European Society of Intensive Care Medicine. Intensive Care Med 22 (7):707–710 31. Amieva M, Peek RM Jr (2016) Pathobiology of helicobacter pylori-induced gastric cancer. Gastroenterology 150(1):64–78 32. Banaei N, Anikst V, Schroeder LF (2015) Burden of Clostridium difficile infection in the United States. N Engl J Med 372(24):2368–2369 33. Kaur N, Chen CC, Luther J, Kao JY (2011) Intestinal dysbiosis in inflammatory bowel disease. Gut Microbes 2(4):211–216 34. Rodríguez-Carrio J, López P, Sánchez B, González S, Gueimonde M, Margolles A, de Los Reyes-Gavilán CG, Suárez A (2017) Intestinal dysbiosis is associated with altered short-chain fatty acids and serum-free fatty acids in systemic lupus erythematosus. Front Immunol 8:23 35. Hu J, Luo H, Wang J, Tang W, Lu J, Wu S, Xiong Z, Yang G, Chen Z, Lan T, Zhou H, Nie J, Jiang Y, Chen P (2017) Enteric dysbiosis-linked gut barrier disruption triggers early renal injury induced by chronic high salt feeding in mice. Exp Mol Med 49(8):e370 36. Merrick MV, Eastwood MA, Anderson JR, Ross HM (1982) Enterohepatic circulation in man of a gamma-emitting bile-acid conjugate, 23-selena-25-homotaurocholic acid (SeHCAT). J Nucl Med 23(2):126–130 37. Chassaing B, Etienne-Mesmin L, Gewirtz AT (2014) Microbiota-liver axis in hepatic disease. Hepatology 59(1):328–339 38. Wu W, Lv L, Shi D, Ye J, Fang D, Guo F, Li Y, He X, Li L (2017) Protective effect of Akkermansia muciniphila against immune-mediated liver injury in a mouse model. Front Microbiol 8:1804 39. Grander C, Adolph TE, Wieser V, Lowe P, Wrzosek L, Gyongyosi B, Ward DV, Grabherr F, Gerner RR, Pfister A, Enrich B, Ciocan D, Macheiner S, Mayr L, Drach M, Moser P, Moschen AR, Perlemuter G, Szabo G, Cassard AM, Tilg H (2018) Recovery of ethanol-induced Akkermansia muciniphila depletion ameliorates alcoholic liver disease. Gut 67(5):891–901 40. Gong S, Lan T, Zeng L, Luo H, Yang X, Li N, Chen X, Liu Z, Li R, Win S, Liu S, Zhou H, Schnabl B, Jiang Y, Kaplowitz N, Chen P (2018) Gut microbiota mediates diurnal variation of acetaminophen induced acute liver injury in mice. J Hepatol 69(1):51–59 41. Clayton TA, Baker D, Lindon JC, Everett JR, Nicholson JK (2009) Pharmacometabonomic identification of a significant host-microbiome metabolic interaction affecting human drug metabolism. Proc Natl Acad Sci U S A 106(34):14728–14733 42. Gong J, Noel S, Pluznick JL, Hamad ARA, Rabb H (2019) Gut Microbiota-kidney cross-talk in acute kidney injury. Semin Nephrol 39(1):107–116 43. Tang WH, Kitai T, Hazen SL (2017) Gut microbiota in cardiovascular health and disease. Circ Res 120(7):1183–1196 44. Kapur R, Kim M, Rebetz J, Hallström B, Björkman JT, Takabe-French A, Kim N, Liu J, Shanmugabhavananthan S, Milosevic S, McVey MJ, Speck ER, Semple JW (2018) Gastrointestinal microbiota contributes to the development of murine transfusion-related acute lung injury. Blood Adv. 2(13):1651–1663 45. Cryan JF, O’Riordan KJ, Cowan CSM, Sandhu KV, Bastiaanssen TFS, Boehme M, Codagnone MG, Cussotto S, Fulling C, Golubeva AV, Guzzetta KE, Jaggar M, Long-Smith CM, Lyte JM, Martin JA, Molinero-Perez A, Moloney G, Morelli E, Morillas E, O’Connor R, Cruz-Pereira JS, Peterson VL, Rea K, Ritz NL, Sherwin E, Spichak S, Teichman EM, van de

10

46.

47.

48.

49.

50.

51.

P. Chen Wouw M, Ventura-Silva AP, Wallace-Fitzsimons SE, Hyland N, Clarke G, Dinan TG (2019) The microbiota-gut-brain axis. Physiol Rev 99(4):1877–2013 Barreto FC, Barreto DV, Liabeuf S, Meert N, Glorieux G, Temmar M, Choukroun G, Vanholder R, Massy ZA, European Uremic Toxin Work Group (EUTox) (2009) Serum indoxyl sulfate is associated with vascular disease and mortality in chronic kidney disease patients. Clin J Am Soc Nephrol 4(10):1551–1558 Bennett BJ, de Aguiar Vallim TQ, Wang Z, Shih DM, Meng Y, Gregory J, Allayee H, Lee R, Graham M, Crooke R, Edwards PA, Hazen SL, Lusis AJ (2013) Trimethylamine-N-oxide, a metabolite associated with atherosclerosis, exhibits complex genetic and dietary regulation. Cell Metab 17(1):49–60 Li J, Zhao F, Wang Y, Chen J, Tao J, Tian G, Wu S, Liu W, Cui Q, Geng B, Zhang W, Weldon R, Auguste K, Yang L, Liu X, Chen L, Yang X, Zhu B, Cai J (2017) Gut microbiota dysbiosis contributes to the development of hypertension. Microbiome 5(1):14 Hsiao EY, McBride SW, Hsien S, Sharon G, Hyde ER, McCue T, Codelli JA, Chow J, Reisman SE, Petrosino JF, Patterson PH, Mazmanian SK (2013) Microbiota modulate behavioral and physiological abnormalities associated with neurodevelopmental disorders. Cell 155(7):1451–1463 Gong S, Yan Z, Liu Z, Niu M, Fang H, Li N, Huang C, Li L, Chen G, Luo H, Chen X, Zhou H, Hu J, Yang W, Huang Q, Schnabl B, Chang P, Billiar TR, Jiang Y, Chen P (2019) Intestinal microbiota mediates the susceptibility to polymicrobial sepsis-induced liver injury by granisetron generation in mice. Hepatology 69(4):1751–1767 Lee KC, Chen P, Maricic I, Inamine T, Hu J, Gong S, Sun JC, Dasgupta S, Lin HC, Lin YT, Loomba R, Stärkel P, Kumar V, Schnabl B (2019) Intestinal iNKT cells migrate to liver and contribute to hepatocyte apoptosis during alcoholic liver disease. Am J Physiol Gastrointest Liver Physiol 316(5):G585–G597

Chapter 2

Gut Microbiota and Alimentary Tract Injury Ye Chen, Guangyan Wu, and Yongzhong Zhao

Abstract The gastrointestinal (GI) tract is inhabited by a diverse array of microbes, which play crucial roles in health and disease. Dysbiosis of microbiota has been tightly linked to gastrointestinal inflammatory and malignant diseases. Here we highlight the role of Helicobacter pylori alongside gastric microbiota associated with gastric inflammation and cancer. We summarize the taxonomic and functional aspects of intestinal microbiota linked to inflammatory bowel diseases (IBD), irritable bowel syndrome (IBS), and colorectal cancer in clinical investigations. We also discuss microbiome-related animal models. Nevertheless, there are tremendous opportunities to reveal the causality of microbiota in health and disease and detailed microbe–host interaction mechanisms by which how dysbiosis is causally linked to inflammatory disease and cancer, in turn, potentializing clinical interventions with a personalized high efficacy.









Keywords Microbiota Microbiome Chronic gastritis Gastric cancer Inflammatory bowel disease Irritable bowel syndrome Colorectal cancer Dysbiosis







Y. Chen Department of Gastroenterology, Nanfang Hospital, Southern Medical University, Guangzhou, Guangdong, China e-mail: [email protected] G. Wu Department of Pathophysiology, School of Basic Medical Sciences, Southern Medical University, Guangzhou, China Y. Zhao (&) Department of Cellular and Molecular Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, OH, USA e-mail: [email protected] © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_2

11

12

2.1

Y. Chen et al.

Introduction

The alimentary tract comprises the oral cavity, pharynx, esophagus, stomach (fundus and antrum), duodenum (Du), jejunum (Je), ileum, cecum (Ce), colon (ascending colon, transverse colon, and descending colon), sigmoid colon (SC), and rectum (Re). Owing to this consistent basic structure, the gastrointestinal (GI) tract is colonized by a diverse array of microbiota. The complicated bi-directional relationship between the GI microbiota and host underscores the crucial role of dysbiosis (a microbial imbalance or maladaptation on or inside the body, such as an impaired microbiota) in complex disease etiology [1–5]. Here we review advances on studying gut microbiota in GI diseases, focusing on Helicobacter pylori (H. pylori) infection on the upper GI tract, bacterial dysbiosis in inflammatory bowel disease (IBD), colorectal cancer (CRC), and irritable bowel syndrome (IBS).

2.2

The Role of Helicobacter pylori Infection and Microbiome Dysbiosis in Chromic Gastritis and Gastric Cancer

Chronic gastritis, one of the most common serious pandemic infections with even more than half of people infected worldwide, is a clinical condition due to an inflammation of gastric mucosa with multistep, progressive, and life-long process, which can progress into peptic ulcer or even gastric cancer [6]. The typical symptoms of chronic gastritis patients comprise upper abdominal pain, indigestion, bloating, nausea, vomiting, belching, loss of appetite, and weight loss. Many different conditions and factors cause or contribute to the development of chronic gastritis, including bacteria (bacterial gastritis, H. pylori), bile reflux, HIV/AIDS, and celiac disease. Chronic gastritis is linked to gastric cancer, which is the fifth leading cause of cancer and the third leading cause of death from cancer, making up 7% of case and 9% of deaths [7, 8]. In 1982, Australian scientists Barry Marshall and Robin Warren identified that H. pylori was present in a person with chronic gastritis and gastric ulcers [9]. H. pylori belongs to Campylobacter pylori, a germ-negative, microaerophilic bacterium generally found in the stomach as an essential risk factor in 65–80% of gastric cancers, but only 2% of people with H. pylori infections develop stomach cancer [10]. Given that the disease-etiology paradigm has been shifting to germ-organ theory, in which human microbiota is the essential component of health and disease beyond the canonical Koch’s postulates, gastric microbiota is not an exception [1, 7]. Thus, as anticipated, the bacterial diversity has altered in the present with H. pylori infection, following the criteria of gastric mucosal microbiome dysbiosis definition [11–13]. Taxonomic changes have been revealed, including blooming of Escherichia Shigella, Burkholderia of the Proteobacteria

2

Gut Microbiota and Alimentary Tract Injury

13

phylum, Lactobacillus, Lachnospiraceae, Streptococcus, and Veillonella of the Firmicutes phylum, and Prevotella of the Bacteroidetes phylum [14, 15]. Taxonomic composition and diversity of microbial communities in the stomach are associated with the development of gastric cancer [11–19]; however, functional impact is anticipated to play critical molecular role, such as urea and lipopolysaccharide and the encoding biosynthetic clusters [20–24]. Of particular note, cagA encoded in H. pylori is the first identified bacterial oncogene [25]. Moreover, H. pylori could have a beneficial effect to human health, indicating a promising complexity of studying the H. pylori-centered gastric microbiome [26].

2.3

Microbiome Dysbiosis and Inflammatory Bowel Disease

IBD with a complex disease etiology comprises ulcerative colitis (UC) and Crohn’s disease (CD), resulting from dysregulated mucosal immune response to commensal intestinal microbiota and epithelium, which usually happens in genetically susceptible individuals [27]. UC is characterized by diffuse inflammation restricted to the mucosal layer of the colon that causes recurrent symptoms, while the inflammation of CD can be typically transmural, patchy, and segmental, occurring in any part of the GI tract [28]. The prevalence of IBD is more common in developed than in developing nations as reported that an estimate of 1.8 million (0.9%) U.S. adults in 1999 and 3.1 million (1.3%) U.S. adults in 2015 were diagnosed of IBD, with a rising disease rate in developing nations, including China and India [29–34]. Given that IBD is a complex genetics-microbiome-environment disease with a myriad of studies published [28, 35], here we focus on microbiome dysbiosis and IBD, including significant discoveries with 16S DNA sequencing, metaproteomics, metagenome shot-gun sequencing, integrative human microbiome project (HMP2) [36], as well as fecal microbiota transplantation and mouse models. Microbiome can be measured by 16S ribosomal DNA sequencing, albeit with the limitation of resolution at the genus level [37]. Reduced alpha-diversity has been consistently observed, including reduced Firmicutes abundance and decreased strict anaerobic bacteria abundance and increased facultative anaerobic proteobacteria [38–40]. Blakeley-Ruiz et al. found that gut microbiota was sustained, and revealed phylum-redundant metabolic functional stability in CD patients in the year after resection surgery [41]. Besides bacteria, fungi also play critical role in the pathogenesis of IBD, albeit with a great potential to clarify due to conflicted reports. Fungi Saccharomyces cerevisiae and Candida albicans increased with CD, particularly in setting of more serious bacterial dysbiosis, while increased Candida albicans alongside decreased Saccharomyces cerevisiae was observed in patients with IBD [38, 42–46].

14

Y. Chen et al.

IBD is fundamentally caused by the aberrant host–microbe interactions. Thus, microbiome-related host response is pivotal to the pathogenesis of IBD. Dextran sodium sulfate (DSS), a chemical reagent that can damage the integrity of the epithelial barrier, is an extensively induced experimental colitis. With the development of intestinal organoid culture systems, the modular gut inflammation-on-a-chip provides a research tool for DSS-induced colitis models. Shin et al. have provided significant experimental evidence that the integrity of the epithelial barrier is necessary to maintain the normal intercellular host–microbiome cross-talk [47]. In a cohort of sigmoid colon biopsies of active UC, 29 core proteins were reduced, including MUC2, a secretory mucin glycoprotein as one of the major structural components with the function of protecting the host against the pathogenic bacterial invasion in the mucus layer [48]. The Predicting Response to Standardized Pediatric Colitis Therapy (PROTECT) study for UC patients showed that severity and treatment response gene signatures of mucosal transcriptomes are associated with changes in flora involved in mucosal homeostasis [49]. Mucosa-associated bacterial species, such as Burkholderia cepacia, Flavonifractor plautii, and Rumminococcus sp., were more prone to be recognized selectively by pediatric IBD patient-derived IgG at intestinal mucosal surface compared with non-IBD [50, 51]. Exclusive enteral nutrition (EEN) appears to modulate the gut microbiome, for example, the CD Treatment-with-EATing diet (CD-TREAT), an individualized, food-based diet-enabled replicating EEN changes in the inflammatory gut microbiome [52–54]. Since FMT was broadly reported as an effective and safe treatment for recurrent or refractory Clostridium difficile infection, it has attracted increasing research attention in FMT as a potential therapeutic option in other diseases related to the gut microbiome, such as the IBD. In recent randomized controlled trials, Costello et al. revealed that patients with mild-to-moderate UC received treatment with anaerobically prepared donor FMT compared to patients who received autologous stool; the former was more effective in achieving clinical remission at 8 weeks [55]. According to the transcriptomic and microbiome data, Hyams et al. revealed that the gut microbiome before treatment with standardized protocols was associated with achieving clinical remission in the PROTECT study. For example, patients with lower relative abundance of Sutterella and higher relative abundance of Ruminococcaceae were more prone to 52-week corticosteroid-free remission [56]. We summarized the known microbiome pathophysiology of IBD, as shown in Fig. 2.1. It appears that microbiome lays the foundation of pathogenesis of IBD, thereby leading to potential interventions, including synthetic ecology approach of integrating prebiotics and engineered probiotics [57]. Meanwhile, the host–microbe interactions are pivotal in the disease mechanism of IBD.

2

Gut Microbiota and Alimentary Tract Injury

15

Fig. 2.1 Altered microbiome-linked pathogenesis features of IBD. Nine aspects have been illustrated, with an emphasis of the immune mechanism of host–microbe interaction

2.4

Irritable Bowel Syndrome

IBS, a functional GI disorder, is characterized by recurrent abdominal pain related to defecation and associated with a change in frequency and/or form of stool, with diarrhea or constipation, or mixed symptoms of diarrhea and constipation, according to the Rome IV diagnostic criteria. IBS is one of the most common diagnoses in functional GI disorder associated with the dysbiosis of gut microbiota [58, 59]. A variety of gut microbial signatures have been generated, for example, relative abundance of Clostridiales and Prevotella and Bacteroides based on 16S ribosomal DNA sequencing, being claimed to have the capacity to discriminate the IBS symptom severity and healthy, however, highly controversial [60–63]. It is most likely that the poor replicability is due to the pathogenesis of IBS at the strain level, that is, those available data had not reached the right resolution. Distinct intestinal fungal dysbiosis was observed in patients with IBS [64–66], as well as well-established rat maternal separation model [64–66]. By combining fecal calprotectin and the top 20 selected taxonomies, it led to highest prediction accuracy in distinguishing patients with IBD from those with IBS [64–66]. A fermentable oligo-, di-, monosaccharides, and polyols (FODMAPs) restricted diet (low FODMAPs diet) has been reported to ameliorate symptoms in IBS patients [67]. Yet, previous studies demonstrated that the low FODMAPs diet could lead to altered gut microbiota, such as the reductions of probiotic Bifidobacterium

16

Y. Chen et al.

[68–85]. In a randomized controlled trial, it showed that the low FODMAPs diet was more effective in IBS patient group compared with the placebo diet. And co-administration of the multistrain probiotic and low FODMAPs diet could restore Bifidobacterium [86]. Bennet et al. demonstrated that change in Bifidobacterium in low FODMAPs diet was related to the lactose consumption [66]. Future research needs to explore whether a combination of low FODMAPs, probiotic, and prebiotic will play an even greater role in predicting and modulating gut microbiota for IBS patients. The first randomized, double-blind, placebo-controlled trial aimed at exploring the effect of FMT in patients with IBS was implemented by Johnsen and colleagues. They reported that IBS patients in the active treatment group treated with FMT via colonoscopy-induced effective symptom relief at 3 months post-FMT, yet less remarkable at 12 months post-FMT compared with those in placebo group [87]. In addition, unlike delivered to the lower-gut through colonoscopy, FMT via oral capsules is a form of upper-gut delivery method. Halkjær et al. reported that patients with IBS orally administered FMT capsules were exhibited changes to the gut microbiota, however, those individuals after 3 months have poor effects in symptom improvement compared to patient treatment with placebo [88, 89]. Therefore, there is an urge to determine whether FMT protocol is related to the outcome of patients with IBS.

2.5

Colorectal Cancer

Colorectal cancer (CRC) is a common malignancy, including colon cancer and rectal cancer. According to the global cancer statistics in 2018, CRC ranks third in terms of incidence but second in the matter of mortality among estimated cases and deaths for males and females [90]. Several species of bacteria and bacterial metabolites have been investigated in CRC. For example, Fusobacterium nucleatum, a causal opportunistic infections pathogen, has been implicated in CRC. Several systematic reviews have emphatically described that the intestinal microbiota Fusobacterium nucleatum, colibactin-producing Escherichia coli, and enterotoxigenic Bacteroides fragilis were involved in DNA damage and tumor progression in CRC [91, 92]. In a large-scale, long-term retrospective study, Kwong et al. reported that the risk of CRC was enhanced in patients with bacteremia from specific intestinal microbes, including Bacteroides fragilis, Streptococcus gallolyticus, Fusobacterium nucleatum, Clostridium septicum, Clostridium perfringens, and Gemella morbillorum [93]. Predictive microbiome signature could facilitate the early intervention of CRC. Thus, a set of signatures have been revealed [94–98]. In a large cohort of 616 participants who underwent colonoscopy to assess taxonomic and functional characteristics of gut microbiota and metabolites, the relative abundance of Fusobacterium nucleatum spp. was significantly (P < 0.005) elevated continuously from intramucosal carcinoma to more advanced stages, and Atopobium parvulum

2

Gut Microbiota and Alimentary Tract Injury

17

and Actinomyces odontolyticus, which co-occurred in intramucosal carcinomas, were significantly (P < 0.005) increased only in multiple polypoid adenomas and/ or intramucosal carcinomas [94]. A meta-analysis of eight geographically and technically diverse fecal shotgun metagenomic studies of 768 cases of CRC identified a core set of 29 species significantly enriched in CRC metagenomes (false discovery rate (FDR) < 1  10(−5)), with that functional analysis of CRC metagenomes revealed enriched protein and mucin catabolism genes and depleted carbohydrate degradation genes [95]. On two additional cohorts with total 969 fecal metagenomes, the gut microbiome in CRC showed reproducibly higher richness than controls partially due to expansion of species typically derived from the oral cavity, with identification of gluconeogenesis and the putrefaction and fermentation pathways as being associated with CRC [96]. Nevertheless, the microbial molecules and/or pathways contribute to the pathogenic drivers of CRC that is still in its infancy at this stage. The intestinal microbiota also consists of both fungi and viruses, yet their role in the contribution of CRC is still largely unknown. Commensal gut fungi are one of the most important components of the gut microbiota. Coker et al. reported that fecal mycobiome is a predictive biomarker for the diagnosis of CRC with independent bacteria and other clinical parameters, and disruptions are involved in colorectal carcinogenesis [99]. Targeting one of the immune sensing pathways associated with fungi, Malik et al. reported that gut fungi through SYK-CARD9 pathway enhanced the inflammasome activation and IL-18 maturation, and then promoted epithelial barrier repair and IFN-c production, which was protected from colitis and CRC [100]. In another study, aiming at the gut virome, Nakatsu and colleagues first reported that the diversity of bacteriophage virome was increased in CRC patients, and the dysbiosis of the enteric virome was related to CRC prognosis [101]. These findings collectively indicated the potential effectiveness of gut fungi and virome in the development of CRC. Recent study has revealed that activating transcription factor 6 (ATF6) in intestinal epithelial cells contributes to microbial dysbiosis, and promotes microbiota-dependent colorectal tumorigenesis in mice [102]. It indicated that interruption of the ATF6 signaling pathways and modulation of the gut microbiome to reverse dysbiosis may provide the precise therapeutic targets in CRC. This study underscores the critical role of the host side in the pathogenesis of IBD. In addition, Cremonesi et al. reported that CRC cells were recognized as an essential chemokine source, expressing a range of chemokines. And specific gut microbiota promoted production of chemotactic factors in CRC cells and recruited more T cells into tumor sites, and thereby improved the prognosis of CRC [99]. He et al. recently reported that Campylobacter jejuni, a strain that produces cytolethal distending toxin (CDT), could disturb both host and microbial genes expression, and finally promotes tumorigenesis in mice [103]. This novel piece of evidence strengthened the tight association between gut microbiota and CRC development.

18

2.6

Y. Chen et al.

Future Aspects

Gut microbiota study in common gastrointestinal complex diseases is still in its early-infancy, looking forward to discovery of mechanistic predictive microbiome biomarkers and host–microbe interaction mechanisms. In addition to examining the replicability of association studies in clinical investigations, the future focus will lie in the causality approaches, such as randomized controlled trials, gnotobiotic rodent models, mediation analysis, as well as Mendelian randomization [104, 105]. We envision that the future research will be focusing on the host–microbe interface, or treating host and microbiome as a whole, namely holobiont [106, 107]. Under the germ-organ theory framework, novel individualized clinical interventions, such as targeting the microbiome and the host–microbe interface, are emerging in the arena of fighting those devastating diseases, including IBD and CRC.

References 1. Byndloss MX, Baumler AJ (2018) The germ-organ theory of non-communicable diseases. Nat Rev Microbiol 16(2):103–110 2. Lamont RJ, Koo H, Hajishengallis G (2018) The oral microbiota: dynamic communities and host interactions. Nat Rev Microbiol 16(12):745–759 3. Marchesi J, Shanahan F (2007) The normal intestinal microbiota. Curr Opin Infect Dis 20 (5):508–513 4. Stecher B, Maier L, Hardt WD (2013) ‘Blooming’ in the gut: how dysbiosis might contribute to pathogen evolution. Nat Rev Microbiol 11(4):277–284 5. Tuddenham S, Sears CL (2015) The intestinal microbiome and health. Curr Opin Infect Dis 28(5):464–470 6. Sipponen P, Maaroos HI (2015) Chronic gastritis. Scand J Gastroenterol 50(6):657–667 7. Karimi P et al (2014) Gastric cancer: descriptive epidemiology, risk factors, screening, and prevention. Cancer Epidemiol Biomarkers Prev 23(5):700–713 8. Thrift AP, El-Serag HB (2020) Burden of gastric cancer. Clin Gastroenterol Hepatol 18 (3):534–542 9. Marshall BJ (1988) The Campylobacter pylori story. Scand J Gastroenterol Suppl 146:58–66 10. Crowe SE (2019) Helicobacter pylori infection. N Engl J Med 380(12):1158–1165 11. Wang L et al (2016) Bacterial overgrowth and diversification of microbiota in gastric cancer. Eur J Gastroenterol Hepatol 28(3):261–266 12. Liu X et al (2019) Alterations of gastric mucosal microbiota across different stomach microhabitats in a cohort of 276 patients with gastric cancer. EBioMedicine 40:336–348 13. Coker OO et al (2018) Mucosal microbiome dysbiosis in gastric carcinogenesis. Gut 67 (6):1024–1032 14. Zmora N et al (2018) Personalized gut mucosal colonization resistance to empiric probiotics is associated with unique host and microbiome features. Cell 174(6):1388–1405 e21 15. Hold GL, Hansen R (2019) Impact of the gastrointestinal microbiome in health and disease: co-evolution with the host immune system. Curr Top Microbiol Immunol 421:303–318 16. Zhang C et al (2017) The gastric microbiome and its influence on gastric carcinogenesis: current knowledge and ongoing research. Hematol Oncol Clin North Am 31(3):389–408 17. Yu G et al (2014) Association between upper digestive tract microbiota and cancer-predisposing states in the esophagus and stomach. Cancer Epidemiol Biomarkers Prev 23(5):735–741

2

Gut Microbiota and Alimentary Tract Injury

19

18. Tan MP et al (2007) Chronic Helicobacter pylori infection does not significantly alter the microbiota of the murine stomach. Appl Environ Microbiol 73(3):1010–1013 19. Merrifield DL et al (2011) Effect of dietary alginic acid on juvenile tilapia (Oreochromis niloticus) intestinal microbial balance, intestinal histology and growth performance. Cell Tissue Res 344(1):135–146 20. Marshall BJ, Langton SR (1986) Urea hydrolysis in patients with Campylobacter pyloridis infection. Lancet 1(8487):965–966 21. Marshall BJ et al (1990) Urea protects Helicobacter (Campylobacter) pylori from the bactericidal effect of acid. Gastroenterology 99(3):697–702 22. Li H et al (2017) The redefinition of Helicobacter pylori lipopolysaccharide O-antigen and core-oligosaccharide domains. PLoS Pathog 13(3):e1006280 23. Li H et al (2018) Lipopolysaccharide structural differences between western and Asian Helicobacter pylori strains. Toxins (Basel) 10(9) 24. Li H et al (2016) Lipopolysaccharide structure and biosynthesis in Helicobacter pylori. Helicobacter 21(6):445–461 25. Backert S, Blaser MJ (2016) The role of CagA in the gastric biology of Helicobacter pylori. Cancer Res 76(14):4028–4031 26. Blaser MJ (2010) Helicobacter pylori and esophageal disease: wake-up call? Gastroenterology 139(6):1819–1822 27. Abraham C, Cho JH (2009) Inflammatory bowel disease. N Engl J Med 361(21):2066–2078 28. Ashton JJ, Beattie RM (2019) Personalised therapy for inflammatory bowel disease. Lancet 393(10182):1672–1674 29. Sood A et al (2003) Incidence and prevalence of ulcerative colitis in Punjab. North India. Gut 52(11):1587–1590 30. Zhao J et al (2013) First prospective, population-based inflammatory bowel disease incidence study in mainland of China: the emergence of “western” disease. Inflamm Bowel Dis 19(9):1839–1845 31. Peery AF et al (2019) Burden and cost of gastrointestinal, liver, and pancreatic diseases in the United States: update 2018. Gastroenterology 156(1):254–272 e11 32. Nguyen GC, Chong CA, Chong RY (2014) National estimates of the burden of inflammatory bowel disease among racial and ethnic groups in the United States. J Crohns Colitis 8(4):288–295 33. Dias CC et al (2019) Hospitalization trends of the inflammatory bowel disease landscape: a nationwide overview of 16 years. Dig Liver Dis 51(7):952–960 34. Furlow B (2019) The challenge of inflammatory bowel disease in India. Lancet Gastroenterol Hepatol 4(9):670–671 35. Baumgart DC, Sandborn WJ (2007) Inflammatory bowel disease: clinical aspects and established and evolving therapies. Lancet 369(9573):1641–1657 36. Lloyd-Price J et al (2019) Multi-omics of the gut microbial ecosystem in inflammatory bowel diseases. Nature 569(7758):655–662 37. Morgan XC et al (2012) Dysfunction of the intestinal microbiome in inflammatory bowel disease and treatment. Genome Biol 13(9):R79 38. Pascal V et al (2017) A microbial signature for Crohn’s disease. Gut 66(5):813–822 39. Ott SJ et al (2004) Reduction in diversity of the colonic mucosa associated bacterial microflora in patients with active inflammatory bowel disease. Gut 53(5):685–693 40. Manichanh C et al (2006) Reduced diversity of faecal microbiota in Crohn’s disease revealed by a metagenomic approach. Gut 55(2):205–211 41. Blakeley-Ruiz JA et al (2019) Metaproteomics reveals persistent and phylum-redundant metabolic functional stability in adult human gut microbiomes of Crohn’s remission patients despite temporal variations in microbial taxa, genomes, and proteomes. Microbiome 7(1):18 42. Sokol H et al (2017) Fungal microbiota dysbiosis in IBD. Gut 66(6):1039–1048 43. Singh V et al (2019) Microbiota fermentation-NLRP3 axis shapes the impact of dietary fibres on intestinal inflammation. Gut 68(10):1801–1812

20

Y. Chen et al.

44. Sabino J et al (2016) Primary sclerosing cholangitis is characterised by intestinal dysbiosis independent from IBD. Gut 65(10):1681–1689 45. Lemoinne S et al (2019) Fungi participate in the dysbiosis of gut microbiota in patients with primary sclerosing cholangitis. Gut 69(1):92–102 46. Imhann F et al (2018) Interplay of host genetics and gut microbiota underlying the onset and clinical presentation of inflammatory bowel disease. Gut 67(1):108–119 47. Shin W, Kim HJ (2018) Intestinal barrier dysfunction orchestrates the onset of inflammatory host-microbiome cross-talk in a human gut inflammation-on-a-chip. Proc Natl Acad Sci U S A 115(45):E10539–E10547 48. van der Post S et al (2019) Structural weakening of the colonic mucus barrier is an early event in ulcerative colitis pathogenesis. Gut 68(12):2142–2151 49. Haberman Y et al (2019) Ulcerative colitis mucosal transcriptomes reveal mitochondriopathy and personalized mechanisms underlying disease severity and treatment response. Nat Commun 10(1):38 50. Armstrong H et al (2019) Host immunoglobulin G selectively identifies pathobionts in pediatric inflammatory bowel diseases. Microbiome 7(1):1 51. Palm NW et al (2014) Immunoglobulin A coating identifies colitogenic bacteria in inflammatory bowel disease. Cell 158(5):1000–1010 52. Svolos V et al (2019) Treatment of active Crohn’s disease with an ordinary food-based diet that replicates exclusive enteral nutrition. Gastroenterology 156(5):1354–1367 e6 53. Ashton JJ, Beattie RM (2019) Treatment of active Crohn’s disease with an ordinary food-based diet that replicates exclusive enteral nutrition. Gastroenterology 157(4):1160– 1161 54. Quince C et al (2015) Extensive modulation of the fecal metagenome in children with Crohn’s disease during exclusive enteral nutrition. Am J Gastroenterol 110(12):1718–29; quiz 1730 55. Costello SP et al (2019) Effect of fecal microbiota transplantation on 8-week remission in patients with ulcerative colitis: a randomized clinical trial. JAMA 321(2):156–164 56. Hyams JS et al (2019) Clinical and biological predictors of response to standardised paediatric colitis therapy (PROTECT): a multicentre inception cohort study. Lancet 393 (10182):1708–1720 57. Cleary B et al (2017) Efficient generation of transcriptomic profiles by random composite measurements. Cell 171(6):1424–1436 e18 58. Drossman DA (2016) Functional gastrointestinal disorders: history, pathophysiology, clinical features and Rome IV. Gastroenterology pii: S0016-5085(16)00223-7 59. Ford AC, Lacy BE, Talley NJ (2017) Irritable bowel syndrome. N Engl J Med 376 (26):2566–2578 60. Precup G, Vodnar DC (2019) Gut Prevotella as a possible biomarker of diet and its eubiotic versus dysbiotic roles: a comprehensive literature review. Br J Nutr 122(2):131–140 61. Peter J et al (2018) A microbial signature of Psychological distress in irritable bowel syndrome. Psychosom Med 80(8):698–709 62. Hugerth LW et al (2019) No distinct microbiome signature of irritable bowel syndrome found in a Swedish random population. pii: gutjnl-2019-318717 63. Collins SM (2016) The intestinal microbiota in the irritable bowel syndrome. Int Rev Neurobiol 131:247–261 64. Mandarano AH et al (2018) Eukaryotes in the gut microbiota in myalgic encephalomyelitis/ chronic fatigue syndrome. PeerJ 6:e4282 65. Gu Y et al (2019) The potential role of gut mycobiome in irritable bowel syndrome. Front Microbiol 10:1894 66. Bennet SMP et al (2018) Multivariate modelling of faecal bacterial profiles of patients with IBS predicts responsiveness to a diet low in FODMAPs. Gut 67(5):872–881 67. Staudacher HM, Whelan K (2017) The low FODMAP diet: recent advances in understanding its mechanisms and efficacy in IBS. Gut 66(8):1517–1527

2

Gut Microbiota and Alimentary Tract Injury

21

68. Zhou SY et al (2018) FODMAP diet modulates visceral nociception by lipopolysaccharidemediated intestinal inflammation and barrier dysfunction. J Clin Invest 128(1):267–280 69. Wilder-Smith CH et al (2018) Breath methane concentrations and markers of obesity in patients with functional gastrointestinal disorders. United European Gastroenterol J 6(4): 595–603 70. Wilder-Smith CH et al (2018) Fermentable sugar ingestion, gas production, and gastrointestinal and central nervous system symptoms in patients with functional disorders. Gastroenterology 155(4):1034–1044 e6 71. Valeur J et al (2018) Exploring gut microbiota composition as an indicator of clinical response to dietary FODMAP restriction in patients with irritable bowel syndrome. Dig Dis Sci 63(2):429–436 72. Vakil N (2018) Dietary fermentable oligosaccharides, disaccharides, monosaccharides, and polyols (FODMAPs) and gastrointestinal disease. Nutr Clin Pract 33(4):468–475 73. Tuck CJ et al (2019) The impact of dietary fermentable carbohydrates on a postinflammatory model of irritable bowel syndrome. Neurogastroenterol Motil 31(10):e13675 74. Su H et al (2019) Effects of low-FODMAPS diet on irritable bowel syndrome symptoms and gut microbiome. Gastroenterol Nurs 42(2):150–158 75. Sloan TJ et al (2018) A low FODMAP diet is associated with changes in the microbiota and reduction in breath hydrogen but not colonic volume in healthy subjects. PLoS ONE 13(7): e0201410 76. Rodino-Janeiro BK et al (2018) A review of microbiota and irritable bowel syndrome: future in therapies. Adv Ther 35(3):289–310 77. Rinninella E et al (2019) Food components and dietary habits: keys for a healthy gut microbiota composition. Nutrients 11(10) 78. Rej A, Sanders DS (2018) Gluten-free diet and its ‘cousins’ in irritable bowel syndrome. Nutrients 10(11) 79. Rej A et al (2019) The role of diet in irritable bowel syndrome: implications for dietary advice. J Intern Med 286(5):490–502 80. Rej A et al (2018) Clinical application of dietary therapies in irritable bowel syndrome. J Gastrointestin Liver Dis 27(3):307–316 81. Panacer K, Whorwell PJ (2019) Dietary lectin exclusion: the next big food trend? World J Gastroenterol 25(24):2973–2976 82. Ooi SL, Correa D, Pak SC (2019) Probiotics, prebiotics, and low FODMAP diet for irritable bowel syndrome—what is the current evidence? Complement Ther Med 43:73–80 83. Mitchell H et al (2019) Review article: implementation of a diet low in FODMAPs for patients with irritable bowel syndrome-directions for future research. Aliment Pharmacol Ther 49(2):124–139 84. Konturek PC, Zopf Y (2017) Therapeutic modulation of intestinal microbiota in irritable bowel syndrome. From probiotics to fecal microbiota therapy. MMW Fortschr Med 159 (Suppl 7):1–5 85. Kay E et al (2019) Nonpharmacologic options for treating irritable bowel syndrome. JAAPA 32(3):38–42 86. Huaman JW et al (2018) Effects of prebiotics vs. a diet low in FODMAPs in patients with functional gut disorders. Gastroenterology 155(4):1004–1007 87. Johnsen PH et al (2018) Faecal microbiota transplantation versus placebo for moderate-tosevere irritable bowel syndrome: a double-blind, randomised, placebo-controlled, parallel-group, single-centre trial. Lancet Gastroenterol Hepatol 3(1):17–24 88. Halkjaer SI et al (2018) Faecal microbiota transplantation alters gut microbiota in patients with irritable bowel syndrome: results from a randomised, double-blind placebo-controlled study. Gut 67(12):2107–2115 89. De Palma G et al (2017) Transplantation of fecal microbiota from patients with irritable bowel syndrome alters gut function and behavior in recipient mice. Sci Transl Med 9(379)

22

Y. Chen et al.

90. Bray F et al (2018) Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin 68(6):394–424 91. Brennan CA, Garrett WS (2016) Gut microbiota, inflammation, and colorectal cancer. Annu Rev Microbiol 70:395–411 92. Tilg H et al (2018) The intestinal microbiota in colorectal cancer. Cancer Cell 33(6):954–964 93. Kwong TNY et al (2018) Association between bacteremia from specific microbes and subsequent diagnosis of colorectal cancer. Gastroenterology 155(2):383–390 e8 94. Yachida S et al (2019) Metagenomic and metabolomic analyses reveal distinct stage-specific phenotypes of the gut microbiota in colorectal cancer. Nat Med 25(6):968–976 95. Wirbel J et al (2019) Meta-analysis of fecal metagenomes reveals global microbial signatures that are specific for colorectal cancer. Nat Med 25(4):679–689 96. Thomas AM et al (2019) Metagenomic analysis of colorectal cancer datasets identifies cross-cohort microbial diagnostic signatures and a link with choline degradation. Nat Med 25(4):667–678 97. Nakatsu G et al (2015) Gut mucosal microbiome across stages of colorectal carcinogenesis. Nat Commun 6:8727 98. Zeller G et al (2014) Potential of fecal microbiota for early-stage detection of colorectal cancer. Mol Syst Biol 10:766 99. Cremonesi E et al (2018) Gut microbiota modulate T cell trafficking into human colorectal cancer. Gut 67(11):1984–1994 100. Malik A et al (2018) SYK-CARD9 signaling axis promotes gut fungi-mediated inflammasome activation to restrict colitis and colon cancer. Immunity 49(3):515–530 e5 101. Nakatsu G et al (2018) Alterations in enteric virome are associated with colorectal cancer and survival outcomes. Gastroenterology 155(2):529–541 e5 102. He Z, Gharaibeh RZ, Newsome RC, Pope JL, Dougherty MW, Tomkovich S, Pons B, Mirey G, Vignard J, Hendrixson DR, Jobin C (2019) Campylobacter jejuni promotes colorectal tumorigenesis through the action of cytolethal distending toxin. Gut 68(2): 289–300 103. Coleman OI et al (2018) Activated ATF6 induces intestinal dysbiosis and innate immune response to promote colorectal tumorigenesis. Gastroenterology 155(5):1539–1552 e12 104. Sanna S et al (2019) Causal relationships among the gut microbiome, short-chain fatty acids and metabolic diseases. Nat Genet 51(4):600–605 105. Sun BB et al (2018) Genomic atlas of the human plasma proteome. Nature 558(7708):73–79 106. Richardson LA (2017) Evolving as a holobiont. PLoS Biol 15(2):e2002168 107. Planes S et al (2019) The Tara Pacific expedition-A pan-ecosystemic approach of the “-omics” complexity of coral reef holobionts across the Pacific Ocean. PLoS Biol 17(9): e3000483

Chapter 3

Gut Microbiota and Liver Injury (I)—Acute Liver Injury Guangyan Wu, Sanda Win, Tin A. Than, Peng Chen, and Neil Kaplowitz

Abstract Over the last few decades, intestinal microbial communities have been considered to play a vital role in host liver health. Acute liver injury (ALI) is the manifestation of sudden hepatic injury and arises from a variety of causes. The studies of dysbiosis in gut microbiota provide new insight into the pathogenesis of ALI. However, the relationship of gut microbiota and ALI is not well understood, and the contribution of gut microbiota to ALI has not been well characterized. In this chapter, we integrate several major pathogenic factors in ALI with the role of gut microbiota to stress the significance of gut microbiota in prevention and treatment of ALI. Keywords Gut microbiota

3.1

 Acute liver injury  DILI

Introduction

Acute liver injury (ALI) can present as a life-threatening disorder and has emerged as a public health issue around the world. ALI can be caused by a series of risk factors, including drug-induced liver injury (DILI), acute viral hepatitis, acute alcohol-induced liver injury, and so on. A wide range of medications and herbal supplements can give rise to DILI, and DILI is one of the major causes of ALI. Acute liver failure (ALF) is a severe manifestation of ALI that seriously threatens human life and health. So far liver transplantation is the most effective therapy for ALF [1]. Nevertheless, many patients with ALF die due to the scarcity of liver G. Wu  P. Chen Department of Pathophysiology, Guangdong Provincial Key Laboratory of Proteomics, School of Basic Medical Sciences, Southern Medical University, N.No 1838 Guangzhou Ave., Guangzhou 510515, China e-mail: [email protected] S. Win  T. A. Than  N. Kaplowitz (&) USC Research Center for Liver Disease, Department of Medicine, Keck School of Medicine of USC, Los Angeles, CA 90089, USA e-mail: [email protected] © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_3

23

24

G. Wu et al.

sources and expensive healthcare costs. Therefore, the limited availability of liver transplantation justifies the need to seek other promising therapies in patients with ALF. For decades, accumulating evidence has demonstrated the close relationship between the gut microbiota and liver injury [2]. However, these researches had mostly focused on chronic liver disease. For example, it is widely discussed that dysbiosis of gut microbiota, including composition and function, is associated with nonalcoholic fatty liver disease (NAFLD) [3]. Recently, based on modulating the intestinal microbiota, an increasing number of preventive and adjuvant therapies have been proposed for ALI. In this section, we highlight the current knowledge of preclinical work to clinical studies about gut microbiota and its contribution in ALI.

3.2

The Gut Microbiota and Drug-Induced Liver Injury

It is well demonstrated that the majority of drugs are metabolized by liver, and some of them associated with hepatotoxicity causing DILI. DILI is a diagnosis of exclusion and is made only after elimination of common causes of liver disease, such as alcoholic liver disease, metabolic and genetic liver diseases, bile duct obstruction, hepatitis virus infection, hepatic ischemia, and autoimmune hepatitis. DILI is often detected during routine testing of serum chemistries [4]. Alanine aminotransferase (ALT) values greater than three times the upper limits of normal (ULN) have been suggested as a sensitive but not necessarily specific DILI signal. It is commonly recognized that DILI is the most predominant cause for acute liver failure in the United States [5, 6]. It is well known that liver transplantation is the most effective measures for the acute liver failure at the present time, but the limitation of liver source scarcity and high surgery expense is a serious threat against patients’ life. An immediate result of DILI is the withdrawal or restricted usage of otherwise efficacious drugs, leading to deficits in therapy [7, 8]. Therefore, DILI is not only a major financial burden for society but also a leading public health problem. DILI can be broadly classified into two types, intrinsic (predictable) or idiosyncratic DILI (IDILI) (unpredictable). The former is related to the drug dosage and all individuals are susceptible. A typical example is acetaminophen (APAP)induced liver injury. IDILI is defined as toxicity that is dose-independent, occurs in a minority of patients during clinical drug therapy, and is exemplified by isoniazid, troglitazone, various antibiotics, and tacrine. Troglitazone was the first peroxisomal proliferator-activated receptor (PPAR-c) agonist available to treat Type 2 diabetes. It was withdrawn from the market by the U.S. Food and Drug Administration (FDA) in 2000 [9]. Tacrine was the first centrally acting cholinesterase inhibitor approved for the treatment of Alzheimer’s disease was withdrawn from clinical use in 2013 [10]. In the past few decades, the gut microbiota is gradually being accepted as an environmental factor that affects host metabolism and contributes to associated host

3

Gut Microbiota and Liver Injury (I)—Acute Liver Injury

25

health and diseases. The gastrointestinal tract contains trillions of microorganisms that play an important role in host health. Currently, the connection between an altered gut microbiota and liver disorders such as alcoholic liver disease, nonalcoholic fatty liver disease, and primary sclerosing cholangitis [11] is well demonstrated. Although it is also extensively reported that some clinical drugs are metabolized by the liver to metabolites, and then causing DILI. However, the contribution of gut microbiota to interindividual variation in DILI is rarely reported. The process of drug metabolism is complicated and governed by a serious of drug transporters and metabolic enzymes, including cytochrome P450 enzymes (CYPs), N-acetyl transferases, sulfotransferases, and glucuronosyl transferases. However, only a small number of clinical drugs are demonstrated to be metabolically influenced by the gut microbiota. The role of gut microbiota in hepatotoxic responses induced by drugs is not well understood. Thus, it is noteworthy and imperative to investigate the functional link between gut microbiota and DILI. Understanding hepatotoxicity–microbiome interaction is critical for prevention and treatment of DILI. As of 2018, a total of 1356 agents were presented and discussed independently in Livertox Database. It is well known that the liver is a major metabolic organ involved in drug disposition and detoxification activities. It is now generally accepted that the gut microbiota exhibits a broad spectrum of enzymatic activities with profound impact on the metabolism of drugs. Therefore, to elucidate the roles of the gut microbiota on DILI, it has been suggested that the variation in the drug hepatotoxicity may be due to interindividual variability of gut microbiome. DILI is a major public health issue, and improving its prediction remains an enormous challenge. Currently, despite the continual advances of medical science and technology, the most commonly applied metric for potential hepatotoxicity is ALT levels higher than three times ULN. In addition, there are no other effective metrics and therapies available to prevent or treat drug-induced hepatotoxicity. Most of investigations have focused on the effects of drug metabolism or pharmacokinetics on the gut microbiota. In this section, which primarily focuses on the effect of gut microbiota on DILI, we examined the connection between gut microbiota and DILI, and provide several examples of drugs whose activities have been shown to be altered by gut microbiota, including drugs associated with intrinsic and idiosyncratic DILI.

3.2.1

APAP-Induced Liver Injury

Dose-dependent APAP-induced liver injury is referred to as intrinsic DILI, which is predictable and reproducible in preclinical models. In 1966, the first reports of APAP-induced hepatotoxicity were described [12]. From that time on, as a widely used analgesic and antipyretic drug, the overdose of APAP is the most striking example of direct hepatotoxicity and the cause of severe intrinsic DILI. Even at therapeutic dose, a study reported that healthy adults who used the maximum daily

26

G. Wu et al.

recommended dose of 4 g APAP for 5 days frequently cause elevations in serum aminotransferases [13]. However, another study demonstrated that it is safe to take APAP at half the maximum recommended daily dosage (2 g) in the long-term, and the elevation in aminotransferases are clinically insignificant [14]. Further research is warranted to determine the effects on liver function of maximum therapeutic doses of APAP. Case reports of acute liver injury due to APAP toxicity have appeared but the extent of the problem is not fully known. Overdose, either in suicidal or therapeutics settings, represents nearly half of all cases of ALF in the US and UK. Because of APAP dose dependence and similar metabolism in rodents, toxicity in mice is the reproducible experimental animal model and has helped to define the pathways leading to hepatotoxicity. Generally, the majority of a therapeutic dose of APAP is conjugated by O-sulfation or glucuronidation and then excreted into the bile or urine. More than 20 different CYPs exist in the CYP450 family in human liver. Especially, isozymes of families CYP1, CYP2, and CYP3 are commonly involved in most phase I biotransformation of drugs and other xenobiotics [15, 16]. A small part of APAP dose is transformed to toxic N-acetyl-p-benzoquinone imine (NAPQI) by the CYPs, (mainly CYP2E1 and CYP3A4) [17]. NAPQI can be detoxified by conjugation with glutathione (GSH) in the hepatocyte. Early mechanistic studies of APAP demonstrated that N-acetylcysteine (NAC) scavenges the reactive metabolite NAPQI of APAP by replenishment of GSH [18]. Subsequent studies suggested that mitochondrial damage and nuclear DNA fragmentation are likely to be predominant terminal events in APAP-induced cellular necrosis [19]. With the technologic development of metagenomics and pharmacometabonomics in recent decades, increasing evidence shows that gut microbiota and microbial metabolites contribute to the hepatotoxicity of APAP [20]. P-cresol is largely produced by the intestinal anaerobic flora [21] and recognized as a substrate for the human cytosolic sulfotransferases SULT1A1 related to the O-sulfation of APAP [22]. In addition, it has been reported that the conjugated APAP is deconjugated by b-glucuronidase and sulfatase of the liver and gut microbiota [23]. Clayton et al. reported that a person’s intestinal tract exhibits high p-cresol generation, leading to competitive inhibition of SULT1A1 by p-cresol which reduces the hepatic sulphation capacity of APAP [22], indicating the gut-derived microbial metabolite p-cresol may contribute to APAP-induced liver injury. Circadian clocks play a vital role in coordinating host daily physiology and metabolism. In the recent investigation, it has been shown that APAP-induced liver injury exhibits diurnal variation and chronotoxicity [24–26]. Chronology has largely been focused on the central clock and hepatocyte clock. The circadian feeding/ fasting-associated metabolism is controlled by the central clock and the transcriptional oscillations of APAP-metabolizing genes are driven by the hepatocyte clock, which may have mutual influence on APAP chronotoxicity. However, the detailed mechanisms responsible for the APAP chronotoxicity are not fully demonstrated. Interestingly, it is established that gut microbiota exhibits diurnal oscillations at the compositional and functional level and affects metabolic homeostasis in both mice and humans [27]. Thaiss et al. also showed that the diurnal oscillations in intestinal

3

Gut Microbiota and Liver Injury (I)—Acute Liver Injury

27

microbiota regulate the circadian liver transcriptome and detoxification pattern [28]. A novel area of recent work relates to the analysis of the interaction of the APAP-induced diurnal variation of liver injury with the gut microbiota. APAP-induced liver injury exhibits diurnal variation, it shows more severe liver injury when APAP was given at night (ZT12 = when the light is off—start of active period) compared with in the morning (ZT0 = when the light is on—start of resting period). Mice received fecal microbiota transplantation (FMT) separately from ZT12 and ZT0 and ZT12 microbiota enhanced morning toxicity. Specifically, as a type of gut microbial metabolite, the level of 1-phenyl-1,2-propanedione (PPD) was significantly enhanced at ZT12 compared with ZT0 based on metabonomics. Further study found that the PPD is involved in the rhythmic liver injury induced by APAP, by depleting hepatic GSH levels. What’s more, oral administration of Saccharomyces cerevisiae could reduce the PPD level and then ameliorate APAP-induced liver injury at ZT12 [29]. It indicated that a novel mechanism is providing the hepatotoxic rhythmicity of APAP mediated by gut microbial functional oscillations (Fig. 3.1). It is widely reported that dietary polyphenols are natural compounds existing in plants and contribute to health through gut microbiota in humans [30] and animals [31]. It has been estimated that majority of polyphenols (more than 90% of total polyphenol intake) are subjected to the enzymatic activities of the gut microbiota in the colon [32, 33]. The flavonol quercetin is one of the most abundant polyphenols

ZT 0

ZT 12

APAP

APAP

GSH

PPD

GSH

PPD

Saccharomyces cerevisiae

Fig. 3.1 Gut microbiota and the diurnal variation of APAP hepatotoxicity

28

G. Wu et al.

which exist in vegetables, fruits, and officinal plants. It has been reported that 3,4-dihydroxyphenylacetic acid (DOPAC), a microbiota-derived metabolite of quercetin, protected against APAP-induced liver injury via Nrf-2 activation [34]. Previous studies showed that 4-hydroxyphenylacetic acid (4-HPA), the major microbial metabolite of polyphenols, is associated with the antianxiety, antiplatelet, and antioxidative activities. Recently, Zhao et al. [35] reported that 4-HPA has a beneficial impact on the detoxification of APAP hepatotoxicity. It is related to the inhibition of CYP2E1 and the enhancement of antioxidant enzymes and phase II enzymes via Nrf2 activation. What’s more, molecular docking study by Zhao and coworkers pointed out that 4-HPA has a higher docking score for binding sites on CYP2E1 compared with the antidote for APAP poisoning, NAC [35]. Taken together, this finding indicated 4-HPA, the microbial metabolite of polyphenol, could be an effective preventive intervention for APAP-induced liver injury, although such a hypothesis warrants future investigation. These studies support the notion that gut microbiota and microbial metabolites play a key role in APAP-induced liver injury. Gut microbiota produce metabolites from dietary sources, which can promote or inhibit APAP toxicity. The gut microbiota exhibited diurnal differences which may influence exposure to the gut-derived metabolites. Furthermore, gut permeability promotes inflammatory responses in the liver which may influence APAP toxicity. Continued research within this field is warranted in order to predict the progression of APAP chronotoxicity and help in better understanding of the role and the mechanisms of the gut microbiota in APAP metabolism.

3.2.2

Tacrine-Induced Liver Injury

Alzheimer’s disease (AD) is a complex age-related neurodegenerative disease and one of the major factors that cause dementia, mainly occurring in the elderly [36]. With the aggravating trend of an aging population, AD prevention and therapy studies require significant attention. As we have known, tacrine, a reversible acetylcholinesterase inhibitor, was the first drug approved for the treatment of dementia of AD. However, due to its hepatotoxicity characterized by elevation in aminotransferase [37], tacrine was discontinued by FDA. It is important to explore the mechanism of tacrine-induced liver injury, so that the researchers can better develop new derivative drugs for the prevention and cure of AD in the future. A recent study conducted by Yip et al. revealed the role of gut microbiota in tacrine-induced liver injury [10]. According to pharmacokinetic studies, they showed enhanced enterohepatic recycling of deglucuronidated tacrine correlated with tacrine hepatotoxicity in rats. Metabonomic studies suggested that gut microbial activities play a pivotal role in tacrine-induced transaminitis. Metagenomics studies demonstrated that an enrichment of b-glucuronidase-producing bacteriae (Bacteroides and Enterobacteriaceae), an abatement of b-glucuronidase-inhibiting bacteriae (Lactobacillus) and an increase in b-glucuronidase genes was exhibited in

3

Gut Microbiota and Liver Injury (I)—Acute Liver Injury

29

strong responders of tacrine. Ultimately, in the validation study, pretreatment with b-glucuronidase increased the tacrine-induced hepatotoxicity in rats, and pretreatment of antibiotics (vancomycin and imipenem) decreased the susceptibility to the hepatotoxicity of tacrine. Together with the new scientific techniques, Yip et al. supported the hypothesis that the gut microbiota, especially the intestinal b-glucuronidase activities, have an impact on tacrine-induced liver injury through enhancing enterohepatic exposure of the liver to the toxic parent drug. b-glucuronidase, as a toxicologically relevant intestinal bacterial enzyme, which has been demonstrated as newly developed selective bacterial b-glucuronidase inhibitors became available [38]. The specific inhibition of bacterial b-glucuronidase was shown to ameliorate tacrine-induced hepatotoxicity. The relevance of this research to other therapeutic agents with hepatotoxic potential and large interindividual variation is intriguing. Therefore the contribution of the gut microbiota in personalized medicine research is strongly suggested and needs further investigation.

3.2.3

Diclofenac-Induced Liver Injury

Diclofenac (DCLF) is a widely used nonsteroidal anti-inflammatory drug (NSAIDs) that commonly causes intestinal damage as a side effect. The main metabolic pathway of DCLF is through 4′-hydroxylation by CYP2C9 [39], and the formation of a toxic metabolite leads to acute lethal cell injury [40]. DCLF is partly metabolized to DCLF-acyl glucuronides which then are cleaved by bacterial b-glucuronidase, producing the potentially harmful aglycone. Besides, in vitro, it was confirmed that DCLF-acyl glucuronides were wildly catalyzed to DCLF by purified Escherichia coli b-glucuronidase, and bacterial-specific b-glucuronidase inhibitor restrained this reaction [41]. In addition, another major adverse reaction of DCLF is implicated in hepatotoxicity that is associated with IDILI [42]. Owing to the dose-independence, the variable temporal relationship with drug exposure, and the lack of experimental animal models, the mechanisms of DCLF-induced liver injury remain mostly unclear. However, a few years ago, Yano et al. developed a DCLF-induced ALF model in normal mice and demonstrated the Th17 and IL-1b-related immunological factors are involved in the pathogenesis of DCLF-induced liver injury in mice [43]. Therefore, it indicated that gut microbiota and its enzymes, such as bacterial b-glucuronidase, may be involved in DCLF-induced liver injury. A potential model of DCLF-induced IDILI was established by Deng et al. in 2006. Bacteria/lipopolysaccharide (LPS) contributed to a nontoxic dose of DCLF to induce liver injury. Gut sterilization attenuated DCLF-induced liver injury, indicating that hepatotoxicity induced by high concentrations of DCLF is caused in part by its ability to enhance intestinal permeability to LPS or other bacterial products [44]. Further study found that intestinal bacteria/LPS are involved in DCLF-induced liver injury in

30

G. Wu et al.

a neutrophil-independent manner and indicate that hypoxia plays a crucial role in the pathogenesis [45]. In this section, gut microbiota involving DILI were categorized based on published documented cases of hepatotoxicity in recent years. We mainly integrate several typical drugs with the role of gut microbiota in DILI to stress the clinical significance of gut microbiota for safe and effective use of drugs. In addition to the abovementioned drugs whose toxicity may be mediated by metabolism by gut microbiota, there are many studies of drug–drug interactions influenced by the gut microbiota. The ability to elucidate the mechanisms underlying the gut microbiota responses in DILI is paramount in developing personalized medicine research. Much progress has been recently made in the field of the gut microbiota in DILI. This section provided several cases of the interindividual responses to DILI which contributed to both the influence of the host and the gut microbiota. Accumulating evidence indicates that the gut microbiota may be a target for monitoring susceptibility to DILI. Understanding the relationship of gut microbiota and DILI may enhance our ability to identify and predict susceptible interindividual variability, and then prevent serious adverse reactions. These studies shed new light on the interaction of gut microbiota in DILI. This is an emerging field with enormous potential.

3.3

The Gut Microbiota and D-Galactosamine-Induced Acute Liver Injury

D-galactosamine (GalN) specifically inhibits RNA and protein synthesis in the hepatocytes, while the hypomethylation of ribosomal RNA is involved in protein synthesis defect [46]. GalN is a well-established ALI animal model induced by intraperitoneal injection of GalN that is similar to human viral infections (hepatitis A, B, and E). More generally, the combined injection of a subtoxic dose of GalN and LPS or TNF has been extensively used in the highly reproducible experimental animal model, which more closely resembles fulminant hepatic failure in the clinic. Previous study has shown that bacterial translocation from gut both to the blood and intraabdominal organs occurred in ALI induced by GalN and then decreased by 70% liver resection with improvement of liver function in rats [47]. However, liver resection in patients can be performed only under strict indications; it is important to explore the relationship of gut microbiota and GalN-induced liver injury. It is reported that the gut microbiota was altered in rats with GalN-induced ALF, which is associated with the translocation of endotoxin from lumen [48]. Studies lead to a hypothesis that administration of probiotic may decrease the bacterial translocation and reduce the bacterial endotoxin exposure. Administration of probiotic Lactobacillus, Bifidobacterium with or without blueberry exhibited therapeutic effects on LPS/GalN-induced liver injury in various degree [49]. Further, when rats were pretreated by gavage with probiotic Bifidobacterium adolescentis CGMCC

3

Gut Microbiota and Liver Injury (I)—Acute Liver Injury

31

15058, GalN-induced liver injury were alleviated with anti-inflammatory effects. This indicated that these protective effects might be correlated with enhanced intestinal barrier and modified composition of the gut microbiota [50]. It is generally known that the mitogen-activated protein kinase (MAPK) signaling pathway is involved in the mediation and regulation of a variety of physiological processes, such as proliferation, differentiation, stress response, and apoptosis. However, little research had been published to discuss the effect of MAPK signaling pathway on LPS/GalN-induced ALF. A report by Wang et al. showed that when rats given LPS/ GalN were orally administrated pretreatment with probiotic Lactobacillus casei Zhang, pro-inflammatory cytokine production were reduced and liver injury was attenuated by modulating the TLR-MAPK-PPAR-c signaling pathways and gut microbiota [51]. Taken together, these findings suggest that the administration of probiotic might provide a new and effective way for the prevention and treatment of ALI.

3.4

The Gut Microbiota and Acute Alcohol-Induced Liver Injury

Alcohol abuse has been recognized as a major cause of liver injury, including acute and chronic liver injury. The metabolic mechanism of alcohol is complicated, which has been described elsewhere [52]. CYP2E1 in the endoplasmic reticulum is also involved in process of oxidative metabolism of ethanol to acetaldehyde. It has been well demonstrated that acetaldehyde, as the main toxic metabolite of ethanol, can cause direct damage to mitochondria [53]. In addition to alcohol consumption, endogenous ethanol production can be fermented by the commensal microbiota, then being metabolized via the portal vein in the liver, but also further metabolized by the gut microbiota [54]. Many previous studies have addressed the effects of gut microbiota on long-term ethanol consumption, focusing on alcoholic liver disease both in animals and humans. It is widely documented that long-term alcohol consumption disrupts the intestinal epithelial barrier and tight junctions, resulting intestinal hyperpermeability [55]. Alcoholic liver disease is the most common chronic liver injury, and numerous reviews have described the bidirectional communication between gut and liver [56– 58]. It indicates that gut microbiota is inextricably linked to alcohol-induced liver injury. However, published studies on the effects of gut microbiota on acute alcohol-induced liver injury are encouraging but are limited. Studies have shown that acute exposure to ethanol at high concentration can cause the dysbiosis of gut microbiota and mucosal damage, and then increase intestinal permeability, leading to the translocation of enteral bacteria and its endotoxin from the intestine into the liver [59, 60]. It was reported that probiotics Lactobacillus rhamnosus GG culture supernatant pretreatment for 5 days improved intestinal integrity and liver injury in a “binge-drinking” mouse model [60]. Pretreatment with rhubarb extract for 17 days

32

G. Wu et al.

ameliorated alcohol-induced liver damage by improving intestinal homeostasis and restoring gut barrier function in a mouse model of binge drinking, especially the relative abundance of Akkermansia muciniphila was increased in the cecal content [61]. Akkermansia muciniphila, a Gram-negative, anaerobic bacterium, is recognized as potential probiotic that strengthens the integrity of intestinal barrier [62], as well as reverses high-fat diet-induced abnormal intestinal permeability [63]. Whether the protective effects observed here extend to the progress of ethanol-induced mucosal damage is less clear. A recent study by Grander et al. revealed that patients with alcoholic steatohepatitis showed a decreased abundance of faecal Akkermansia muciniphila. In experimental acute and chronic alcoholic liver disease, oral supplementation with Akkermansia muciniphila protected against ethanol-induced gut leakiness and ameliorated alcohol-induced liver injury [64]. Consequently, the administration of probiotic may have a beneficial effect in patients with acute ethanol-induced liver injury, although such a hypothesis warrants future investigation in clinical studies. Canesso and colleagues reported that germ-free mice showed no liver injury and less inflammation compared with conventional mice after acute alcohol intake, which indicated gut microbiota contribute either directly or indirectly to liver injury due to acute alcohol consumption [65]. Interestingly, another study reached completely different conclusions. In a binge drinking-induced liver injury model, germ-free mice exhibited more liver damage and inflammation compared with the conventional mice, which suggested the commensal microbiota have a certain protective effect on acute alcohol-induced liver injury [66]. A possible explanation for this difference may be due to the different doses of ethanol and the duration of exposure in mouse models. Furthermore, this could represent an example of hormesis, e.g., microbiota exposure may sustain expression of defensive factors whereas complete absence of such phenomena may weaken defense. In any case, these studies shed new light on the gut microbiota in acute alcohol-induced liver injury. In the future, more research is needed to confirm these preliminary results and to determine whether the gut microbiota is a contributor or a therapeutic target in this area.

3.5

The Gut Microbiota and Viral Hepatitis-Related ALI

Viral hepatitis can cause acute or chronic liver injury. Hepatitis A, B, C, D, and E viruses infection can lead to ALI and even ALF [67]. In recent decades, more and more studies on gut microbiota have provided novel insights for the prevention and treatment of hepatitis viruses infection, mostly about the chronicity of hepatitis viruses infection [68, 69]. So far, however, there is no evidence for a direct relation between gut microbiota and viral hepatitis-related ALI. A study by Zhang et al. showed that patients with hepatitis B-related acute-on-chronic liver failure have increased circulating bacterial DNA and decreased microbial diversity. Also, gut-derived bacterial translocation may be one of the causes of the increased

3

Gut Microbiota and Liver Injury (I)—Acute Liver Injury

33

peripheral bacterial load [70]. Although less well described, it suggests that gut microbiota may be associated with the development of acute viral hepatitis.

3.6

The Gut Microbiota and Transplantation Linked Liver Injury

Liver transplantation is the most effective approaches to cure end-stage liver diseases. However, due to ischemia and reperfusion (ischemia/reperfusion injury, IRI) during the surgery, recipient may develop hepatic injury, which is the main problem for the allograft dysfunction. The mechanism of IRI is complex, involving redox imbalance and inflammation responses. Recently, Nakamura et al. reported that amoxicillin pretreatment could alleviate liver transplantation linked hepatic damage, which may be through decreasing CCAAT/enhancer-binding protein homologous protein (CHOP) expression, enhancing autophagy, and inhibiting inflammation [71]. They further validate the finding in humans. This interesting study directly linked gut microbiota and IRI. However, the authors used amoxicillin alone which could only deplete part of the gut microbiota. Whether depletion of all of gut microbiota would differently influence IRI still requires further investigation.

3.7

Conclusion

It is recognized that the gut microbiota plays a vital role in body health maintenance, and a well-balanced gut microbiota is critical in the regulation of host homeostasis. Actually, the gut microbiota is a complex community, including bacteria and the nonbacterial microbiome such as fungi, archaea, and viruses. Most studies focus on bacteria in liver disease. So far, little progress has been made toward understanding the contribution of nonbacterial microbiome in this field. This chapter mainly summarizes the current state of knowledge on the role of the gut microbiota in ALI, especially bacteria. Future studies need to pay more attention to investigation of both the bacteria and the nonbacterial components of microbiota in the development and progression of liver disease, especially in ALI.

References 1. Bernal WAG, Dhawan A, Wendon J (2010) Acute liver failure. Lancet 376(9736):190–201 2. Chassaing B, Etienne-Mesmin L, Gewirtz AT (2014) Microbiota-liver axis in hepatic disease. Hepatology 59(1):328–339 3. Boursier J, Mueller O, Barret M (2016) The severity of nonalcoholic fatty liver disease is associated with gut dysbiosis and shift in the metabolic function of the gut microbiota. Hepatology 63(3):764–775

34

G. Wu et al.

4. Davern TJ, Chalasani N, Fontana RJ et al (2011) Acute hepatitis E infection accounts for some cases of suspected drug-induced liver injury. Gastroenterology 141(5):1665–1672 5. Lee WM (2013) Drug-induced acute liver failure. Clin Liver Dis 17(4):575–586 6. Goldberg DS, Forde KA, Carbonari DM et al (2015) Population-representative incidence of drug-induced acute liver failure based on an analysis of an integrated health care system. Gastroenterology 148(7):1353–1361 7. Kullak-Ublick GA, Andrade RJ, Merz M et al (2017) Drug-induced liver injury: recent advances in diagnosis and risk assessment. Gut 66(6):1154–1164 8. Medina-Caliz I, Robles-Diaz M, Garcia-Munoz B et al (2016) Definition and risk factors for chronicity following acute idiosyncratic drug-induced liver injury. J Hepatol 65(3):532–542 9. Graham DJ, Green L, Senior JR et al (2003) Troglitazone-induced liver failure: a case study. Am J Med 114(4):299–306 10. Yip LY, Aw CC, Lee SH (2018) The liver-gut microbiota axis modulates hepatotoxicity of Tacrine in the rat. Hepatology 67:282–295 11. Kummen M, Holm K, Anmarkrud JA et al (2017) The gut microbial profile in patients with primary sclerosing cholangitis is distinct from patients with ulcerative colitis without biliary disease and healthy controls. Gut 66(4):611–619 12. Davidson DG, Eastham WN (1966) Acute liver necrosis following overdose of paracetamol. Br Med J 2(5512):497–499 13. Watkins PB, Kaplowitz N, Slattery JT, Colonese CR, Colucci SV, Stewart PW, Harris SC (2006) Aminotransferase elevations in healthy adults receiving 4 grams of acetaminophen daily: a randomized controlled trial. JAMA 5;296(1):87–93 14. Ioannides SJ, Siebers R, Perrin K et al (2015) The effect of 1 g of acetaminophen twice daily for 12 weeks on alanine transaminase levels–a randomized placebo-controlled trial. Clin Biochem 48(10–11):713–715 15. Lewis DF (2004) 57 varieties: the human cytochromes P450. Pharmacogenomics 5(3):305– 318 16. Zanger UM, Turpeinen M, Klein K et al (2008) Functional pharmacogenetics/genomics of human cytochromes P450 involved in drug biotransformation. Anal Bioanal Chem 392 (6):1093–1108 17. Laine JE, Auriola S, Pasanen M et al (2009) Acetaminophen bioactivation by human cytochrome P450 enzymes and animal microsomes. Xenobiotica 39(1):11–21 18. McGill MR, Jaeschke H (2013) Metabolism and disposition of acetaminophen: recent advances in relation to hepatotoxicity and diagnosis. Pharm Res 30(9):2174–2187 19. McGill MR, Sharpe MR, Williams CD et al (2012) The mechanism underlying acetaminophen-induced hepatotoxicity in humans and mice involves mitochondrial damage and nuclear DNA fragmentation. J Clin Invest 122(4):1574–1583 20. Lee SH, An JH, Lee HJ et al (2012) Evaluation of pharmacokinetic differences of acetaminophen in pseudo germ-free rats. Biopharm Drug Dispos 33(6):292–303 21. Bone E, Tamm A, Hill M (1976) The production of urinary phenols by gut bacteria and their possible role in the causation of large bowel cancer. Am J Clin Nutr 29(12):1448–1454 22. Clayton TA, Baker D, Lindon JC et al (2009) Pharmacometabonomic identification of a significant host-microbiome metabolic interaction affecting human drug metabolism. Proc Natl Acad Sci U S A 106(34):14728–14733 23. Lee HWCM, Bae EA, Kim DH (2003) b-Glucuronidase inhibitor tectorigenin isolated from the flower of Pueraria thunbergiana protects carbon tetrachloride-induced liver injury. Liver Int 23:221–226 24. Kakan X, Chen P, Zhang J (2011) Clock gene mPer2 functions in diurnal variation of acetaminophen induced hepatotoxicity in mice. Exp Toxicol Pathol 63(6):581–585 25. Kim YC, Lee SJ (1998) Temporal variation in hepatotoxicity and metabolism of acetaminophen in mice. Toxicology 128(1):53–61 26. Johnson BP, Walisser JA, Liu Y et al (2014) Hepatocyte circadian clock controls acetaminophen bioactivation through NADPH-cytochrome P450 oxidoreductase. Proc Natl Acad Sci U S A 111(52):18757–18762

3

Gut Microbiota and Liver Injury (I)—Acute Liver Injury

35

27. Thaiss CA, Zeevi D, Levy M et al (2014) Transkingdom control of microbiota diurnal oscillations promotes metabolic homeostasis. Cell 159(3):514–529 28. Thaiss CA, Levy M, Korem T et al (2016) Microbiota diurnal rhythmicity programs host transcriptome oscillations. Cell 167(6):1495–1510 29. Gong S, Lan T, Zeng L et al (2018) Gut microbiota mediates diurnal variation of acetaminophen induced acute liver injury in mice. J Hepatol 69(1):51–59 30. Hervert-Hernández D, Goñi I (2011) dietary polyphenols and human gut microbiota: a review. Food Rev Int 27(2):154–169 31. Anhê FF, Roy D, Pilon G, Dudonné S, Matamoros S, Varin TV (2015) A polyphenol-rich cranberry extract protects from diet-induced obesity, insulin resistance and intestinal inflammation in association with increased Akkermansia spp. population in the gut microbiota of mice. Gut 64(6):872–883 32. Bowey E, Adlercreutz H, Rowland I (2003) Metabolism of isoflavones and lignans by the gut microflora: a study in germ-free and human flora associated rats. Food Chem Toxicol 41 (5):631–636 33. Cardona F, Andres-Lacueva C, Tulipani S et al (2013) Benefits of polyphenols on gut microbiota and implications in human health. J Nutr Biochem 24(8):1415–1422 34. Xue H, Xie W, Jiang Z et al (2016) 3,4-Dihydroxyphenylacetic acid, a microbiota-derived metabolite of quercetin, attenuates acetaminophen (APAP)-induced liver injury through activation of Nrf-2. Xenobiotica 46(10):931–939 35. Zhao H, Jiang Z, Chang X et al (2018) 4-Hydroxyphenylacetic acid prevents acute APAP-induced liver injury by increasing phase II and antioxidant enzymes in mice. Front Pharmacol 9:653 36. Goedert M, Spillantini MG (2006) A century of Alzheimer’s disease. Science 314(5800):777– 781 37. Watkins PB, Zimmerman HJ, Knapp MJ et al (1994) Hepatotoxic effects of tacrine administration in patients with Alzheimer’s disease. JAMA 271(13):992–998 38. Wallace BD, Wang H, Lane KT et al (2010) Alleviating cancer drug toxicity by inhibiting a bacterial enzyme. Science 330(6005):831–835 39. Leemann T, Transon C, Dayer P (1993) Cytochrome P450TB (CYP2C): a major monooxygenase catalyzing diclofenac 4’-hydroxylation in human liver. Life Sci 52(1):29–34 40. Kretz-Rommel A, Boelsterli UA (1993) Diclofenac covalent protein binding is dependent on acyl glucuronide formation and is inversely related to P450-mediated acute cell injury in cultured rat hepatocytes. Toxicol Appl Pharmacol 120(1):155–161 41. LoGuidice A, Wallace BD, Bendel L et al (2012) Pharmacologic targeting of bacterial beta-glucuronidase alleviates nonsteroidal anti-inflammatory drug-induced enteropathy in mice. J Pharmacol Exp Ther 341(2):447–454 42. Aithal GP (2005) Diclofenac-induced liver injury: a paradigm of idiosyncratic drug toxicity. Expert Opin Drug Saf 3(6):519–523 43. Yano A, Higuchi S, Tsuneyama K et al (2012) Involvement of immune-related factors in diclofenac-induced acute liver injury in mice. Toxicology 293(1–3):107–114 44. Deng X, Stachlewitz RF, Liguori MJ et al (2006) Modest inflammation enhances diclofenac hepatotoxicity in rats: role of neutrophils and bacterial translocation. J Pharmacol Exp Ther 319(3):1191–1199 45. Deng X, Liguori MJ, Sparkenbaugh EM et al (2008) Gene expression profiles in livers from diclofenac-treated rats reveal intestinal bacteria-dependent and -independent pathways associated with liver injury. J Pharmacol Exp Ther 327(3):634–644 46. Clawson GASJ, Milam K, Wang YF, Gabriel C (1990) The hepatocyte protein synthesis defect induced by galactosamine involves hypomethylation of ribosomal RNA. Hepatology 11(3):428–434 47. Kasravi FBWL, Wang XD, Molin G, Bengmark S, Jeppsson B (1996) Bacterial translocation in acute liver injury induced by D-Galactosamine. Hepatology 23(1):97–103 48. Li LJ, Wu ZW, Xiao DS, Sheng JF (2004) Changes of gut flora and endotoxin in rats with D-galactosamine-induced acute liver failure. World J Gastroenterol 10(14):2087–2090

36

G. Wu et al.

49. Osman N, Adawi D, Ahrne S et al (2007) Endotoxin- and D-galactosamine-induced liver injury improved by the administration of Lactobacillus, Bifidobacterium and blueberry. Dig Liver Dis 39(9):849–856 50. Li Y, Lv L, Ye J et al (2019) Bifidobacterium adolescentis CGMCC 15058 alleviates liver injury, enhances the intestinal barrier and modifies the gut microbiota in D-galactosamine-treated rats. Appl Microbiol Biotechnol 103(1):375–393 51. Wang Y, Xie J, Li Y et al (2016) Probiotic Lactobacillus casei Zhang reduces pro-inflammatory cytokine production and hepatic inflammation in a rat model of acute liver failure. Eur J Nutr 55(2):821–831 52. Zakhari S (2006) Overview how is alcohol metabolized by the body? Alcohol Res Health 29 (4):245–254 53. Arteel GE (2003) Oxidants and antioxidants in alcohol-induced liver disease [J]. Gastroenterology 124(3):778–790 54. Salaspuro V, Nyfors S, Heine R et al (1999) Ethanol oxidation and acetaldehyde production in vitro by human intestinal strains of Escherichia coli under aerobic, microaerobic, and anaerobic conditions. Scand J Gastroenterol 34(10):967–973 55. Elamin EE, Masclee AA, Dekker J et al (2013) Ethanol metabolism and its effects on the intestinal epithelial barrier. Nutr Rev 71(7):483–499 56. Tripathi A, Debelius J, Brenner DA et al (2018) The gut-liver axis and the intersection with the microbiome. Nat Rev Gastroenterol Hepatol 15(7):397–411 57. Schnabl B, Brenner DA (2014) Interactions between the intestinal microbiome and liver diseases [J]. Gastroenterology 146(6):1513–1524 58. Bajaj JS (2019) Alcohol, liver disease and the gut microbiota. Nat Rev Gastroenterol Hepatol 59. Lippai D, Bala S, Catalano D et al (2014) Micro-RNA-155 deficiency prevents alcohol-induced serum endotoxin increase and small bowel inflammation in mice [J]. Alcohol Clin Exp Res 38(8):2217–2224 60. Wang YLY, Sidhu A, Ma Z, McClain C, Feng W (2012) Lactobacillus rhamnosus GG culture supernatant ameliorates acute alcohol-induced intestinal permeability and liver injury. Am J Physiol Gastrointest Liver Physiol 303(1):G32-41 61. Neyrinck AM, Etxeberria U, Taminiau B et al (2017) Rhubarb extract prevents hepatic inflammation induced by acute alcohol intake, an effect related to the modulation of the gut microbiota Mol Nutr Food Res 61(1) 62. Reunanen J, Kainulainen V, Huuskonen L et al (2015) Akkermansia muciniphila adheres to enterocytes and strengthens the integrity of the epithelial cell layer. Appl Environ Microbiol 81(11):3655–3662 63. Everarda A, Belzer C, Geurtsa L (2013) Cross-talk between Akkermansia muciniphila and intestinal epithelium controls diet-induced obesity. Proc Natl Acad Sci U S A, 110(22):9066– 71 64. Grander CAT, Wieser V, Lowe P (2018) Recovery of ethanol-induced Akkermansia muciniphila depletion ameliorates alcoholic liver disease. Gut 67(5):891–901 65. Canesso MCC, Lacerda NL, Ferreira CM, Gonçalves JL (2014) Comparing the effects of acute alcohol consumption in germ-free and conventional mice: the role of the gut microbiota. BMC Microbiol 14:240 66. Chen P, Miyamoto Y, Mazagova M et al (2015) Microbiota protects mice against acute alcohol-induced liver injury. Alcohol Clin Exp Res 39(12):2313–2323 67. Vinholt Schiødt F (2003) Viral hepatitis-related acute liver failure. Am J Gastroenterol 98 (2):448–453 68. Chou HH, Chien WH, Wu LL et al (2015) Age-related immune clearance of hepatitis B virus infection requires the establishment of gut microbiota. Proc Natl Acad Sci U S A 112 (7):2175–2180 69. Ren YDYZ, Yang LZ, Jin LX, Wei WJ (2017) Fecal microbiota transplantation induces hepatitis B virus e-antigen (HBeAg) clearance in patients with positive HBeAg after long-term antiviral therapy. Hepatology 65(5):1765–1768

3

Gut Microbiota and Liver Injury (I)—Acute Liver Injury

37

70. Zhang Y, Zhao R, Shi D et al (2019) Characterization of the circulating microbiome in acute-on-chronic liver failure associated with hepatitis B. Liver Int 39(7):1207–1216 71. Nakamura K, Kageyama S, Ito T, Hirao H, Kadono K, Aziz A, Dery KJ, Everly MJ, Taura K, Uemoto S, Farmer DG, Kaldas FM, Busuttil RW, Kupiec-Weglinski JW (2019) Antibiotic pretreatment alleviates liver transplant damage in mice and humans. J Clin Invest 129 (8):3420–3434

Chapter 4

Gut Microbiota and Liver Injury (II): Chronic Liver Injury Susan S. Baker and Robert D. Baker

Abstract Chronic liver injury mainly comprises viral hepatitis, fatty liver disease, autoimmune hepatitis, cirrhosis and liver cancer. It is well established that gut microbiota serves as the key upstream modulator for chronic liver injury progression. Indeed, the term “gut–liver axis” was mostly applied for chronic liver injury. In the current chapter, we will summarize the relationship between gut microbiota and chronic liver injury, including the interaction between them based on latest clinic and basic research.





Keywords Gut microbiota Viral hepatitis Nonalcoholic fatty liver disease Alcoholic liver disease Autoimmune hepatitis Cirrhosis Hepatocellular carcinoma



4.1







Overview

The term “chronic liver disease” includes any condition that affects the liver and is present for a prolonged period of time. The etiology is variable and encompasses infection, metabolic disease, inflammatory, autoimmune, toxin and, at times, the cause cannot be identified. Chronic liver disease progresses to cirrhosis and liver failure unless treatment is available and implemented. It is important to note that the initiation and development of chronic liver disease is complex and likely involves the juxtaposition and interaction of several factors, including genetics, environmental factors, toxins, infections, behaviors and the gut microbiome. Chronic liver disease is the 15th leading cause of mortality in the United States and is responsible for the deaths of more than 25,000 Americans each year [1]. Globally, cirrhosis was S. S. Baker (&)  R. D. Baker Department of Pediatrics, Jacobs School of Medicine and Biomedical Sciences, State University of New York at Buffalo, Buffalo, USA e-mail: [email protected] S. S. Baker 39 Irving Place, Buffalo, NY 14201, USA © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_4

39

40

S. S. Baker and R. D. Baker

the 11th leading cause of death and accounted for more than 1.2 million deaths in 2016 [2]. The gut microbiome is considered to be a complex ecosystem that involves communication among the organisms, bacteria, fungi, phages, that reside in the ecosystem as well as within the gastrointestinal tract itself and likely other organs, notably the liver, that have access to this ecosystem through the portal circulation. The organisms are estimated to number about 100 trillion and vary with the site along the approximately 9 m adult gastrointestinal tube and location within any given site (luminal, mucosal, submucosal [3]). There is evidence that diet has an important effect on the gut microbiome [4–6]. The ecosystem is generally considered beneficial as its members serve to strengthen gut integrity, harvest energy, contribute to the regulation of host immunity and make vitamin B12, among other functions. Disruption of this ecosystem, often times labeled as dysbiosis, has been associated with many chronic diseases. The liver is particularly vulnerable to the influence of the gut microbiome, its metabolites and inflammatory factors because it has direct access to these through the portal vein. The relationship between the liver and the gut microbiome as a cause or contributor to chronic disease is intriguing because manipulation of the microbiome could lead to improvement of the disease. However, to date the relationship between liver disease and the gut microbiome is descriptive as a causal relationship has not been identified in humans.

4.2

Hepatitis B Virus (A DNA Virus)

Hepatitis B continues to be a global health problem. It is estimated that 270 million people worldwide are infected with the hepatitis B virus. However, among this large number of individuals, there are illuminating details. In endemic areas, the majority of the infected individuals have vertically acquired chronic hepatitis B which they contracted at the time of birth from their chronically infected mothers [7]. This is due to a difference in the way neonates and adults deal with the infection. Hepatitis B infection in neonates becomes chronic 90% of the time, while almost all who become infected as adults eventually clear the virus [8]. The difference in viral clearance has led to a close investigation of the way neonates and adults respond to the infection. There are several obvious differences between neonates and adults. Importantly for the present discussion is the fact that the neonate’s gastrointestinal tract is initially characterized by a paucity of microbes, but rapidly gains bacteria. The early gut microbiome varies considerably but at about a year of age switches to a more adult-like distribution and subsequently remains relatively stable [9]. In a mouse study of age dependency and chronicity of hepatitis B virus infection, Chou et al. report that in C3H/HeN mice the gut microbiome stabilized at about 8 weeks. Mice injected with HBV at 6 weeks (prior to stabilization) became persistently infected while mice injected at 12 weeks (post stabilization) cleared the

4 Gut Microbiota and Liver Injury (II): Chronic Liver Injury

41

virus. If the gut microbiome was obliterated using antibiotics, persistent infection occurred in the 12-week-old injected mice. Chou et al. hypothesized that their findings could be due to TLR4 and its ligand, LPS. When young mice with either a mutation in TLR4 or Kupffer cell depletion were injected, they cleared the virus rapidly [10]. Another obvious difference is that the neonatal gut has the function of defining what is recognized as a foreign protein. To accomplish this function the neonatal gut is necessarily more permeable than the adult gut. Lipopolysaccharide (LPS), derived from gram negative bacteria, enters the portal system and reaches the liver, where with its ligand TLR4 participates in a number of important immunologic pathways. However, there are other bacterial cell components, designated pathogen-associated molecular patterns (PAMPs) that enter the portal circulation, exerting immunologic reactions in the liver and beyond. Conversely, hepatitis B infection alters the gut microbiome. Lu et al. looked at the microbiome of adults with hepatitis B infections: asymptomatic carriers, individuals with chronic hepatitis B, and those with decompensated cirrhosis due to hepatitis B and normal controls. The microbiome of the individuals who were asymptomatic carriers revealed an increase in Faecalibacterium prausnitzii, Enterococcus faecalis and Enterobacteriaceae when compared to controls. The differences were even more pronounced for those with chronic hepatitis B and decompensated cirrhosis [11]. In summary, there is evidence that in hepatitis B infection, the gut microbiome affects changes in the liver and likewise the liver infection induces changes in the microbiome. It should be made clear that at this point these changes are associations. Whether they represent detrimental attacks on liver or a positive response to infection or merely neutral is not known.

4.3

Hepatitis C Virus (An RNA Virus)

Hepatitis C is another infection of global importance. It is estimated that 79 million people currently have hepatitis C [12]. Unlike hepatitis B vertical transmission is infrequent. There is presently no effective vaccine for hepatitis C; however, recently there has been a proliferation of direct-acting antivirals that can eradicate the virus. This opens up an opportunity to study the long-term effects of the viral infection after elimination of the virus itself. Clinical studies have revealed alterations in the microbiota of hepatitis C patients compared to controls. There is an increase in Enterobacteriaceae and Bacteriodetes abundance and a decreased abundance of Firmicutes [13]. Several hypotheses have been put forward to explain the alteration in the microbiota. The hepatitis virus invades not only hepatocytes but also gastric cells and B lymphocytes. B lymphocytes regulate synthesis of secretory IgA, which, in turn, modulates the intestinal microbiota and intestinal permeability. A second mechanism that could theoretically influence the microbiome is increased carbohydrate present in the gut of patients with hepatitis C, allowing fermenting bacteria species such as Prevotella

42

S. S. Baker and R. D. Baker

to overgrow. The increase in gastrointestinal carbohydrate is part of generalized malabsorption presumably driven by changes in bile acids [14]. An avenue of research into the pathogenesis of hepatitis C infection and its progression to cirrhosis has focused on the increased permeability of the GI tract and the translocation of bacteria and bacterial products. There is clear evidence that LPS and PAMPs gain access to the liver via the portal system. In a cross-sectional study that used normal values as reference values, Manteanu et al. found increased LPS and tissue necrosis factor-a (TNFa) levels in hepatitis C patients; however, they were unable to demonstrate a correlation between levels of LPS and TNFa and the degree of fibrosis present in the liver [15]. Taking advantage of the newly available treatment for hepatitis C that results in sustained viral response (SVR = no evidence of viremia for 12 months), Bajaj et al. hypothesized that following successful treatment of the virus there would be a return to a “normal” microbiome and a decrease in systemic inflammation. Contrary to their hypothesis, they could not demonstrate a change either in the microbiome or in the systemic inflammation, the two factors important in progression to cirrhosis and liver failure and hepatic cancer [16]. This finding suggests that patients who successfully clear the virus need to be monitored for further progression toward fibrosis. Possible manipulation of the microbiome may be efficacious in the future. As with other hepatotrophic viruses, hepatitis C virus infection is associated with changes in the microbiome and these changes may influence the course of the infection. Manipulation of the gut microbiota as a means of altering the course of hepatitis C seems worth pursuing, especially as the changed microbiome in hepatitis C appears to persist even after viral eradication. Prior to such attempts at treatment, wide-ranging, in-depth study of the microbiome of hepatitis C patients both before and after virus eradication must be undertaken.

4.4

Autoimmune Hepatitis

Autoimmune hepatitis (AIH) is a rare cause of chronic, progressive hepatic inflammation which exists as two phenotypes, type 1 and type 2; both occur in all ethnic groups with a female predominance. Type 1 is characterized by the presence of the auto antibodies ANA (antinuclear antibody) and/or ASMA (anti smooth muscle antibody), while type 2 is characterized by anti-liver-kidney microsomal antibody (ALKM) and liver cytosolic antibody (LC-1). Both types are associated with elevated serum IgG and elevated aminotransferases. There is a strong association with the alleles HLA-DR3 and HLA-DR4. Reports of prevalence range from 1 per 200,000 to 20 per 100,000 [17]. The pathogenesis is as yet unknown; however, it may involve molecular mimicry and promiscuous targeting. For example, ALKM cross reacts with hepatitis C virus [18]. However, changes in the microbiome may help initiate and sustain the disease. In accordance with the general pattern of gut microbiome and liver disease, gut dysbiosis, including a decrease in Bifidobactderium and Lactobacillus, and increased plasma LPS has

4 Gut Microbiota and Liver Injury (II): Chronic Liver Injury

43

been documented in AIH. Increased permeability is implied [19, 20]. Bacteria, viruses and their derivatives can stimulate Toll-like receptors and induce inflammasomes via a number of pathways. These alterations in liver and gut are not all detrimental. Some act as a defense [21]. The occurrence of AIH is clearly multifactorial. Genetics, gender, environment, diet and medications are among the factors that influence the pathogenesis of AIH. The gut microbiome and the gut–liver axis undoubtedly also play important roles. Realization of these roles may lead to new and novel avenues for treatment [22]. Primary Biliary Cholangitis Primary biliary cholangitis (PBC) is a chronic autoimmune disease occurring mostly in women and characterized by destructive cholangitis of intrahepatic small ducts by inflammatory cells, mainly lymphocytes and plasma cells. Elevated serum alkaline phosphatase combined with the presence of serum anti-mitochondrial autoantibodies or PBC-specific antinuclear antibodies and a high serum IgM are characteristics of the disease. In most patients the disease is progressive resulting in bile duct loss, fibrosis and cirrhosis. The mechanism of bile duct injury is unknown, but exposure to infectious microbes and/or xenobiotics has been suggested to initiate an immune reaction in genetically predisposed individuals [23, 24]. Treatment is with ursodeoxycholic acid (UDCA) as first line. If there is no or inadequate response, obetacholic acid (OCA), a synthetically modified bile acid and potent agonist of the farnesoid X nuclear receptor, can be offered. Farnesoid X receptor is a nuclear bile acid receptor found in the liver and gastrointestinal tract that regulates the expression of bile acid transporters. The impaired flow of bile acids results in their accumulation within the liver and contributes to liver disease. This suggests that the regulation of bile acid (BA) synthesis and homeostasis, as well as the resistance of the hepatocytes to BA, is important in the initiation and progression of PBC. The gut microbiome likely interacts with BA in a bi-directional manner and thus holds promise for diagnosis and treatment [25]. Lv et al. assessed the fecal microbiome in 42 subjects with early-stage PBC [26] and 30 healthy controls and found that the gut was depleted of potentially beneficial bacteria, including Acidobacteria, Lachnobacterium sp., Bacteroides eggerthii and Ruminococcus bromii, and enriched with opportunistic pathogens, Proteobacteria, Enterobacteriaceae, Neisseriaceae among others. Tang et al. studied the fecal microbiome in PBC before and after treatment with UDCA [27]. The investigators studied three groups: The first group consisted of 60 UDCA naïve patients with PBC and 80 healthy controls, the second group was an independent cohort used to confirm the findings in the first group and the third group consisted of 37 patients with PBC who were assessed prior to the treatment with UDCA and prospectively followed for 6 months after treatment. In the first cross-sectional study they found biodiversity was decreased in patients with PBC compared to the healthy controls and the eight genera that were enriched in the PBC patients were rare in the normal controls. Interestingly, they found that an unknown genus in the Enterobacteriaceae family was most significantly associated with PBC. This dysbiosis was partially

44

S. S. Baker and R. D. Baker

relieved after treatment with UDCA, suggesting a link between the gut microbiome and bile metabolism in PBC. The gut microbiome holds promise for patients with PBC by providing diagnostic markers and perhaps altering the course of the disease by targeting the gut microbiome, either directly, or through the homeostasis of bile acids. Primary Sclerosing Cholangitis Primary sclerosing cholangitis (PSC) is a progressive, cholestatic, inflammatory disease of the intra- and extra-hepatic ducts. It is characterized by a fibro-obstructive process. It typically, but not always, affects patients with inflammatory bowel disease. The only approved drug treatment is UDCA; but even this limited treatment option is controversial. Up to 25% of transplanted patients will develop recurrent disease. While the cause of PSC is not known, four hypotheses exist: (1) alteration in the bile acid pool (toxic bile hypothesis), (2) abnormal homing of lymphocytes (aberrant gut lymphocyte hypothesis), (3) abnormal pathways of fibrogenesis and (4) changes in microbiome leading to gut permeability (leaky gut hypothesis) [28]. The four hypotheses are not mutually exclusive. For this discussion only the “leaky gut” hypothesis will be addressed, although other mechanisms may also play roles. Evidence to support the leaky gut hypothesis include bacteria found in the bile of explanted livers, H. pylori in proximal hepatic ducts and serum antibodies against Chlamydia in the patients with PSC [29–31]. The suggestion that gut bacteria may play a role in PSC has triggered a number of trials employing oral antibiotics. The first report, 60 years ago, documented improvement using oral tetracycline [32]. This study was followed by several others using other antibiotics [33–35]. To date, the trials of antibiotic use have been small; however, results have been positive and larger, hence controlled studies are warranted. PSC is a serious, life-threatening disease which frequently occurs as a complication of IBD but also can stand alone. Oral antibiotics appear to offer a long-term treatment. The optimal antibiotic or combination of antibiotics needs to be determined. A definitive cure remains elusive.

4.5

Nonalcoholic Fatty Liver Disease

Nonalcoholic fatty liver disease (NAFLD) is the most common liver disease worldwide, affecting 20–30% of the general population. The hallmark of NAFLD is hepatic steatosis, which is characterized by fat accumulation in more than 5% of hepatocytes in the absence of excessive alcohol intake, or other cause of fat deposition such as viral infections, medications and genetic disorders. Nonalcoholic steatohepatitis (NASH) is the severe form of NAFLD. The histology of NASH includes inflammation and fibrosis that can progress to hepatic cirrhosis and hepatocellular carcinoma [36]. NASH is associated with obesity, insulin resistance and cardiovascular disease. Liver histology in NASH recalls that of alcoholic liver disease. Children [37] and adults [38] with NASH have elevated serum alcohol

4 Gut Microbiota and Liver Injury (II): Chronic Liver Injury

45

concentrations. Children also have increased hepatic levels of genes related to inflammation and fibrosis and most strikingly have increased gene transcription of the alcohol dehydrogenases [39]. Since there is no identifiable source of dietary alcohol in children [37], and the altered gut microbiome in NASH produces alcohol [37], likely the gut microbiome is a source of alcohol and this alcohol contributes to the initiation and/or the progression of the disease. Wu et al. [40] showed that diet strongly affects the gut microbiome composition and although changes can be seen within 24 h of a dietary change, it is long-term diet that correlates with the microbiome. The microbiome in NAFLD is similar to the microbiome in obesity. Firmicutes are increased and Bacteroidetes are decreased and the ratio is a potential phenotypic marker for obesity. However, in NASH, Firmicutes are decreased. Zhu et al. [37, 41] found that the increased abundance of Proteobacteria in the obese and NASH groups was mainly explained by the increased abundance of Enterobacteriaceae. Importantly, Enterobacteriaceae was the only abundant family (within the whole bacteria domain) exhibiting a significant difference between the obese group and the NASH group and most of the Enterobacteriaceae sequences belonged to Escherichia, which is the only abundant genus within the whole bacteria domain exhibiting a significant difference between the obese group and the NASH group. This genus had been shown to produce significant amounts of alcohol [42]. Alcohol is known to increase permeability of the gut [43] and allow bacterial products as well as toxins and infectious agents to relocate from the gut through the portal system assuring that these agents have direct access to the liver. This suggests the pathology could be driven by gut microbial generation of alcohol. The gut microbiota may contribute to NAFLD via energy salvage of diets. Short-chain fatty acids (SCFA), mostly identified as acetate, butyrate and propionate, are the major fermentation products of the gut microbes [41]. The concentrations of SCFA can reach 50–100 mM and provide a significant portion of energy. It is estimated that SCFA produced by the gut microbiota account for about 30% of energy extraction from the diet [44] and in obese individuals SCFA production is increased [45], suggesting that the gut microbiota could contribute to the development of steatosis by increasing delivery of SCFA to the liver. Although this hypothesis is appealing, others have suggested that SCFA are protective against the development of the metabolic syndrome [46, 47]. An area of gut–hepatocyte interaction that has not been explored in NASH is that of the metabolism of the D-amino acids. Among all domains of life, bacteria have the largest capacity to produce a wide variety of D-amino acids, major components of their cell walls, while mammals are thought to synthesize generally two kinds of D-amino acids, D-serine and D-aspartate, both of which are important in the neurological system. D-amino acid oxidase (DAO) is distributed in the proximal tubules in the kidney with the highest expression, hepatocytes (except not in mice), astrocytes in the hindbrain, neutrophils and epithelium of the small intestines [48]. The product of DAO is hydrogen peroxide, a strong oxidizing agent. Oxidative stress is considered to be a factor in the initiation and progression of NASH. Dr. Lixin Zhu (personal communication) noted that children with a histologic

46

S. S. Baker and R. D. Baker

diagnosis of NASH showed increased gene expression of DAO and peptidoglycan recognition protein 2 (PGLYRP-2, a liver secreted protein that recognizes and hydrolyzes bacterial peptidoglycans) compared to normal controls by microarray analysis, confirmed by quantitative real-time PCR analysis. As articulated by Dr. Zhu, these findings support his hypothesis that the bacterial cell wall components derived from the gastrointestinal tract microbiome cause increased oxidative stress via generation of hydrogen peroxide and contribute to the pathogenesis of NASH. The microbiome in patients with NASH appears to be important in the initiation and progress of NASH and represents an opportunity to alter the course of the disease if the microbiota can be appropriately manipulated.

4.6

Alcoholic Liver Disease

Alcohol-related liver disease consists of steatohepatitis in its mildest form and can progress to alcoholic hepatitis, fibrosis, cirrhosis and liver cancer. According to the American Liver Foundation [49], alcohol is the second most common cause of liver cirrhosis after hepatitis C infection in the USA. About 88,000 people die from alcohol-related causes annually, making alcohol the fourth leading preventable cause of death in the USA. It is likely that the gut microbiota play a role in alcoholic liver disease, but it is unclear if changes in the gut microbiome drive the progression of liver disease in some alcoholics or the ingestion of alcohol drives the changes in the microbiome. Alcohol itself causes increased permeability of the intestinal barrier as studied in CaCo2 cells [43]. The diet of alcoholics is not always of high quality [50], and as noted above, diet can influence the composition of the gut microbiome. Whether or not liver disease is present, patients with chronic alcohol abuse have increased gut permeability [51] as suggested by an increased level of plasma endotoxin compared to healthy controls. In cirrhosis a worsening of the clinical condition is thought to be associated with the translocation of gut bacteria and their products; treatment with antibiotics and or lactulose improve the condition [52]. Mutlu et al. [53] studied alcoholics with liver disease (19 subjects), who fulfilled the National Institute on Alcohol Abuse and Alcoholism and DSM-IV criteria for alcoholism, had a regular drinking history of at least 10 years, clinically significant liver disease and radiographic or histological evidence of liver disease. They also studied active alcoholics without liver disease (14 subjects), sober alcoholics without liver disease (15 subjects) and 18 healthy controls. They assessed biopsies from an unprepped flexible sigmoidoscopy. They did not assess stool specimens. They found that some, but not all subjects in the alcoholic groups were “dysbiotic” as compared to the control group. The dysbiotic group had no significant differences in phyla, but at the class level there was a uniform reduction of Bacteroidetes and Clostridia, and an increase in Bacilli and Gammaproteobacteria. The investigators concluded that alcohol alone appeared to affect the microbiome. They then

4 Gut Microbiota and Liver Injury (II): Chronic Liver Injury

47

correlated some clinical features of alcoholic liver disease with the microbiome and found that the dysbiotic group had a higher frequency of diabetes and a higher mean hemoglobin A1c. Further they showed that in the dysbiotic group of alcoholics, the microbiome was persistent and correlated with endotoxemia. In contrast, Tuomisto et al. [54] assessed stool samples from 13 autopsies of alcoholics with cirrhosis, 15 alcoholics without liver cirrhosis and 14 nonalcoholic men. Sterile samples were also obtained from their liver. Finally, stool samples from seven healthy volunteers were analyzed. Similar to the findings of Mutlu, stool from cirrhotics had a significant increase in Enterobactericaea. When compared to all other groups, the stool of alcoholic cirrhotics had significantly more gram negative Enterobactericaea, Enterobacter and bacteroides, but no differences in Bifidobacterium and Lactobacillus. Bacterial DNA was found in all autopsy liver samples and was not significantly different among the three groups. Bacterial DNA was also found in ascitic fluid, but all had a white blood count below 250. These results differed from those of Chen [55] who found decreased Bifidobacterium and increased Clostridium. The ability of the intestinal microbiome to promote the development of HCC is not well known, although it is likely that TLR4 signaling is involved.

4.7

Cirrhosis

The end stage of chronic liver disease is cirrhosis, characterized histologically by fibrosis and the formation of regenerative nodules. Fibrosis progresses with the length of time cirrhosis is present, culminating in liver failure, which results in death if liver transplantation is not possible. Life expectancy is reduced in cirrhosis, and cirrhosis is associated with several complications including variceal hemorrhage, ascites, spontaneous bacterial peritonitis, hepatic encephalopathy, hepatocellular carcinoma, hepatorenal syndrome and hepatopulmonary syndrome. Clinically cirrhosis can be compensated or decompensated and the time to progression to decompensation is not fixed. Compensated cirrhosis has few, if any, clinical symptoms and no major complications [56]. Decompensated cirrhosis is characterized by the development of any complication of cirrhosis, most commonly variceal bleeding, spontaneous bacterial peritonitis, hepatocellular carcinoma, hepatorenal syndrome and hepatopulmonary syndrome. In a study of 219 cirrhotic patients, 121 had compensated cirrhosis and 54 decompensated, and in 25 age-matched healthy controls Bajaj et al. [57] found the fecal microbiome in cirrhosis was characterized by a lack of diversity of organisms and a significant reduction in the beneficial organisms, Clostridiales XIV, Ruminococcaceae and Lachnospiraceae with a significant increase in the potentially toxic organisms, Enterococcaeae, Staphylococcaceae and Enterobacteriaceae compared to normal controls. As cirrhosis progressed, a reduction in Veillonellaceae and Porphyromonadaceae occurred. These investigators also showed that the microbiome in the cirrhotic patients who remained stable did not

48

S. S. Baker and R. D. Baker

change over the 4 months of the study. Based on these observations the investigators developed a cirrhosis–dysbiosis ratio (CDR), Lachnospiraceae + Ruminococcacea + Clostridiales incertase sedic XIV + Veillonellaceae divided by the sum of Enterobacteriaceae + Bacteroidaceae. They found that the CDR decreased as patients went from a compensated to a decompensated state characterized by hepatic encephalopathy or infection. They explained this change by an increase in the denominator organisms. Changes in the gut microbiome in cirrhosis can contribute to the explanation for some of the complications such as spontaneous bacterial peritonitis and hepatic encephalopathy. The beneficial organisms produce short-chain fatty acids that are a source of fuel for the colonocyte and may prevent or ameliorate the increased gut permeability that occurs with cirrhosis as well as reduce colonic inflammation. In addition, they can compete with pathogenic bacteria for nutrients and produce antibacterial peptides. These observations are intriguing, but do not necessarily describe the changes in the microbiome that are relevant to the progression of disease since the fecal microbiome differs significantly from that found at the sigmoid mucosal level [58]. Acute-on-chronic liver failure (ACLF) occurs in patients who have chronic liver disease and then experience a triggering event [59]. According to the 2013 European study [60], ACLF is characterized by an acute decompensation of cirrhosis (ascites, encephalopathy, gastrointestinal hemorrhage and/or bacterial infection) associated with organ/system failure(s) (liver, kidney, brain, coagulation, circulation and/or lung); may develop at any time during the course of the disease from compensated to long-standing decompensated cirrhosis; and is associated with a short-term (28-day) mortality rate ranging from 32 to 78.6%, depending on the number of organ failures despite standard supportive medical treatment. In Western countries, ACLF usually occurs in the context of a systemic inflammatory response, characterized by high leukocyte count and plasma C-reactive protein level, as a result of bacterial infections, severe alcoholic hepatitis or yet to be identified mechanisms [57]. Because of the high short-term mortality associated with ACLF any possible trigger that could be recognized early in the onset of ACLF could have an impact on the mortality. The systemic inflammatory response associated with ACLF suggests that microbiota may be involved with the development of ACLF especially because the gut microbiota plays an important role in the development of the immune system and its response [61]. To that end, Chen et al. [62] assessed gut microbiota in 79 Chinese patients with ACLF and followed them for four weeks after the onset of ACLF. They compared the results with 50 healthy controls. The investigators found a significant decrease in the microbial diversity and richness in the patients with ACLF; Proteobacteria phylum was significantly increased in nonsurvivors; at the family level Pasteurellaceae was significantly increased in nonsurvivors, while Lachnospiraceae was significantly lower. The 4-month longitudinal sampling showed that a new equilibrium was reached and remained stable after the onset of ACLF. Antibiotics showed a moderate impact on the composition of the gut microbiota. They concluded that gut dysbiosis occurs and the increased abundance of the family Pasteurella, a large and diverse species of Gram-negative bacteria, can serve as an independent predictor of mortality.

4 Gut Microbiota and Liver Injury (II): Chronic Liver Injury

49

In a North American study Bajaj et al. [57] collected stool from 181 patients with cirrhosis at the time of hospital admission and followed the patients for 30 days. They found that 8% of the subjects developed ACLF, 21% died. Beta-diversity was significantly different between those who developed ACLF and those who did not, but alpha-diversity was not. Similar to the observations of the Chinese group, Bajaj et al. found an increase in Proteobacteria (Enterobacteriaceae, Campylobacteriaceae and Pasteurellaceae) [58] in hospitalized patients with cirrhosis and was associated with an increased risk of extra-hepatic organ failure, ACLF, and death, independent of clinical factors. The fact that alpha-diversity did not differ over the course of the study suggests that the gut microbiome present on admission to the hospital could not be used to predict who would develop organ failure. Hepatic encephalopathy (HE) is a complication of cirrhosis. Cirrhosis is the end stage of practically all liver disease. The same pathogenic changes in the gut–liver axis evoked for other liver diseases have been suggested for HE. In the case of HE, ammonia appears to play a prominent role. Ammonia is predominantly produced by gut bacteria, where it gains entry to the portal system. In health the ammonia is metabolized and detoxified via the urea cycle. However, in end-stage liver disease, the liver can no longer processes ammonia sufficiently and serum and brain levels increase. There is poor correlation between arterial ammonia concentration and level of brain dysfunction, probably because ammonia’s neurotoxic effects are indirect via metabolites [63]. It is generally accepted that bacteria and bacterial byproducts play an important role in HE. Supporting this concept, the CDR index correlates with symptoms of HE [64]. Thus much of the effort in treating HE has focused on the microbiome. One of the more novel approached to the treatment of HE was by way of fecal microbiota transplantation. Weekly transplantation from a universal donor resulted in general improvement but the improvement was not durable after transplantations were discontinued [65]. Lactulose, a nonabsorbable pre biotic/laxative, has long been used in the treatment of overt HE. Its mechanisms of action are (1) as a laxative, reducing the number of gut bacteria and limiting the absorption of toxins, (2) as a prebiotic, stimulating bacterial proliferation and hence nitrogen uptake and (3) decreasing of glutamine uptake via inhibition of glutaminase [66]. Rifaximin, a broad spectrum, nonabsorbable antibiotic, has been shown to have beneficial effects on HE; it not only acts as an antimicrobial but also improves intestinal permeability [67].

4.8

Hepatocellular Carcinoma

Approximately 80–90% of cases of hepatocellular carcinoma (HCC) occur at the end stage of chronic liver disease, cirrhosis. Spontaneous HCC in the absence of liver disease is rare [68]. Because there are many possible causes of chronic liver disease, the tumors are heterogeneous and a possible microbiome pattern has been reported for cirrhosis as well as the specific chronic liver disease. So it would seem that there may not be specific signature for HCC. The gut microbiome in cirrhosis is

50

S. S. Baker and R. D. Baker

associated with inflammation, increased gastrointestinal permeability and the presentation of gut constituents, including those of the altered microbiome, to the liver. It is also likely that the microbiome in cirrhosis modulates the immune system, further disrupting normal functions. Cirrhosis often develops on a background of inflammation that causes cellular damage and the combination of these factors leads to hepatocellular mutations that cause HCC. In general, the gut microbiome in cirrhosis is described as exhibiting a decrease in beneficial microbes and an increase in those associated with systemic inflammation. For example, in one of the few human studies on the gut microbiota and HCC, Ponziani et al. [69] studied three groups of patients, those with NAFLD and HCC, those with NAFLD-related cirrhosis, but no HCC and healthy controls. They found that plasma levels of the interleukins IL8, IL13 and chemokines CCL3, 4 and 5 were higher in the HCC group and associated with activated status of circulating monocytes. The microbiome in all the cirrhotic subjects had increased abundance of Enterobacteriaceae and Streptococcus and a reduction in Akkermansia. Bacteroides and Ruminococcacea were increased in the HCC group, while Bifidobacterium was reduced. Akkermansia and Bacteroides were inversely correlated with fecal calprotectin, a marker for intestinal inflammation. These findings support the association of the microbiota and inflammation in HCC. In another human study of the microbiome in HCC, Ren et al. [70] assessed the microbiome in 486 stool samples from East, Central and Northwest China prospectively. They characterized the gut microbiome, identified microbial markers and constructed HCC classifier in 75 early HCC, 40 cirrhosis and 75 healthy controls. They found that fecal microbial diversity was increased from cirrhosis to early HCC and that the phylum Actinobacteria was increased in early HCC compared to cirrhosis. Thirteen genera including Gemmiger and Parabacteroides were enriched in early HCC compared to cirrhosis. Butyrate-producing genres were decreased; the lipopolysaccharide producing genera were increased. They identified 30 microbial markers that had diagnostic potential and they concluded that this was the first study to characterize gut microbiome in patients with HCC. These few human studies suggest that although that gut dysbiosis occurs in HCC and inflammation and immune dysfunction is involved in the initiation and progression of HCC, many unknown factors remain.

4.9

Conclusion

A common pattern is emerging that characterizes the interaction of the gut microbiome and liver disease. The pattern involves (1) variable change in the microbiome itself. The change seems to be disease-specific but tends to favor pro-inflammatory organisms. The driving force behind the change is not understood, but one possibility is that alteration in the bile acids reaching the GI tract directly or indirectly alters the gut microbiome makeup. Another common alteration is (2) increased permeability of the GI tract. Whether this is the result of the

4 Gut Microbiota and Liver Injury (II): Chronic Liver Injury

51

Inflammation Fibrosis Cancer

Hepatocytes Stellate cell

Dysbiosis

Kupffer cell

Leaky gut Bacterial products portal vein

Fig. 4.1 General mechanism of “gut-liver axis” in chronic liver injury

changed microbiome or due to the disease itself is unclear and may differ depending on the liver disease. In some cases the increase in permeability may be the result of altered secretory IgA; however, change in sIgA does not appear to account for the totality of the change in permeability. The increased permeability allows (3) translocation of bacterial components to the portal system. LPS, PAMPs, and so on gain access to heptaocytes, kupffer cells and stellate cells via the portal system where they promote inflammation and fibrosis through their influence on Toll-like receptors (TLR) and inflammatory mediators, IL6, IL17, TNFa and others (Fig. 4.1).

References 1. Xu J, Murphy SL, Kochanek KD, Bastian B, Arias E (2018) Deaths: final data for 2016. Natl Vital Stat Rep 67(2):1–75 2. Geneva WHO (2016) Global health estimates 2016: deaths by cause, age, sex, by country and by region 3. Zmora N, Zilberman-Schapira G, Suez J et al (2018) Personalized gut mucosal colonization resistance to empiric probiotics is associated with unique host and microbiome features. Cell 174(6):1388–1405.e21 4. Sartor RB (2012) Gut microbiota: diet promotes dysbiosis and colitis in susceptible hosts. Nat Rev Gastroenterol Hepatol 9(10):561–562 5. David LA, Maurice CF, Carmody RN et al (2014) Diet rapidly and reproducibly alters the human gut microbiome. Nature 505(7484):559–563 6. Dong TS, Gupta A (2019) Influence of early life, diet, and the environment on the microbiome. Clin Gastroenterol Hepatol 17(2):231–242 7. Hepatitis B. World Health Organization Fact Sheet 204 (2017) 8. Trepo C, Chan HL, Lok A (2014) Hepatitis B virus infection. Lancet 384(9959):2053–2063

52

S. S. Baker and R. D. Baker

9. Faith JJ, Guruge JL, Charbonneau M et al (2013) The long-term stability of the human gut microbiota. Science 341(6141):1237439 10. Chou HH, Chien WH, Wu LL et al (2015) Age-related immune clearance of hepatitis B virus infection requires the establishment of gut microbiota. Proc Natl Acad Sci U S A 112 (7):2175–2180 11. Lu H, Wu Z, Xu W et al (2011) Intestinal microbiota was assessed in cirrhotic patients with hepatitis B virus infection. Intestinal microbiota of HBV cirrhotic patients. Microb Ecol 61 (3):693–703 12. Hepatitis C. World Health Organization Fact Sheet (2018) 13. Aly AM, Adel A, El-Gendy AO et al (2016) Gut microbiome alterations in patients with stage 4 hepatitis C. Gut Pathog 8(1):42 14. Preveden T, Scarpellini E, Milic N et al (2017) Gut microbiota changes and chronic hepatitis C virus infection. Expert Rev Gastroenterol Hepatol 11(9):813–819 15. Munteanu D, Negru A, Radulescu M et al (2014) Evaluation of bacterial translocation in patients with chronic HCV infection. Rom J Intern Med 52(2):91–96 16. Bajaj JS, Sterling RK, Betrapally NS et al (2016) HCV eradication does not impact gut dysbiosis or systemic inflammation in cirrhotic patients. Aliment Pharmacol Ther 44(6):638– 643 17. Czaja AJ (2013) Autoimmune hepatitis in diverse ethnic populations and geographical regions. Expert Rev Gastroenterol Hepatol 7(4):365–385 18. Kerkar N, Choudhuri K, Ma Y et al (2003) Cytochrome P4502D6(193–212): a new immunodominant epitope and target of virus/self cross-reactivity in liver kidney microsomal autoantibody type 1-positive liver disease. J Immunol 170(3):1481–1489 19. Yuksel M, Wang Y, Tai N et al (2015) A novel “humanized mouse” model for autoimmune hepatitis and the association of gut microbiota with liver inflammation. Hepatology 62 (5):1536–1550 20. Lin R, Zhou L, Zhang J et al (2015) Abnormal intestinal permeability and microbiota in patients with autoimmune hepatitis. Int J Clin Exp Pathol 8(5):5153–5160 21. Czaja AJ (2016) Factoring the intestinal microbiome into the pathogenesis of autoimmune hepatitis. World J Gastroenterol 22(42):9257–9278 22. Czaja AJ (2017) Review article: next-generation transformative advances in the pathogenesis and management of autoimmune hepatitis. Aliment Pharmacol Ther 46(10):920–937 23. Tsuneyama K, Baba H, Morimoto Y et al (2017) Primary biliary cholangitis: its pathological characteristics and immunopathological mechanisms. J Med Invest 64(1.2):7–13 24. Goet JC, Hirschfield GM (2019) Guideline review: British Society of Gastroenterology/ UK-PBC Primary Biliary Cholangitis treatment and management guidelines. Frontline Gastroenterol 10(3):316–319 25. Li Y, Tang R, Leung PSC et al (2017) Bile acids and intestinal microbiota in autoimmune cholestatic liver diseases. Autoimmun Rev 16(9):885–896 26. Lv L-X, Fang D-Q, Shi D et al (2016) Alterations and correlations of the gut microbiome, metabolism and immunity in patients with primary biliary cirrhosis. Environ Microbiol 18 (7):2271–2286 27. Tang R, Wei Y, Li Y et al (2018) Gut microbial profile is altered in primary biliary cholangitis and partially restored after UDCA therapy. Gut 67(3):534–541 28. Cheung AC, Lazaridis KN, LaRusso NF et al (2017) Emerging pharmacologic therapies for primary sclerosing cholangitis. Curr Opin Gastroenterol 33(3):149–157 29. Tabibian JH, O’Hara SP, Lindor KD (2014) Primary sclerosing cholangitis and the microbiota: current knowledge and perspectives on etiopathogenesis and emerging therapies. Scand J Gastroenterol 49(8):901–908 30. Olsson R, Bjornsson E, Backman L et al (1998) Bile duct bacterial isolates in primary sclerosing cholangitis: a study of explanted livers. J Hepatol 28(3):426–432 31. Krasinskas AM, Yao Y, Randhawa P et al (2007) Helicobacter pylori may play a contributory role in the pathogenesis of primary sclerosing cholangitis. Dig Dis Sci 52(9):2265–2270

4 Gut Microbiota and Liver Injury (II): Chronic Liver Injury

53

32. Rankin JG, Boden RW, Goulston SJ et al (1959) The liver in ulcerative colitis; treatment of pericholangitis with tetracycline. Lancet 2(7112):1110–1112 33. Tabibian JH, Weeding E, Jorgensen RA et al (2013) Randomised clinical trial: vancomycin or metronidazole in patients with primary sclerosing cholangitis—a pilot study. Aliment Pharmacol Ther 37(6):604–612 34. Tabibian JH, Gossard A, El-Youssef M et al (2017) Prospective clinical trial of rifaximin therapy for patients with primary sclerosing cholangitis. Am J Ther 24(1):e56–e63 35. Davies YK, Cox KM, Abdullah BA et al (2008) Long-term treatment of primary sclerosing cholangitis in children with oral vancomycin: an immunomodulating antibiotic. J Pediatr Gastroenterol Nutr 47(1):61–67 36. Liu W, Baker RD, Bhatia T et al (2016) Pathogenesis of nonalcoholic steatohepatitis. Cell Mol Life Sci 73(10):1969–1987 37. Zhu L, Baker SS, Gill C et al (2013) Characterization of gut microbiomes in nonalcoholic steatohepatitis (NASH) patients: a connection between endogenous alcohol and NASH. Hepatology 57(2):601–609 38. Volynets V, Kuper MA, Strahl S et al (2012) Nutrition, intestinal permeability, and blood ethanol levels are altered in patients with nonalcoholic fatty liver disease (NAFLD). Dig Dis Sci 57(7):1932–1941 39. Baker RD, Greer FR (2010) Diagnosis and prevention of iron deficiency and iron-deficiency anemia in infants and young children (0–3 years of age). Pediatrics 126(5):1040–1050 40. Wu GD, Chen J, Hoffmann C et al (2011) Linking long-term dietary patterns with gut microbial enterotypes. Science 334(6052):105–108 41. Zhu L, Baker RD, Baker SS (2015) Gut microbiome and nonalcoholic fatty liver diseases. Pediatr Res 77(1–2):245–251 42. Dawes EA, Foster SM (1956) The formation of ethanol in Escherichia coli. Biochim Biophys Acta 22(2):253–265 43. Rao RK (1998) Acetaldehyde-induced increased in paracellular permeability in Caco-2 cell monolayer. Alcohol Clin Exp Res 22(8):1724–1730 44. Wostmann BS, Larkin C, Moriarty A et al (1983) Dietary intake, energy metabolism, and excretory losses of adult male germfree Wistar rats. Lab Anim Sci 33:46–50 45. Schwiertz A, Taras D, Schafer K et al (2010) Microbiota and SCFA in lean and overweight healthy subjects. Obesity (Silver Spring) 18(1):190–195 46. den Besten G, van Eunen K, Groen AK et al (2013) The role of short-chain fatty acids in the interplay between diet, gut microbiota, and host energy metabolism. J Lipid Res 54(9):2325– 2340 47. den Besten G, Bleeker A, Gerding A et al (2015) Short-chain fatty acids protect against high-fat diet-induced obesity via a PPARgamma-dependent switch from lipogenesis to fat oxidation. Diabetes 64(7):2398–2408 48. Sasabe J, Suzuki M (2018) Emerging role of D-amino acid metabolism in the innate defense. Front Microbiol 9:933 49. Foundation AL (2019) Liver disease statistics. https://liverfoundation.org/liver-diseasestatistics/#alcohol-related-liver-disease-and-cirrhosis. Accessed 30 June 2019 50. DiCecco SR, Francisco-Ziller N (2006) Nutrition in alcoholic liver disease. Nutr Clin Pract 21 (3):245–254 51. Elamin EE, Masclee AA, Dekker J et al (2013) Ethanol metabolism and its effects on the intestinal epithelial barrier. Nutr Rev 71(7):483–499 52. Riordan SM, Williams R (2010) Gut flora and hepatic encephalopathy in patients with cirrhosis. N Engl J Med 362(12):1140–1142 53. Mutlu EA, Gillevet PM, Rangwala H et al (2012) Colonic microbiome is altered in alcoholism. Am J Physiol Gastrointest Liver Physiol 302(9):G966–G978 54. Tuomisto S, Pessi T, Collin P et al (2014) Changes in gut bacterial populations and their translocation into liver and ascites in alcoholic liver cirrhotics. BMC Gastroenterol 14:40 55. Chen Y, Yang F, Lu H et al (2011) Characterization of fecal microbial communities in patients with liver cirrhosis. Hepatology 54(2):562–572

54

S. S. Baker and R. D. Baker

56. D’Amico G, Garcia-Tsao G, Pagliaro L (2006) Natural history and prognostic indicators of survival in cirrhosis: a systematic review of 118 studies. J Hepatol 44(1):217–231 57. Bajaj JS, Vargas HE, Reddy KR et al (2019) Association between intestinal microbiota collected at hospital admission and outcomes of patients with cirrhosis. Clin Gastroenterol Hepatol 17 (4):756–765.e3 58. Bajaj JS, Hylemon PB, Ridlon JM et al (2012) Colonic mucosal microbiome differs from stool microbiome in cirrhosis and hepatic encephalopathy and is linked to cognition and inflammation. Am J Physiol Gastrointest Liver Physiol 303(6):G675–G685 59. Jalan R, Gines P, Olson JC et al (2012) Acute-on chronic liver failure. J Hepatol 57(6):1336– 1348 60. Laleman W, Verbeke L, Meersseman P et al (2011) Acute-on-chronic liver failure: current concepts on definition, pathogenesis, clinical manifestations and potential therapeutic interventions. Expert Rev Gastroenterol Hepatol 5(4):523–537; quiz 37 61. Hooper LV, Littman DR, Macpherson AJ (2012) Interactions between the microbiota and the immune system. Science 336(6086):1268–1273 62. Chen YF, Guo J, Qian GR et al (2015) Gut dysbiosis in acute-on-chronic liver failure and its predictive value for mortality. J Gastroenterol Hepatol 30(9):1429–1437 63. Arora S, Martin CL, Herbert M (2006) Myth: interpretation of a single ammonia level in patients with chronic liver disease can confirm or rule out hepatic encephalopathy. CJEM 8 (6):433–435 64. Bajaj JS (2014) The role of microbiota in hepatic encephalopathy. Gut Microbes 5(3):397– 403 65. Kao D, Roach B, Park H et al (2016) Fecal microbiota transplantation in the management of hepatic encephalopathy. Hepatology 63(1):339–340 66. van Leeuwen PA, van Berlo CL, Soeters PB (1988) New mode of action for lactulose. Lancet 1(8575–8576):55–56 67. Zhang Y, Feng Y, Cao B et al (2015) Effects of SIBO and rifaximin therapy on MHE caused by hepatic cirrhosis. Int J Clin Exp Med 8(2):2954–2957 68. Yu LX, Schwabe RF (2017) The gut microbiome and liver cancer: mechanisms and clinical translation. Nat Rev Gastroenterol Hepatol 14(9):527–539 69. Ponziani FR, Bhoori S, Castelli C et al (2019) Hepatocellular carcinoma is associated with gut microbiota profile and inflammation in nonalcoholic fatty liver disease. Hepatology 69 (1):107–120 70. Ren Z, Li A, Jiang J et al (2019) Gut microbiome analysis as a tool towards targeted non-invasive biomarkers for early hepatocellular carcinoma. Gut 68(6):1014–1023

Chapter 5

Gut Microbiota and Lung Injury Ji-yang Tan, Yi-chun Tang, and Jie Huang

Abstract Gut microbiota are known to impact multiple organs including the lung. The cross talk between gut microbes and lungs, termed as the “gut–lung axis,” is vital for immune response and homeostasis in the airways. In this chapter, we summarized the coordinated development of microorganisms in the gut and lung, exogenous and endogenous factors related to the cross talk, the mechanisms of the gut–lung axis and their dysbiosis in lung diseases. Although the current understanding of the gut–lung axis is in its infancy, several gut microbiota-associated strategies have been designed to treat and prevent lung diseases.



Keywords Gut–lung axis Gut microbiota Cystic fibrosis Lung cancer



5.1

 Pneumonia  COPD  Asthma 

Introduction

Gut microbes, as the main component of human microbiota, benefit from a stable nutrient-balanced microenvironment and more importantly, they perform crucial roles in maintaining immune homeostasis in exchange [1]. Disturbance in the gut microbiota impacts not only gut, but also distal organs such as brain, liver, and lungs [2]. For example, about half of adults with inflammatory bowel diseases (IBD) and 33% with irritable bowel syndrome (IBS) have pulmonary involvement, J. Tan Peking Union Medical College, Chinese Academy of Medical Sciences, Plastic Surgery Hospital, Graduate School of Peking Union Medical College, Beijing, China Y. Tang The Second School of Clinical Medicine, Southern Medical University, Guangzhou, China J. Huang (&) Guangdong Provincial Key Laboratory of Translational Medicine in Lung Cancer, Guangdong Lung Cancer Institute, Guangdong Provincial People’s Hospital and Guangdong Academy of Medical Sciences, Guangzhou, China e-mail: [email protected] © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_5

55

56

J. Tan et al.

such as inflammation or impaired lung function, although many patients have no history of acute or chronic respiratory disease [3, 4]. Accumulating evidence indicates the association between the gut microbiota and lung immunity, referred to as the “gut–lung axis,” though the underlying mechanisms are still being studied and many questions remain unanswered. In this chapter, we summarize the influence of gut microbiota on respiratory tract in physiologic and pathologic conditions and the main mechanisms that have already been uncovered. Moreover, we highlight the potential therapeutic strategies of gut microbes for lung disease.

5.2

Structural and Functional Similarities and Differences Between the Gut and Lung

The epithelial of lung and gastrointestinal tract differs in structure. The epidermis of the lung is covered with pseudostratified ciliated columnar epithelial cells while gastrointestinal epidermis is covered with laminated squamous epithelium or single columnar epithelium [5]. However, they both develop from the endoderm and are exposed to outside. Both epithelia have abundant bacterial communities [6]. The balance between the bacterial communities and the barrier on the epithelial surface protects them from pathogens.

5.3

Coordinated Development of Microbiota in Gastrointestinal Tract (GIT) and Respiratory Tract

The microflora of human epidermis is not always invariable during the whole life. In fact, since the birth of the newborn, the microflora has undergone rapid dynamic change [7]. Certain factors of birth, such as the mode of delivery, the way of breastfeeding, gestational age, etc., only have an impact on the baby surface flora in the early infancy. Afterward, after six weeks of birth, the microflora of different parts of the body is gradually differentiated, and the influence of different parts on the microflora was greater than that from the mother. The flora is first determined by different body parts, then by individuals, last by age [8–10], and gradually formed a stable structure eventually. Meanwhile, microbes from different parts of the body are not isolated from each other. On the contrary, everybody is composed of a complex and interconnected ecology. The interrelated microbiome sites are either in direct contact, proximal contact, or similar in environmental exposure, thus providing mechanisms supporting similar microbiota and demonstrating high levels of microbial coexistence [11]. And the observation of a specific community state types (CST) at a given body site was predictive of contemporaneous CSTs at other body sites [7]. But how the body forms its unique flora is not fully understood. One study

5 Gut Microbiota and Lung Injury

57

explored the progression of human flora spatiotemporal interaction changes, which showed that the co-occurrence pattern between different body positions is very significant. Nasal CST 1, intestinal CST 2, laryngeal CST 1, and laryngeal CST 6 had the highest CST correlation (the number assigned to each CST indicates the overall frequency of occurrence at each respective site), and all trans-body pairs were positively correlated. The closest relationship shared between operational taxonomic units and coordinate is between the nose and throat, followed by the throat and intestines, and then the nose and intestines [7]. It indicates that the microflora between lung and gut influence each other during the development of differentiation. Not only in the healthy condition, but also in the pathological condition, the microflora of different parts of the body are correlated and co-developed. Serial analysis of the gut and respiratory microbiome in cystic fibrosis in infancy found that a group of microbes found in early life samples of the gut is also present in late life samples of the respiratory tract. This group of bacteria includes Enterococcus, Coprococcus, Escherichia, and Parabacteroides. Mapping populations over time showed that intestinal colonization occurred before they appeared in the respiratory tract. The microbial diversity of the airways and intestines increased significantly over time, and the diversity of the airways increased more acutely than this over time [12]. In mechanically ventilated patients, bacterial diversity tends to decrease, and a single population (usually opportunistic pathogens) may end up being an advantage. Intestinal permeability increases during sepsis and may also increase in other acute conditions involving the digestive tract, such as intestinal obstruction, ischemia, severe pancreatitis, or enterocytotoxicity due to chemotherapy [13]. Studies have shown that the lung microbiota of experimental sepsis mice and patients with confirmed ARDS has become rich in bacteria related to gut, among which Bacteroides, an anaerobic intestinal symbiotic bacteria, is the most common in both environments. Other intestinal symbionts, such as enterococcus faecalis and species belonging to the enterobacteriaceae family, were also found in the lungs of septic mice [14]. Therefore, lung microbiomes can be influenced by gut microbiomes in pathological condition.

5.4

Factors of Intestinal Microbiota that Influence the Lung

Dysbiosis of gut microbiota, with the exposure to smoke, antibiotic, or change of their diet, can lead to inflammation in the whole body and pathogenic bacterial infection, which can cause diseases at distal sites such as the lung [5]. The pathogenic factors can be divided into exogenous factors and endogenous factors. Generally, exogenous factors are comprised of smoking, antibiotics, diet, and probiotic bacteria. While endogenous factors mainly include the delivery mode of the newborns (Fig. 5.1).

58

J. Tan et al.

Fig. 5.1 Many exogenous and endogenous factors like diet, smoking, antibiotics, probiotic bacteria, and delivery mode can exert a significant influence on the composition of the gut microbiota. In some cases, it will lead to the dysbiosis of the intestinal tract and then cause allergic lung inflammation through the gut–lung axis. As a result, persistent inflammation of lung can finally increase the morbidity of asthma, COPD, pneumonia, cystic fibrosis, and even lung cancer

5.4.1

Exogenous Factors

1. Diet Obviously, diet can greatly change the composition of intestinal microbiota and plays an important factor in the reproduction and the stability of the intestinal microbiota. Animal experiments have demonstrated that change of the diet can affect the intestinal microbiota composition both rapidly and significantly [15]. Through the alteration of the metabolic production of intestinal microbiota, different diets can produce changes which can alter the host’s physiological state and therefore contribute to the development of lung injury.

5 Gut Microbiota and Lung Injury

59

Different diets can promote the growth of different pathogenic microbiota strains. Human’s daily food consists mainly of protein, fat, and carbohydrates. A high-fat diet may promote the reproduction of the pro-inflammatory gut microbiota and influence the intestinal permeability consequently [16]. High-fat diets can also play an important role in the chronic intestinal inflammation [17]. As for consumption of carbohydrates, for example, dietary fiber is a kind of polysaccharides which most of the intestinal microbiota can produce and use it to escape from being digested by human digestive juice. Carbohydrate metabolism of intestinal bacteria appears to be a highly cooperative process and the knowledge of it has been supported by numerous studies [18, 19]. When it comes to the digestion of protein, which is much less studied than the consumption of carbohydrates, the dietary protein intake pattern can in some ways influence the metabolic health and the outcomes of host’s intestinal bacteria [20, 21]. As mentioned above, the composition of the intestinal microbiota can change in response to deferent consumption of fat, sugar, fiber, and protein. In addition, the diet of a certain species can exert influence on viruses, fungi, and even archaea besides bacteria in human’s intestinal tract [22]. As for the connection between dietary composition and the lung disease, recent studies have found out that dietary fibers and its products called short-chain fatty acids (SCFA) can protect human from the allergic airway inflammation [23]. Mouse studies show that high-fiber diet can increase the amount of short-chain fatty acid (SCFA) and reduce the prevalence of asthma through an increase on the number of T-regulatory cell and strengthen their function [24]. In other words, short-chain fatty acid (SCFA) can prevent the aggravation of allergic lung inflammation caused by the dysbiosis of the intestinal tract [25]. While, on the contrary, high-fat diet is closely associated with the airway hyperreactivity (AHR), which can consequently lead to asthma [26]. Excess consumption of saturated fat and limited consumption of dietary fiber like fruit, grains, and vegetables have been greatly associated with high risk of inflammatory lung diseases [27]. The consumption of raw milk, however, can decrease the morbidity of asthma and allergies, which has been confirmed by mouse studies [28]. 2. Smoking In the past, smoking was associated with lung disease mostly due to its direct damage on the respiratory mucosa and the contribution of airway inflammation, while recent studies have confirmed that smoking can cause lung injury through the change of the composition of intestinal microbiome [29]. Biedermann L’s research, performed by the observation of 5 non-smokers and 5 continuing smokers, has found that smoke cessation can change the bacterial composition of human intestinal tract with a higher proportion of Firmicutes and Actinobacteria and a decrease of Bacteroidetes and Proteobacteria. Moreover, his study also showed that

60

J. Tan et al.

smoking cessation can cut down on the diversity of intestinal microbiota [30]. The mechanism that can explain the ways of smoking’s effect on intestinal microbiota includes changes in intestinal epithelial tight junctions, the enhancement of oxidative stress which can lead to free radical damage and the alterations of acid– base balance [31]. 3. Antibiotics As we all know, antibiotic treatment can greatly inhibit the development of certain kinds of bacteria in intestinal tract and alter the composition of intestinal microbiome, which can eventually increase the prevalence rate of the allergic pulmonary allergic diseases such as asthma. Although the studies in this field are limited and the results of which are partly inconsistent, recent research has confirmed that the antibiotic therapy conducted in the first 3 years of life has the ability of making children more susceptible to the onset of asthma (rather than the exacerbation of asthma) [32]. Furthermore, not only the antibiotic treatment in the early life but also the antibiotic use in the pregnancy can play an important role in the development of allergic disease in lung. According to a case-control analysis, if a pregnant woman received antibiotic treatment in the third trimester of pregnancy, her baby would be more likely to get asthma at pre-school stage [33]. The association between antibiotic exposure and lung disease can also enlighten new therapies toward asthma. Allergen immunotherapy is confirmed to have a protective effect of keeping people who has been treated with antibiotics from allergic asthma [34]. 4. Probiotic Bacteria Probiotic bacteria are important components of the intestinal microbiome. They have been considered to provide healthy benefit and maintain the acid–base balance in the intestinal tract for a long time. To date, numerous researches have been conducted to interpret how probiotic bacteria can exert impact on the chronic respiratory inflammation, though for different bacterial species the results may seem inconsistent. As has been shown in many researches, the intake of lactobacillus rhamnosus GG (LGG) tablets can play an anti-inflammatory role in the inhibition of OVA-induced chronic allergic respiratory disease and therefore offers a possible treatment of airway inflammation [35]. Similarly, Bifidobacterium lactis have the same effect on alleviating airway inflammation by reducing the ratio of Th17 and Treg cells [36]. In addition, according to mouse studies, Bifidobacterium breve M-16 V (Bb) are able to decrease the number of Th2 cells and intestinal DC and contribute to the increase of Treg, thus easing the inflammatory symptoms [37]. However, the studies mentioned above are largely based on in vitro experimental data, which means that there is no firm evidence of the protective effect of the probiotics on lungs [38]. Further studies are still required.

5 Gut Microbiota and Lung Injury

5.4.2

Endogenous Factors

5.4.2.1

Delivery Mode

61

Many studies have revealed the relationship between the birth mode of the newborn and the morbidity of allergic lung disease. Children who were born by cesarean section, compared to those who were born by natural section, are more likely to get asthma when they grow up, partly because of the decrease of the diversity of their intestinal microbiota and the alteration of the composition of the bacteria in their gastrointestinal tract [39]. Likewise, a study conducted in Vietnam and India also confirmed the result [40]. In addition, another research performed by Sevelsted et al. has pointed out that the cesarean delivery’s effect on the morbidity of asthma can be even more pronounced if the operation is conducted before the rupture of fetal membrane [41].

5.5

The Mechanism of the Gut–Lung Axis

Recent studies have reached a deeper understanding of the complex interrelationship between the gut and lung, which can be mediated in mainly two ways— bacterial products in the circulatory system and the immune cells in the lymph node of gut and lung (Fig. 5.2). In one way, cell wall fragments, the protein parts of dead bacteria and surviving bacteria generated in the intestinal tract can travel along the mesenteric lymph node to systemic circulation and subsequently enter the pulmonary circulation, which will lead to the stimulation of DCs, macrophage, T cells, and the influx of neutrophil. As a result, lung injury and inflammation caused by uncontrolled macrophage and neutrophil activation will be observed. Another way of the gut and lung to exert influence on each other is the migration of T and B cells. First, the bacterial products and the surviving bacteria will be transferred to the mesenteric lymph node (MLN) by macrophage and DCs after phagocytosis and elimination. In the MLN, translocated bacteria and their parts could prime naïve B cells and T cells, thereby activating the differentiation of B cells to plasma cells. Plasma cells, known for its ability to produce antigen-specific immunoglobulin, can cut down the production of anti-inflammatory molecules like IL-10, leading to the priming and the differentiation of T cells. After that, some of the T cells will migrate out of the gut-associated lymphatic tissue (GALT) and finally reach the lung epithelium because the DCs of lung can influence the expression of gut-homing integrin 47 and CCR9, thus activating local APC and T cells. The activation of the immune cells mentioned above is able to improve pulmonary immune response and consequently increase the pathogen clearance and antitumor activity of lungs.

62

J. Tan et al.

Fig. 5.2 The pathways of the interrelationship between the gut and lung. The bacteria in the lung can improve immune response of the gut or cause chronic inflammation and cancer through the lymphatic and blood flow (the blue arrows). Similarly, microorganisms in the gut also could change the immune state of the lungs in the same way (the yellow arrows). DC: dendritic cell; IEC: intestinal epithelial cell; APC: antigen-presenting cell; LLN: lung lymph node; MLN: mesenteric lymph node; GALT: gut-associated lymphatic tissue

Likewise, the microorganisms in the lung can improve immune response of the gut, but they also may cause chronic inflammation and cancer through the lung lymphatic system and the blood flow. Additionally, DCs in gut can increase the amount of lung-homing integrin-CCR4 [42, 43].

5.6

Intestinal Dysbiosis and Lung Disease

To date, many studies have shown that various pulmonary diseases are closely related to the dysbiosis of the intestinal tract and the consequent changes of immune response. According to recent researches, the pulmonary diseases mainly include

5 Gut Microbiota and Lung Injury

63

asthma, COPD, pneumonia, cystic fibrosis, and even lung cancer, which is usually accompanied by a deceased intestinal microbial diversity and the disorder of immune system [44].

5.6.1

Asthma

Asthma, mainly caused by the hyper-responsiveness of airway, is widely known to be one of the most common chronic allergic respiratory diseases in the world. To date, there are emerging evidences confirming the connection between asthma and the alteration of the composition of intestinal microbiome [45]. A study performed by Melli et al. compared the composition of the intestinal microorganism of allergic people to that of non-allergic people and revealed the differences between them. According to the experiment data, lower microbiome diversity and a predominance of Firmicutes can be observed in the children with asthma [46]. As mentioned above, early life exposures such as birth mode and antibiotic treatment can change the risk of asthma in childhood. While further studies conducted among Canadian infants at the age of 3 mouth interpret the relationship between the increased morbidity of asthma caused by the alteration of intestinal bacteria and the feeding mode and point out the protective effect of the direct breastfeeding [47]. Similarly, a nationwide survey performed in Japan reveals that direct breastfeeding can also inhibit the risk of asthma of children between the age of 6 and 42 months [48]. As to the interaction between the intestinal microbiome and the onset of asthma, mouse researches showed strong evidence of the view that the intestinal microbiome composition can exert impact on the function of the immune system by the way of the Treg cells’ expansion and the reduction in the IgE production [49]. What’s more, the alteration of Th2 response can also be observed to interpret the mechanism of how can intestinal microbiome be connected to asthma [38]. In addition, the studies supporting that microbiome in intestinal tract plays an important role in asthma enlighten the development of new therapies of asthma like administration of few special bacterial strains [49]. While further researches are required to evaluate the exact curative effect on the children of different ages.

5.6.2

COPD

Chronic obstructive pulmonary disease (COPD), characterized by increasing breathlessness, is a leading cause of morbidity and mortality worldwide. The mechanism of its occurrence and development has been extensively studied. It is confirmed that the pneumonia bacteria can play an important role in its development and deterioration. As is mentioned above, an increasing number of recent studies hold the view that the microbiome of our body can be considered as a whole and the composition of pneumonia bacteria can be altered by the change of

64

J. Tan et al.

intestinal microbiota. Thus, the bacteria in the intestinal tract can exert influence for either good or bad on the occurrence and the development of lung injury. Additionally, more and more studies suggest that the altered intestinal microbiome composition can be considered as an important component of the multisystem disorder in COPD patients [50, 51]. Research conducted by Ottiger and Nickler has also proved that the increased level of gut microflora-dependent metabolite trimethylamine-N-oxide (TMAO) in the circulation can heighten the long-term mortality of the COPD patients [52]. Still, further experiments are needed to interpret the mechanism of the interaction between COPD and the gut bacteria.

5.6.3

Pneumonia

Many recent studies have confirmed that the intestinal microorganism can play a protective role in the host defense against pneumonia. During the neonatal period, the exposure to gut commensal bacteria and the influx of the group 3 innate lymphoid cells (ILC3) produced by IL-22 induce the immunity of the host lung against pulmonary infections [53, 54]. According to recent studies, relevant pathogens include Pneumocystis pneumonia, Klebsiella pneumoniae, and Pneumococcal pneumonia. Mouse experiments made by Schuijt et al. have confirmed that the intestinal microbiota are able to resist pneumococcal pneumonia through an enhanced capacity of alveolar macrophages to phagocytose the pathogenic bacteria [55]. Likewise, gut microflora dysbiosis caused by chronic alcohol consumption can weaken the pulmonary host defense against Klebsiella pneumoniae through intestinal T cell sequestration as well [56]. In addition, besides bacterial pneumonia, the Interaction between intestinal and pulmonary microorganisms can also be regarded as an important factor in host defense against fungal pathogens like Pneumocystis carinii [57].

5.6.4

Cystic Fibrosis

Cystic fibrosis (CF) is an inherited metabolic disorder, the chief symptom of which is the production of a thick, sticky mucus that clogs the respiratory tract and the gastrointestinal tract. The occurrence of cystic fibrosis is always accompanied by intestinal inflammation and dysbiosis. Increased Firmicutes, decreased Bacteroidetes, and a significant change of bacterial metabolites can be observed in the intestinal tracts of the CF patients [58]. In the development of CF, the bacterial imbalance in digestive tract can not only contribute to inflammation through the decrease of short-chain fatty acids (SCFA) but also cause higher incidence rate of gastrointestinal malignancy by increasing the amount of carcinogenic bacteria

5 Gut Microbiota and Lung Injury

65

[59, 60]. While the mechanism of the interactions among bacterial imbalance, gut inflammation, and the gastrointestinal malignancy in CF patients still needs further study.

5.6.5

Lung Cancer

According to previous research, fruits, hereditary and environmental factors play an important role in the occurrence and the development of the lung cancer. While the involvement of the microflora has also been confirmed by recent researches. By using the Lewis lung cancer mouse model, some researchers have found that the intestinal commensal microbiota have anti-lung cancer effect and the consumption of probiotics can contribute to the antigrowth and proapoptotic effects of cisplatin [61]. While, on the contrary, the dysbiosis in intestinal tract and airway can induce a statement of inflammation and increase the risk of developing lung cancer, the underlying mechanism is still not clear [62]. In addition, the intestinal tract is an important place for immune cells to develop, which influence the immunity of not only the intestine but also other distal organs like lung [43]. Thus, with the emerging concept of more systemic effects of the intestinal bacteria, many teams have made a huge leap in confirming the important role intestinal microbiota plays in the immunotherapy of the lung cancer [63], which will be interpreted in detail in the next section.

5.7

Intestinal Microbiome-Associated Strategies for the Treatment and Prevention of Lung Disease

Because of the link existing between gut microbiome and the lung injury, targeting intestinal flora to treat lung injury has become a strategy.

5.7.1

The Asthma and the Gut Microbiome

Asthma and allergic airway inflammation have been linked to bacterial exposure. However, it has been found that intestinal flora plays an indispensable role in the occurrence and development of asthma. A study which performed a longitudinal comparison of stool samples shows that the early development of intestinal microorganism is unique but plastic in infants at high risk for asthma [64]. Controlling the airway and gut microbiota may be a strategy to prevent asthma from starting and worsening [64]. Now, probiotics are widely used to prevent asthma, but there is no sufficient information to suggest the use of specific probiotics in allergy

66

J. Tan et al.

and asthma prevention [38]. Though Lactobacillus rhamnosus GG (LGG) supplementation may partially rescue endogenous disparities in gut microbial distributions, it appears unable to compensate for reduced exposure to environmental microbes [64]. For high-risk infants, a randomized controlled trial finds that supplementation with Lactobacillus rhamnose in the first 6 months of life does not seem to prevent eczema or asthma at age 2 [65]. Another randomized, placebo-controlled trial supports lactic acid bacteria is beneficial to children with asthma. They found that both Lactobacillus paracasei and Lactobacillus fermentum can reduce the severity of asthma in school-age children and improve the control of asthma [66]. How to identify the ideal effective bacteria to shape the early infants flora to prevent asthma remains to be studied. In addition to probiotics, dietary fiber is an important way to change the intestinal flora. Dietary fiber has been well demonstrated to be metabolized by local intestinal microflora into short-chain fatty acids (SCFAs), which can inhibit the inflammatory pathways of macrophages and dendritic cells (DC), promote the development of regulatory T (T reg) cells, and maintain the integrity and health of intestinal epithelium. Different amounts of dietary fiber can alter allergic airway responses and alter susceptibility to inflammatory diseases [67].

5.7.2

The Pneumonia and the Gut Microbiome

A study depleted the gut microbiota in C57BL/6 mice and subsequently infected them intranasally with S. pneumonia, and they found that the gut microbiota enhances primary alveolar macrophage function which highlights the possibility that broad-spectrum antibiotics and disruption of the intestinal microbiota may diminish innate immune defence to infection [55]. It indicates that novel therapies to treat severe pneumonia could focus on the gut–lung axis in bacterial infection. A prospective, randomized, double-blind, placebo-controlled trial in 146 mechanically ventilated patients at high risk for VAP found that in addition to routine care, patients who received probiotics twice daily received fewer antibiotic days of VAP therapy [68]. A study on mice showed that LGG treatment can significantly improve intestinal permeability after Pseudomonas aeruginosa pneumonia and normalize claudin2 expression, suggesting that probiotic treatment can reduce ventilator-associated pneumonia, especially in critically ill patients with pseudomonas aeruginosa pneumonia [69]. These evidences suggest that intestinal flora and related products can improve respiratory immunity during the process of pneumonia infection, which is also one of the directions to be considered in the future treatment of pulmonary infection. But when it comes to meta-analysis, previous studies failed to come to an agreement. Even though probiotic group showed a trend toward lower incidence of VAP, the preventative effect was not statistically valid [70]. The meta-analysis comprising 10 RCTs (1,218 patients) argued that administration of probiotics

5 Gut Microbiota and Lung Injury

67

reduced the incidence of ICU-acquired pneumonia and was associated with a shorter stay in the ICU [71]. Clinical and statistical heterogeneity and inaccurate estimates preclude strong clinical recommendations. Further research on probiotics in critically ill patients is necessary [72]. And the specific mechanisms of benefit are complex and not fully described.

5.7.3

The Immunotherapy and the Gut Microbiome

Immunotherapy has been a great success in cancer treatment, especially in non-small cell lung cancer fields [73]. The currently marketed checkpoint inhibitors are monoclonal antibodies targeting cytotoxic T lymphocyte-associated protein 4 (CTLA4) or the programmed death 1 (PD1) located on T cell surfaces, or its ligand, programmed death ligand 1 (PD-L1), expressed by the APCs [74]. In mice and patients, T cell responses specific for B. thetaiotaomicron or B. fragilis were associated with the efficacy of CTLA-4 blockade. Fecal microbial transplantation from humans to mice confirmed that treatment of melanoma patients with antibodies against CTLA-4 favored the outgrowth of B. fragilis with anticancer properties [75]. Sequencing of the 16S ribosomal RNA showed that bifidobacterium had antitumor effects. Oral administration of Bifidobacterium alone can improve tumor control to a certain extent in programmed cell death protein 1 ligand1 (PD-L1)-specific antibody therapy [76]. In a Chinese cohort of patients with advanced non-small cell lung cancer, antibiotics treatment was significantly associated with the attenuated clinical results of immune checkpoint inhibitors based on anti-PD-1 [77]. This study provides us with a positive signal that to improve the efficacy of PD-1, we may start from the human flora. Microbiota can influence the outcome of cancer immunotherapy in a more specific way. Some symbiotic bacteria may suppress tumor immunity by tilting the balance of the immune subsets to the inhibitory phenotypes such as Tregs and MDSCs. In addition, bacterial metabolites with immunosuppressive properties may be released into the circulation to promote the function of TdLN and TME immunosuppressive cells. Chronic inflammation caused by persistent PAMPs/ MAMPs stimulation or epithelial injury may also eventually lead to immunosuppression. The immunostimulatory effects of the gut microbiota could be mediated by augmented antigenicity, adjuvanticity, or bystander T cell activation [78]. Tanoue et al., firstly, isolated a combination of 11 strains from healthy human feces which can improve the effectiveness of tumor immunotherapy without causing inflammation [79]. These species are present in the human body in low abundance. Although the exact mechanism is unclear, this study provides an idea that some rare strains have a strong potential to promote immunotherapy effect. Zitvogeld et al. found patients treated with antibiotics (which destroyed the gut microbiome) relapsed quickly and did not live long. The researchers then analyzed the differences between the intestinal bacteria of the PD-1 antibody responders and the non-responders. It was found that among responders, the bacterium

68

J. Tan et al.

Akkermansia muciniphila played an important role. Oral supplementation with A. muciniphila after fecal microbiota transplantation with nonresponder feces restored the efficacy of PD-1 blockade in an interleukin-12-dependent manner by increasing the recruitment of CCR9+ CXCR3+ CD4+ T lymphocytes into mouse tumor beds [80]. The strains reported in the above studies are different, but all of them suggest that human flora is very important in the process of immunotherapy. The intestinal tract contains rich flora, and the adjustment of intestinal flora is bound to be the focus of future research on the efficacy of immunotherapy. But more detailed studies and clinical trials are needed on the specific strains and effects in different people.

5.8

Conclusion

With the increasing understanding of gut microbiota, more and more attentions are paid to gut–lung axis that has important role in health and disease. The microbiota of the gut and the microbiota of the lungs interact with each other and restrict each other. However, it is unlikely this interaction is enough responsible for the functions of the microbiota. The cross talk between microbiome and the host is complex and it is widely accepted that the microbiota, especially the gut microbiota, modulate innate and adaptive immune responses in a global way [5, 81]. Dysbiosis in gut microbiota affects local and distal organs, but it is still unclear whether changes in gut microbes affect distal sites equally, and whether distal sites then would also interplay with each other. A further understanding of the systemic effects of gut microbiota requires more broad studies and clinical intervention trials. Several factors are reported to be involved in affecting gut–lung axis, including smoking, antibiotics, diet, probiotic bacteria, and the delivery mode of the newborns, which could be divided into exogenous factors and endogenous factors. But thus far, we could only say the above are related factors, and the significance of each factor is unknown. A more precise assessment model needs to be set up in the future by weighing in significant related factors. More and more evidence supports that bacterial products and immune cells mediate gut–lung cross talk. The microorganisms could be identified as beneficial and pathogenic species according to their immunomodulatory effects [81, 82]. But the mechanism is so complex that our current understanding is only the tip of the iceberg. Further studying the characteristics of the microorganisms will provide us with new insights into ways to manipulate them. In summary, it is clear that gut microbiota is closely associated to respiratory tract, and the specific species are able to influence the pathogenesis of respiratory diseases, which might be mediated by reshaping systemic immune responses. Further researches will help to present the whole picture of gut–lung axis and to identify new and effective therapeutics.

5 Gut Microbiota and Lung Injury

69

References 1. Krajmalnik-Brown R et al (2012) Effects of gut microbes on nutrient absorption and energy regulation. Nutr Clin Pract 27(2):201–214 2. Shreiner AB, Kao JY, Young VB (2015) The gut microbiome in health and in disease. Curr Opin Gastroenterol 31(1):69–75 3. Keely S, Talley NJ, Hansbro PM (2012) Pulmonary-intestinal cross-talk in mucosal inflammatory disease. Mucosal Immunol 5(1):7–18 4. Yazar A et al (2001) Respiratory symptoms and pulmonary functional changes in patients with irritable bowel syndrome. Am J Gastroenterol 96(5):1511–1516 5. Budden KF et al (2017) Emerging pathogenic links between microbiota and the gut-lung axis. Nat Rev Microbiol 15(1):55–63 6. Mjosberg J, Rao A (2018) Lung inflammation originating in the gut. Science 359(6371): 36–37 7. Grier A et al (2018) Neonatal gut and respiratory microbiota: coordinated development through time and space. Microbiome 6(1):193 8. Costello EK et al (2009) Bacterial community variation in human body habitats across space and time. Science 326(5960):1694–1697 9. Cho I, Blaser MJ (2012) The human microbiome: at the interface of health and disease. Nat Rev Genet 13(4):260–270 10. Costello EK et al (2013) Microbiome assembly across multiple body sites in low-birthweight infants. MBio 4(6):e00782-13 11. Faust K et al (2012) Microbial co-occurrence relationships in the human microbiome. PLoS Comput Biol 8(7):e1002606 12. Madan JC et al (2012) Serial analysis of the gut and respiratory microbiome in cystic fibrosis in infancy: interaction between intestinal and respiratory tracts and impact of nutritional exposures. MBio 3(4):e00251-12 13. Mukherjee S, Hanidziar D (2018) More of the gut in the lung: how two microbiomes meet in ARDS. Yale J Biol Med 91(2):143–149 14. Dickson RP et al (2016) Enrichment of the lung microbiome with gut bacteria in sepsis and the acute respiratory distress syndrome. Nat Microbiol 1(10):16113 15. Zhang M, Yang XJ (2016) Effects of a high fat diet on intestinal microbiota and gastrointestinal diseases. World J Gastroenterol 22(40):8905–8909 16. Bibbo S et al (2016) The role of diet on gut microbiota composition. Eur Rev Med Pharmacol Sci 20(22):4742–4749 17. Gulhane M et al (2016) High fat diets induce colonic epithelial cell stress and inflammation that is reversed by IL-22. Sci Rep 6:28990 18. Salonen A, de Vos WM (2014) Impact of diet on human intestinal microbiota and health. Annu Rev Food Sci Technol 5:239–262 19. Goldsmith JR, Sartor RB (2014) The role of diet on intestinal microbiota metabolism: downstream impacts on host immune function and health, and therapeutic implications. J Gastroenterol 49(5):785–798 20. Holmes AJ et al (2017) Diet-microbiome interactions in health are controlled by intestinal nitrogen source constraints. Cell Metab 25(1):140–151 21. Beaumont M et al (2017) Epithelial response to a high-protein diet in rat colon. BMC Genomics 18(1):116 22. Albenberg LG, Wu GD (2014) Diet and the intestinal microbiome: associations, functions, and implications for health and disease. Gastroenterology 146(6):1564–1572 23. He Y et al (2017) Gut-lung axis: the microbial contributions and clinical implications. Crit Rev Microbiol 43(1):81–95 24. Thorburn AN et al (2015) Evidence that asthma is a developmental origin disease influenced by maternal diet and bacterial metabolites. Nat Commun 6:7320

70

J. Tan et al.

25. Cait A et al (2018) Microbiome-driven allergic lung inflammation is ameliorated by short-chain fatty acids. Mucosal Immunol 11(3):785–795 26. MacDonald KD et al (2017) Maternal high-fat diet in mice leads to innate airway hyperresponsiveness in the adult offspring. Physiol Rep 5(5):e13082 27. Wood LG (2017) Diet, obesity, and asthma. Ann Am Thorac Soc 14(Supplement_5), S332– S338 28. Sozanska B (2019) Raw cow’s milk and its protective effect on allergies and asthma. Nutrients 11(2):e469 29. Capurso G, Lahner E (2017) The interaction between smoking, alcohol and the gut microbiome. Best Pract Res Clin Gastroenterol 31(5):579–588 30. Biedermann L et al (2013) Smoking cessation induces profound changes in the composition of the intestinal microbiota in humans. PLoS One 8(3):e59260 31. Savin Z et al (2018) Smoking and the intestinal microbiome. Arch Microbiol 200(5):677–684 32. Ahmadizar F et al (2017) Early life antibiotic use and the risk of asthma and asthma exacerbations in children. Pediatr Allergy Immunol 28(5):430–437 33. Mulder B et al (2016) Antibiotic use during pregnancy and asthma in preschool children: the influence of confounding. Clin Exp Allergy 46(9):1214–1226 34. Woehlk C et al (2018) Allergic asthma is associated with increased risk of infections requiring antibiotics. Ann Allergy Asthma Immunol 120(2):169–176.e1 35. Wu CT et al (2016) Effects of immunomodulatory supplementation with Lactobacillus rhamnosus on airway inflammation in a mouse asthma model. J Microbiol Immunol Infect 49(5):625–635 36. Liu Q, Jing W, Wang W (2018) Bifidobacterium lactis ameliorates the risk of food allergy in chinese children by affecting relative percentage of Treg and Th17 cells. Can J Infect Dis Med Microbiol 2018:4561038 37. de Kivit S et al (2017) Dietary, nondigestible oligosaccharides and Bifidobacterium breve M-16V suppress allergic inflammation in intestine via targeting dendritic cell maturation. J Leukoc Biol 102(1):105–115 38. Mennini M et al (2017) Probiotics in asthma and allergy prevention. Front Pediatr 5:165 39. Chu S et al (2017) Cesarean section without medical indication and risk of childhood asthma, and attenuation by breastfeeding. PLoS One 12(9):e0184920 40. Lavin T, Franklin P, Preen DB (2017) Association between caesarean delivery and childhood asthma in India and Vietnam. Paediatr Perinat Epidemiol 31(1):47–54 41. Sevelsted A, Stokholm J, Bisgaard H (2016) Risk of asthma from cesarean delivery depends on membrane rupture. J Pediatr 171:38–42.e1–4 42. Bingula R et al (2017) Desired turbulence? Gut-lung axis, immunity, and lung cancer. J Oncol 2017:5035371 43. McAleer JP, Kolls JK (2018) Contributions of the intestinal microbiome in lung immunity. Eur J Immunol 48(1):39–49 44. Marsland BJ, Trompette A, Gollwitzer ES (2015) The gut-lung axis in respiratory disease. Ann Am Thorac Soc 12(Suppl 2):S150–S156 45. Kang YB, Cai Y, Zhang H (2017) Gut microbiota and allergy/asthma: from pathogenesis to new therapeutic strategies. Allergol Immunopathol (Madr) 45(3):305–309 46. Melli LC et al (2016) Intestinal microbiota and allergic diseases: a systematic review. Allergol Immunopathol (Madr) 44(2):177–188 47. Klopp A et al (2017) Modes of infant feeding and the risk of childhood asthma: a prospective birth cohort study. J Pediatr 190:192–199.e2 48. Yamakawa M et al (2015) Breast-feeding and hospitalization for asthma in early childhood: a nationwide longitudinal survey in Japan. Public Health Nutr 18(10):1756–1761 49. Ozturk AB et al (2017) The potential for emerging microbiome-mediated therapeutics in asthma. Curr Allergy Asthma Rep 17(9):62 50. Chotirmall SH et al (2017) Microbiomes in respiratory health and disease: an Asia-Pacific perspective. Respirology 22(2):240–250

5 Gut Microbiota and Lung Injury

71

51. Shukla SD et al (2017) Microbiome effects on immunity, health and disease in the lung. Clin Transl Immunol 6(3):e133 52. Ottiger M et al (2018) Gut, microbiota-dependent trimethylamine-N-oxide is associated with long-term all-cause mortality in patients with exacerbated chronic obstructive pulmonary disease. Nutrition 45:135–141.e1 53. Gray J et al (2017) Intestinal commensal bacteria mediate lung mucosal immunity and promote resistance of newborn mice to infection. Sci Transl Med 9(376):eaaf4912 54. Tamburini S, Clemente JC (2017) Gut microbiota: neonatal gut microbiota induces lung immunity against pneumonia. Nat Rev Gastroenterol Hepatol 14(5):263–264 55. Schuijt TJ et al (2016) The gut microbiota plays a protective role in the host defence against pneumococcal pneumonia. Gut 65(4):575–583 56. Samuelson DR et al (2017) Alcohol-associated intestinal dysbiosis impairs pulmonary host defense against Klebsiella pneumoniae. PLoS Pathog 13(6):e1006426 57. Samuelson DR et al (2016) Analysis of the intestinal microbial community and inferred functional capacities during the host response to Pneumocystis pneumonia. Exp Lung Res 42 (8–10):425–439 58. Fouhy F et al (2017) A pilot study demonstrating the altered gut microbiota functionality in stable adults with Cystic Fibrosis. Sci Rep 7(1):6685 59. Garg M, Ooi CY (2017) The enigmatic gut in cystic fibrosis: linking inflammation, dysbiosis, and the increased risk of malignancy. Curr Gastroenterol Rep 19(2):6 60. Dorsey J, Gonska T (2017) Bacterial overgrowth, dysbiosis, inflammation, and dysmotility in the Cystic Fibrosis intestine. J Cyst Fibros 16(Suppl 2):S14–S23 61. Gui QF et al (2015) Well-balanced commensal microbiota contributes to anti-cancer response in a lung cancer mouse model. Genet Mol Res 14(2):5642–5651 62. Garcia-Castillo V et al (2016) Microbiota dysbiosis: a new piece in the understanding of the carcinogenesis puzzle. J Med Microbiol 65(12):1347–1362 63. Bingula R et al (2018) Characterisation of gut, lung, and upper airways microbiota in patients with non-small cell lung carcinoma: Study protocol for case-control observational trial. Medicine (Baltimore) 97(50):e13676 64. Durack J et al (2018) Delayed gut microbiota development in high-risk for asthma infants is temporarily modifiable by Lactobacillus supplementation. Nat Commun 9(1):707 65. Cabana MD et al (2017) Early probiotic supplementation for eczema and asthma prevention: a randomized controlled trial. Pediatrics 140(3):e20163000 66. Huang CF, Chie WC, Wang IJ (2018) Efficacy of Lactobacillus administration in school-age children with asthma: a randomized, placebo-controlled trial. Nutrients 10(11):e1678 67. Huffnagle GB (2014) Increase in dietary fiber dampens allergic responses in the lung. Nat Med 20(2):120–121 68. Morrow LE, Kollef MH, Casale TB (2010) Probiotic prophylaxis of ventilator-associated pneumonia: a blinded, randomized, controlled trial. Am J Respir Crit Care Med 182(8):1058– 1064 69. Khailova L et al (2017) Lactobacillus rhamnosus GG treatment improves intestinal permeability and modulates inflammatory response and homeostasis of spleen and colon in experimental model of Pseudomonas aeruginosa pneumonia. Clin Nutr 36(6):1549–1557 70. Wang J et al (2013) Probiotics for preventing ventilator-associated pneumonia: a systematic review and meta-analysis of high-quality randomized controlled trials. PLoS One 8(12): e83934 71. Barraud D, Bollaert PE, Gibot S (2013) Impact of the administration of probiotics on mortality in critically ill adult patients: a meta-analysis of randomized controlled trials. Chest 143(3):646–655 72. Petrof EO et al (2012) Probiotics in the critically ill: a systematic review of the randomized trial evidence. Crit Care Med 40(12):3290–3302 73. Valecha GK et al (2017) Anti-PD-1/PD-L1 antibodies in non-small cell lung cancer: the era of immunotherapy. Expert Rev Anticancer Ther 17(1):47–59

72

J. Tan et al.

74. Chen Q et al (2018) Delivery strategies for immune checkpoint blockade. Adv Healthc Mater 7(20):e1800424 75. Vetizou M et al (2015) Anticancer immunotherapy by CTLA-4 blockade relies on the gut microbiota. Science 350(6264):1079–1084 76. Sivan A et al (2015) Commensal Bifidobacterium promotes antitumor immunity and facilitates anti-PD-L1 efficacy. Science 350(6264):1084–1089 77. Zhao S et al (2019) Antibiotics are associated with attenuated efficacy of anti-PD-1/PD-L1 therapies in Chinese patients with advanced non-small cell lung cancer. Lung Cancer 130: 10–17 78. Fessler J, Matson V, Gajewski TF (2019) Exploring the emerging role of the microbiome in cancer immunotherapy. J Immunother Cancer 7(1):108 79. Tanoue T et al (2019) A defined commensal consortium elicits CD8 T cells and anti-cancer immunity. Nature 565(7741):600–605 80. Routy B et al (2018) Gut microbiome influences efficacy of PD-1-based immunotherapy against epithelial tumors. Science 359(6371):91–97 81. Jin Y, Dong H, Xia L, Yang Y, Zhu Y, Shen Y, Zheng H, Yao C, Wang Y, Lu S (2019) The diversity of gut microbiome is associated with favorable responses to anti-programmed death 1 immunotherapy in Chinese patients with NSCLC. J Thorac Oncol 14(8):1378–1389 82. Wypych TP, Marsland BJ, Ubags NDJ (2017) The impact of diet on immunity and respiratory diseases. Ann Am Thorac Soc 14(Supplement_5):S339–S347

Chapter 6

Gut Microbiota and Neurologic Diseases and Injuries T. Tyler Patterson and Ramesh Grandhi

Abstract The brain–gut axis is a bidirectional communication pathway connecting the central nervous system (CNS) and the gastrointestinal tract via nerve transmission, hormone, immune system, and other molecular signals. The bacterial flora of the human gut contributes direct and indirect signals to the CNS along the brain– gut axis. Alterations in gut flora, a state known as dysbiosis, has been tied to systemic inflammation, increased bacterial translocation, and increased absorbance of microbial by-products. An increase in recent literature has highlighted the role of the gut–brain axis in CNS pathology. This chapter reviews the association between gut flora dysbiosis and disorders of the central nervous system including autoimmune disease, developmental disorders, physiologic response to traumatic injury, and neurodegenerative disease.







Keywords Gut microbiota Autism Parkinson’s disease Alzheimer’s disease Multiple sclerosis Anxiety Depression Traumatic brain injury Stroke



6.1









Introduction

The brain–gut axis is a bidirectional pathway of communication between the central nervous system and the gastrointestinal tract. Signals along this path include mechanisms involving neurons, hormones, the immune system, and other neuro-active peptides [1, 2]. In particular, the microbial constitution of the gut appears to be a major factor affecting this axis—significantly enough to popularize the notion of a gut–brain–microbiota axis. T. Tyler Patterson Department of Neurosurgery, University of Texas Health Science Center School of Medicine, San Antonio, TX, USA R. Grandhi (&) Department of Neurosurgery, University of Utah School of Medicine, Salt Lake City, Utah, USA e-mail: [email protected] © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_6

73

74

T. Tyler Patterson and R. Grandhi

The gut–brain–microbiota axis, which includes the microbes that live in the human gastrointestinal tract, has become a point of focus in the literature. A typical human gut has an estimated 100 trillion bacteria from more than 1000 species [3]. These microbes not only live in the intestines, but they also serve functional roles. The commensal bacteria are involved in gut motility, secretory responses, and nutrient processing [4]. A traditional example is that of vitamin K, which humans receive almost exclusively as by-products of intestinal bacteria [5]. In addition, the gut microbiota appear to heavily interact with the central nervous system (CNS). Germ-free murine studies have provided insight into various mechanisms by which the organisms affect the brain. Alterations including lowered neurotrophic factors like brain-derived neurotrophic factor, reduced blood–brain barrier (BBB) tight cell junction proteins like occludin, and impaired immune system response have been identified in germ-free mice, indicating that the commensal bacteria play a role in modulating development of the CNS [6]. In recent years, increasing emphasis has been placed on the role of gut microbiota in pathology of the brain. More specifically, deviation from a healthy or steady-state gut microbiota may be a source for disease states. Gut microbes have previously been implicated in inflammatory conditions such as irritable bowel syndrome and ulcerative colitis, in which there exists an overgrowth in clostridia, bacteroides, and enterobacteriaceae [5, 7]. This departure from the steady-state microbiota composition is termed dysbiosis, and is increasingly cited as contributing to pathology. A number of conditions have similarly been linked to alterations in the gut–brain–microbiome axis, including the neurodevelopmental disorder of autism; neurodegenerative disorders like Parkinson’s disease (PD) and Alzheimer’s disease (AD); the autoimmune-mediated multiple sclerosis (MS); neuropsychiatric diseases such as anxiety, depression, and schizophrenia; and secondary CNS injury following traumatic brain injury (TBI), spinal cord injury (SCI), and stroke. This chapter will examine the current evidence for mechanisms underlying the gut–brain–microbiota axis in CNS pathology.

6.2

Neurodevelopmental Disorder—Autism Spectrum Disorder

Autism spectrum disorder (ASD) is a group of neurodevelopmental disorders that are regularly characterized by repetitive behaviors, impaired social interaction, and deficits in communication. Its prevalence worldwide was estimated at 52 million in 2010, with an increase expected [8]. As the leading cause of disability in patients less than 5 years of age, ASD commands serious clinical consideration. Identifying the cause and possible therapies of this disorder not only has the potential to impact millions of lives currently, but it would also provide avenues to reduce its future incidence.

6 Gut Microbiota and Neurologic Diseases and Injuries

75

The etiology of autism has become a frequent topic in the literature in recent decades. It is largely accepted that ASD is caused by a combination of genetic and environmental factors, as identified polymorphisms and de novo mutations are estimated to account for about 50% of the disease [9, 10]. A wide variety of environmental factors have been and continue to be explored. Favored hypotheses center around abnormal immune responses and neural development in utero and postnatally: viral infections, melatonin deficiencies, teratogen exposure, malnutrition, and maternal diabetes, among others [11]. More recently, the microbiota of the gut has come into focus as a source of ASD pathology. It has been noted that patients with ASD often report gastrointestinal symptoms. Even among siblings, those with ASD have been found to have more frequent symptoms such as constipation and pain compared with those without ASD [12]. These findings, in turn, potentially underscore a relationship between gut dysbiosis and autism, with several studies demonstrating various strains of Clostridium species to be over-represented in the fecal samples of patients with ASD compared with controls [13–15]. ASD patients with altered gut microbiota also demonstrate increased gut permeability when compared with controls [16], a phenomenon known as the leaky-gut state. In this state, an increased level of communication exists between the contents of the gut and the host body. This is particularly important when considering that the dysbiosis might be producing an unfavorable gut environment by disturbing the typical flora. For instance, short fatty acid chains and ammonia levels are increased in fecal samples of ASD patients [17]. This could represent poorer absorption of these bacterial products by the gut, but more likely indicates increased production of short chain fatty acids (SCFAs), ammonia, and other products by a dysbiotic gut flora. In that case, ASD patients with altered flora might be at risk of absorbing higher levels of bacterial by-products. This effect is probably superimposed with the leaky-gut phenomenon, predisposing this population to an elevated level of interaction between the gut and body—including the CNS. The increased permeability in the gastrointestinal mucosa also may contribute to systemic inflammation: lipopolysaccharide (LPS), an endotoxin which is found in many dysbiotic gut bacteria, has been shown to be elevated in the serum of autism patients, [18] resulting in the initiation of pro-inflammatory cascades. In addition, cytokines and interleukins produced locally in the gut can also circulate systemically. Interestingly, alterations in the BBB have also been detected in autism patients—selectin and claudin proteins, which are involved in tight junctions of brain endothelium, have been found to be decreased, for example [19, 20]. Taken together, these findings suggest a role for gut dysbiosis and the leaky-gut state in neuroinflammation, with the potential for causing altered CNS function and development. One proposed mechanism implicates SCFAs in mediating changes related to ASD. Propionic acid, one studied example of SFCAs, can cross the BBB and affect the parenchyma [21, 22]. In murine studies of propionic acid injected into the brain, changes associated with ASD were seen: neurochemical changes, neuroinflammation, oxidative stress, glutathione depletion, and altered phospholipid profiles.

76

T. Tyler Patterson and R. Grandhi

Further avenues exist linking the gut to the CNS. The study of germ-free mice has demonstrated changes in the brain including 5-HT receptors, neurotrophic factors like brain-derived neurotrophic factor, and N-methyl-D-aspartate receptors [6]. Although their mechanisms have not been fully outlined, these changes may be thought of as consequences of the pro-inflammatory state caused by the leaky gut and associated BBB permeability, circulating toxin, and inflammatory cascades. In addition, gut microbiota can produce a number of neurologically active compounds including histamine, gamma-aminobutyric acid, and dopamine [23]. These molecules may act locally within the enteric nervous system of the gut and indirectly affect the CNS via the feedback loop provided by the vagus nerve. Furthermore, these compounds can enter the circulation, cross the porous (BBB), and directly influence CNS function. Taken together, the clinical observation of gut symptomatology in ASD patients may be traced to biophysical changes peripherally and centrally which point to CNS dysfunction. Dysbiosis with leaky gut is integral in facilitating this pathway.

6.3

Neurodegenerative Disease—Parkinson’s Disease and Alzheimer’s Disease

Neurodegenerative disease refers to a variety of pathologies that ultimately cause impaired cognition, and in many cases dementia. PD, a motor disorder and neurodegenerative disease, is the second most common neurodegenerative disease in the developed world with a progressive incidence of 20 per 100,000 at 55 years of age to 120 per 100,000 at 70 years of age [24]. Clinically, PD is characterized by a resting tremor, muscle rigidity, gait instability, and bradykinesia—dementia often occurs sequentially. PD was discovered due to the loss of pigmented neurons in the substantia nigra, an area of particularly dopaminergic neurons. The cause of atrophy in this area is thought to be due to an aggregation of a-synuclein, characterizing PD into a group of synucleinopathies that also includes multiple system atrophy and Lewy body dementia. The accumulation of a-synuclein, which occurs intracellularly as Lewy body inclusions, is the consequence of misfolded proteins. It is proposed that oxidative stress and inflammation cause mitochondrial dysfunction and subsequent impaired dopamine metabolism [24]. For one reason or another, the neurons in the substantia nigra appear especially susceptible to a-synuclein aggregation [25]. The motor features readily apparent in PD are the result of these changes. Increasingly it is recognized that a number of symptoms may precede the motor symptoms and cognitive impairment stages of PD. This prodromal state of PD is thought to include rapid eye movement sleep behavior disorder, hyposmia, or anosmia, and gastrointestinal symptoms like constipation [26–28]. These findings correlate with autopsy studies that have identified incidental Lewy bodies in patients without the classic clinical presentation of PD. The Braak staging system

6 Gut Microbiota and Neurologic Diseases and Injuries

77

proposed a sequence of a-synuclein aggregation beginning in the autonomic nervous system and olfactory bulbs with progression centrally to the brainstem and neocortex [29]. Thus, the earliest signs of PD are thought to occur because of a-synuclein accumulation in various areas of the body, with progression to motor and cognitive symptoms occurring only after progression to the substantia nigra and forebrain occurs. The frequency of gastrointestinal symptoms in PD patients has caught the attention of researchers in recent years. PD patients report higher rates of constipation and gastrointestinal discomfort, and have even been found to statistically significant variances in components of the gut microbiome [30]. Certain species have been found to vary consistently in PD. One group found the species Prevotellaceae to be reduced by 77% compared with control; furthermore, the abundance of Enterobacteriaceae species was positively correlated with clinical severity of gait instability [31, 32]. Recognition of these changes has led to the integration of the gut–brain–microbiome axis into the pathophysiology of PD. Recent studies have continued to support this proposition. Microscopic evaluation of gut samples from PD patients demonstrated increased levels of inflammatory cytokines and glial markers [33]. The presence of cytokines, which included tumor necrosis factor alpha (TNF-a) and interferon gamma (IFN-ϒ), may represent the impact of dysbiosis in the gut of patients with PD. This local inflammation likely has important effects on the integrity of the wall, with increased intestinal permeability in PD patients. Samples have demonstrated increased intestinal staining for E. coli, higher levels of oxidative indicators including nitrotyrosine, and elevated serum LPS binding protein [34]. Taken together, these observations appear to indicate that the gut in PD is compromised and may lead to pathologic changes. Additionally, gut dysbiosis may also allow for increased levels of neurofilament and a-synuclein, as has been seen in 72% of colon samples from PD patients compared with none in a control group [35]. From there, environmental stressors, including toxins or bacterial products, can cause the exocytosis of a-synuclein into the extracellular space. Neighboring neurons can subsequently undergo endocytosis of the a-synuclein. Retrograde transport of the material can occur within the neuron, causing its aggregation in a more central location [28]. Alternatively, systemic inflammatory changes may be to blame as well. Serum LPS binding proteins have been shown to be altered in PD populations, suggesting systemic exposure of the dysbiotic environment. Thus, alterations in the gut microbiome are directly linked to inflammation within the gut, followed by the initiation of pathophysiologic cascades that ultimately lead to the classical microscopic intracranial findings of PD. Once again, the study of germ-free mice has significantly aided the effort to support these claims. One group used a line of a-synuclein overexpressing (ASO) mice to examine the role of the microbiome [25]. The control ASO mice had drastic PD syndromes due to a-synuclein aggregation. Germ-free ASO mice, by contrast, exhibited far less a-synuclein accumulation in the substantia nigra and caudoputamen. Postnatal recolonization of germ-free ASO mice caused a return of a-synuclein accumulation and the dysfunctional phenotype, further inciting the gut

78

T. Tyler Patterson and R. Grandhi

microbiota as propagators of the disease. Perhaps most importantly, germ-free animals caused a halt of microglia maturation in the CNS. The importance of these findings cannot be overstated, as they appear to show the vital role of the gut– brain–microbiota axis and the conversion of peripheral to CNS pathological findings. The aforementioned pathway provides a role for the gut–brain–microbiota axis in the setting of PD. Although there are many details left unexplored, the basic framework has been developed. Interestingly, many parallels exist between the pathophysiology proposed for PD and that of autism described previously. This may be cited as good evidence for the conserved role of the gut–brain–microbiota axis in certain pathologies, a concept evidenced in another neurodegenerative disease—Alzheimer’s disease. AD is a very common neurodegenerative disease and is the leading cause of dementia. Two histopathological findings largely in the neocortex and hippocampus accompany AD [36]. The first are intracellular neurofibrillary tangles of hyperphosphorylated tau protein. The second are extracellular accumulations referred to as senile plaque, which are formed from accumulated beta sheets of amyloid protein. Similar to in PD, changes in AD are thought to be secondary to oxidative stresses and mitochondrial dysfunction of neurons—resulting in the pathological changes described. Various factors have been proposed to cause AD. Familial forms exist because of overexpression of amyloid proteins [37]. Other risk factors have been suggested to be head trauma, chronic infections, and even gender [37, 38]. Despite these known changes in AD, the precise pathophysiology remains unknown. In very recent years, a novel focus of research has shifted to the role of the gut in AD, with a handful of emerging studies purport to incite the gut microbiota. A study from 2016 found that antibiotic administration in mice led to changes in the gut microbiota, which in turn altered the accumulation of amyloid plaques [39]. Additional analysis noted differences in circulating cytokines and chemokines. The implications of this study are large, as changing components of the gut directly lowered the rate of amyloid accumulation. Correlative differences in inflammatory cytokines suggest local and systemic inflammation may be at least partially responsible for mediating these differences. A germ-free murine study from 2017 built upon this model by finding that mice lacking gut microbiota had less cerebral amyloid plaques. Both studies appear to clearly pave the way for the gut–brain– microbiota axis to contribute to AD pathology. Thus, in both PD and AD, strong evidence appears to support the role of intestinal dysbiosis in inducing changes responsible for the accumulation of abnormal proteins. Both states appear to rely on oxidative stress and inflammation, likely due to altered gut constituents. Although the precise mechanisms are still being illuminated, conversion of these peripheral findings to central pathology appears quite likely.

6 Gut Microbiota and Neurologic Diseases and Injuries

6.4

79

Autoimmune Disease—Multiple Sclerosis

MS is an autoimmune inflammatory disorder affecting an estimated 2.5 million people worldwide with a male:female ratio of 1:1.5–2.5. [40, 41] MS is characterized by inflammation, demyelination, and gliosis of various areas of the CNS across different points in time, which can result in deficits of ambulation, vision, and more [41]. A number of risk factors have been proposed such as Epstein–Barr virus infection, cytomegalovirus infection, viral meningitis, and vitamin D deficiency. Family history also confers a predisposition for disease, meaning MS is likely the result of environmental and genetic factors. Although an autoimmune component is clear, the precise mechanism has not been uniformly established. Immune cells either intrinsic to or extrinsic to the CNS may be to blame for the pathophysiology of MS. A growing stock of evidence suggests that the extrinsic model, that is, peripheral immune cells infiltrating the CNS, may underlie MS. The bowel has long been implicated in autoimmune disease, ranging from inflammatory bowel disease to type I diabetes mellitus to rheumatoid arthritis [42]. A more narrow focus has begun to reveal the likely role of the gastrointestinal tract and its microbes in the pathophysiology of MS. Not least, alterations in the bacterial species of MS patients have been recently established. In one report, three groups were found to be significantly reduced in the group of MS patients—Faecalibacterium, Prevotella, and Anaerostipes species [43]. In the same study, Bifidobacterium and Clostridium species were elevated in MS patients, although not at a statistically significant level. In a study of MS patients, the largest of its kind, it was found that elevations in Archaea species—a kingdom separate from bacteria and eukaryotes—were associated with MS; on the other hand, known anti-inflammatory species from the Bacteroidetes and Firmicutes phyla were reduced in MS [44]. Taken together, these findings suggest that certain bacteria may accentuate autoimmune disease while others may be protective. The interface between the gastrointestinal microbes and the immune system has been alluded to previously but can be more precisely illuminated. Commensal bacteria of the gut both contain and produce components that can interact with the mucosa. These materials include but are not limited to short-chain fatty acids, tryptophan metabolites, and capsule components like polysaccharide A (PSA). Epithelial cells are able to interact with these materials via receptors, especially toll-like receptors (TLRs). The interaction of TLRs with gut materials can produce chemotactic signals for myeloid and lymphoid cells of the immune system [45]. Moreover, dendritic cells in the gut wall appear to play a role in directing inflammatory versus non-inflammatory responses when encountering foreign material. As a result, a mechanism for the gut microbiome directing immune cells begins to take shape. Gut bacteria have been consistently shown to induce inflammatory immune responses. It has been noted that germ-free mice have deficiencies in pro-inflammatory cell lines, especially Th17 cells [46]. Germ-free mice also have

80

T. Tyler Patterson and R. Grandhi

significant reductions in plasma cells, mucosal IgA production, and intestinal Peyer’s patches—all indicating that the microbiome is vital in stimulating the immune system [47]. The connection to MS emerges in light of the finding that immune cells from peripheral stimulus are found in the CNS. When germ-free mice were recolonized with segmented filamentous bacterium in the gut, responding Th17 cells were subsequently isolated in the CNS [48]. This demonstration highlights the crucial link between gut bacteria and inflammatory responses of the CNS. Murine models of experimental autoimmune encephalomyelitis (EAE) have shown that when recolonized with bacteria, previously germ-free mice develop demyelinating changes of the CNS that closely mimic those seen in MS, the main difference being that the causal antigen is artificially introduced in EAE by recolonization [49]. This finding has been reproduced in a number of studies, firmly suggesting that gut microbiota contribute to the induction of immune-mediated disease of the CNS [46, 48, 50]. Interestingly, germ-free mice are protected from EAE, even in a population of mice predisposed to the disease state [46]. Therefore, reintroduction of gut bacteria stimulates an intact peripheral immune system to confer changes centrally. Part of the centralization process also appears to involve the inclusion of B cells. To this point, most cellular responses discussed have been mediated by T cell responses or actions of the nonspecific innate immune system. In MS, however, B cells appear vital in provided auto-antibodies. This is modeled in experimental autoimmune encephalomyelitis by production of myelin oligodendrocyte glycoprotein, an auto-antibody toward myelin. Evidence that animals deficient in B cells —and thus the myelin oligodendrocyte glycoprotein auto-antibody—do not develop severe disease suggests that T cells alone cannot propagate the disease [46]. Instead, EAE likely represents the activation of B cells by T cell signaling, which confers the disease state. An important finding related to the PSA capsule of Bacteroides fragilis lends further support to this hypothesis. Specifically, the introduction of B. fragilis to the gut of mice destined to develop EAE resulted in a reduction in severity of the disease compared to a control group [50]. The proposed mechanism cited an increase in IL-10 producing T regulatory (Treg) cells, which have anti-inflammatory properties. The significance of these findings lies in the clear connection between gut microbiota, CNS disease, and secondary reduction in disease following colonization with another commensal organism. What can be concluded, then, is that gut microbiota likely stimulate T and B immune cells to proliferate—especially in a suspected dysbiotic state. Peripheral signals derived in the gut mucosa may be followed to the CNS via immune cells that incite damage on nervous tissue. Furthermore, different species and materials may have a protective effect against disease. These murine models offer vital insight into the likely role of the gut– brain–microbiota axis in the disease of MS. More specifically, the gut–brain–microbiota axis can mediate change by communicating through specialized changes of the immune system.

6 Gut Microbiota and Neurologic Diseases and Injuries

6.5

81

Neuropsychiatric Disorders—Anxiety and Depression

Major depression disorder is the most prevalent psychiatric disorder. Just under 20% of the population will meet criteria for depression in their lifetime [51]. What’s more, nearly three in four individuals with depression also meet criteria for other psychiatric disorders—most frequently an anxiety-related disorder [51]. While sub-diagnoses exist within both depression and anxiety disorders, this section will focus broadly on these two topics without further specification. The pathophysiology of anxiety and depression has been well explored in the literature; however, the entire process remains incompletely understood. From a quite broad perspective, it has been generally accepted that neurotransmitter systems involving serotonin and noradrenaline appear altered in anxiety and depression [52]. This is supported by the fact that therapy for both depression and anxiety often utilizes serotonin reuptake inhibitors to increase its availability. Still, the alterations in neurotransmitter levels most likely occur as the consequences of biochemical pathology from other systems. That is, rather than the root cause of anxiety and depression, the changes in neurotransmitter levels are probably the result of other processes. A very likely contributor to this process is the hypothalamic–pituitary–adrenal (HPA) axis. It was reported as early as the 1950s that cortisol, which is the major product of the HPA axis, is elevated in the blood of depressed patients [53]. Thus, a central theory for depression and anxiety rests on the hyperactivity of this system. Corticotropin-releasing factor or corticotropin-releasing hormone (CRH) is synthesized and released from the hypothalamus onto the pituitary gland to allow adrenocorticotrophic hormone to enter systemic circulation; when adrenocorticotrophic hormone reaches the adrenal glands, the main product cortisol is released into the bloodstream. The hypothalamus, which initiates this cascade, receives neuronal projections from a number of areas including the brainstem, the limbic forebrain, and the lamina terminalis [54]. The integration of the HPA axis with the gut–brain–microbiota axis is substantial because it adds a method of communication between these organ systems. Previous discussions have noted the role of inflammation, the adaptive immune system, and neurotransmission as mediators of signals. Herein, the HPA axis adds the component of neuroendocrine communication. As demonstrated previously, the gut–brain–microbiota axis clearly acts as a bidirectional communication between the gut and central nervous system. It is then reasonable to conclude that the microbiota of the gut might be able to affect the regulation of the HPA axis. Support for this notion arrives from population-based studies that have found an increase in gastrointestinal disorders such as irritable bowel syndrome among patients with affective disorders—including anxiety and depression [55]. So these disorders join the ranks among the formerly discussed PD and ASD as CNS pathology that appears to have a disproportionate increase in gastrointestinal symptoms. As a result, anxiety and depression may reasonably be associated with alterations in the gut microbiome and gut function.

82

T. Tyler Patterson and R. Grandhi

Animal models of depression have further supported this connection. Firstly, dysbiosis has been associated with chronically depressed mice, [56] with altered colonic motility as well. In addition, central expression of CRH was found to be elevated, representing empirical evidence of integration of the HPA axis into the pathophysiology of depression. The elevation of CRH would indicate hyperactivity of the HPA axis, likely resulting in elevated serum cortisol. Germ-free mice have helped to define the role of gut microbiota in the development of anxiety and depression. In the absence of any gut microbes, animals regularly exhibit lower levels of anxiety-related behavior [57, 58]. Conversely, animals with gut microbes developed behaviors consistent with anxiety and decreased motor activity, which can model depression. The microbiota therefore is vital to initiating these neuropsychiatric disorders, possibly via the HPA axis. Indeed, it appears the gut microbiota actually shape the development of the HPA axis. When previously germ-free mice are reconstituted early in life with microbes, the HPA axis was reduced. However, if reconstituted at a late stage, the HPA axis remained hyperactive [58, 59]. The distinction between early and late reconstitution of microbes highlights the role of the gut–brain–microbiota axis in calibrating the HPA axis early on. This suggests that gut constituents at a particular phase or phases of life drive the HPA axis, and subsequently, the formation of depression and anxiety disorders. More than one species have been identified in animal studies as having protective properties. When Bifidobacterium infantis, a species commonly found in the gut flora of infants, was re-introduced to germ-free animals, depression behavior was reversed [60]. Oral ingestion of Lactobacillus strains resulted in lower anxiety and depression behavior in rats, in addition to lower stress-induced serum cortisol levels [61, 62]. In a human study, patients taking a probiotic containing Lactobacillus and Bifidobacterium species lowered scores on anxiety and depression scales, which correlated with decreased serum cortisol. In examining anxiety and depression, much emphasis is placed on the HPA axis. Disruptions in gut microbiota likely alter the HPA axis. This may be through inflammatory effects, metabolites, or even direct neurotransmission. Vagotomized rats in a depression model did not benefit from recolonization when compared with non-vagotomized animals, suggesting a central role for interaction of the microbiome and the neuroendocrine HPA axis [61]. The ultimate conclusion is that gut microbiota are able to induce CNS pathology by at least altering the HPA axis. The downstream mechanisms remain less clear; however, one can infer that the signal alterations may lead to the subsequent development of pathology.

6.6

Acute Pathology—Traumatic Brain Injury and Stroke

The previous sections focused on chronic pathologic disease states, but the gut– brain–microbiota axis also appears to be highly involved in mediating injury in acute pathologies as well. Traumatic brain injury (TBI) and stroke are at the

6 Gut Microbiota and Neurologic Diseases and Injuries

83

forefront of the current literature focuses, not least because they are extremely common and burdensome conditions. An estimated 10 million cases of TBI occur worldwide annually, resulting in millions of emergency department visits and hundreds of thousands of hospitalizations [63, 64]. Acute stroke is also extremely common, with nearly a million US cases annually, and is the number one cause of chronic adult disability [65]. In both cases, TBI and stroke result in primary injury to the brain parenchyma. In this section, the ways in which secondary injury may occur will be explored within the context of the gut–brain–microbiota axis. An initial link between CNS injury in TBI and stroke is the effect of dysautonomia, or dysregulation of the autonomic nervous system. Dysautonomia is a phenomenon that manifests through a myriad of symptoms including episodes of tachycardia, hyperthermia, diaphoresis, hypertonicity, and/or decerebrate or decorticate posturing. The pathophysiology behind this process is thought to result from altered metabolism and release of key neurotransmitters and has been linked to TBI [66]. It is likely that the enteric nervous system conducts the dysautonomia to the gut as changes in sympathetic and parasympathetic input results in symptoms of bowel discomfort and chronic gastrointestinal pain [67]. It may be postulated that dysautonomia could alter the gut microbiome, causing a state of dysbiosis. Although this has not been shown in TBI models, SCI studies have demonstrated a correlation between dysautonomia and altered gut flora. Altered gut flora after SCI also produced worse neurologic outcomes compared with non-altered controls, implicating the gut–brain–microbiota axis in central neurologic recovery after trauma [68]. Other currently unknown mechanisms that alter the microbiome, such as neuroendocrine disruption or changes in vagal nerve signaling to the gut, may also be at play. Ultimately, CNS injury produced by trauma or stroke can be tied to dysbiosis and its subsequent effects. The cascade of effects caused by dysbiosis may once again start with altered gut permeability. Bacterial presence in the gut is well linked to intestinal mucosal integrity, as previously mentioned. Tight junction proteins undergo altered expression in the presence or absence of TLR activation [69]. These TLRs, a type of pattern recognition receptors, are thought to mediate the gut barrier integrity via baseline interactions with commensal bacteria. In a dysbiotic state, then, improper signaling may allow breakdown of this barrier. Evidence for this process has been found in a TBI animal model where gastrointestinal tissue demonstrated an increased presence of inflammatory markers, including TNF-a and IL-6, and upregulation of glycoproteins that function to recruit inflammatory cells [70]. In other words, TBI in rats resulted in protein expression to recruit immune cells and inflammatory signals in the gut, a possible result of altered mucosa integrity. The leaky gut may also allow toxic bacterial components, most notably LPS, to enter circulation. Recent work demonstrated that LPS mediates neuroinflammation after TBI by activating microglia in the CNS [71]. So as previously discussed, a pro-inflammatory state may confer damage to the CNS when circulating cytokines or bacterial components reach the intracranial circulation. Gut leakiness after CNS injury also exerts effects on the immune system. Animal models of stroke and TBI have supported this notion. An area of particular focus is

84

T. Tyler Patterson and R. Grandhi

the group of T cells referred to as Treg cells. Treg cells appear to primarily function as a fulcrum between pro-inflammatory and anti-inflammatory states. Treg cells are activated by interactions with normal gut bacteria, which results in several anti-inflammatory actions including secretion of IL-10 [72]. IL-10 production has been directly linked to a reduction in pro-inflammatory cell lines including Th1 and Th17 populations [69]. What’s more, IL-10 may exert neuroprotective effects centrally by suppressing microglia activation [73]. The mechanism for proliferation of Treg cells in the gut is thought to be driven largely through interactions with dendritic cells. By sampling the gut lumen contents, dendritic cells communicate the need for pro- or anti-inflammatory states. On the other hand, dysbiosis then may shift the balance of immune cells to favor an inflammatory state—which could hinder CNS healing or even cause further damage. Evidence for the role of the immune cells in secondary brain injury has been demonstrated in animal stroke models. After induced ischemic stroke, previously germ-free mice reconstituted with fecal contents of dysbiotic mice showed an increase in lesion volume when compared with control [74]. This finding strongly argues that intestinal bacteria stimulate an immune response that communicates with the CNS. Incredibly, the same stroke model used labeling to show that intestinal lymphocytes traveled to the brain. Further study has continued to support this notion and even hinted at an underlying mechanism. When post-ischemic-stroke mice are administered antibiotics, a reduction in cdT cells, a subset of effector T cells that produce the pro-inflammatory cytokine IL-17, and an increase in Treg cells were noted in the small intestine [75]. It can be inversely inferred that an imbalance in gut bacteria favors an inflammatory state created at least partially by cdT cells. In fact, mice deficient in cdT cells were found to have improved functional outcomes after SCI (Fig. 6.1) [76]. One consideration raised when considering the transposition of gut immune cells to the CNS is that of the BBB. In a healthy state, the BBB is vital for controlling vascular permeability and protecting the brain from unwanted influences including toxins or inflammatory factors. Following a TBI, the permeability of the BBB can increase up to four times its normal amount within just 6 h [67]. In the context of the gut–brain–microbiota axis, an increase in BBB permeability offers potentially crucial access from peripheral sources to the CNS. ICAM-1, which helps direct inflammatory cells, is expressed in the brain tissue of rats after TBI [77]. Upregulation of ICAM-1 would help to explain the presence of peripheral immune cells found in the CNS after injury, possibly providing a precise mechanism for secondary neuroinflammation after injury from TBI or stroke. In parallel, these findings illustrate the likely mechanism for which gut microbiota affect CNS healing after injury. Good evidence has been produced in stroke models, while TBI data is still growing or must be extrapolated from SCI models. Taken together, though, the information collected to date supports the notion that acute CNS injury has the capability to alter gut microbiota. In return, these changes appear to affect CNS healing. A leaky-gut phenomenon appears to promote inflammatory cytokine release and immune cell translocation to the brain. Thus, targets for reducing the morbidity of stroke and TBI may therefore focus on

6 Gut Microbiota and Neurologic Diseases and Injuries

85

Fig. 6.1 CD4+ Regulatory T cells (Treg) serve as a fulcrum in the balance of pro-inflammatory and anti-inflammatory immune modulation. In the gut, microbiota are sampled or may translocate, resulting in an immune response. The CD103+ dendritic cells (CD103+DC), a type of antigen presenting cell of the intestinal mucosa, have been found to drive the proliferation of Treg cells while the microbiota is in a steady state. Subsequent activity by Treg cells results in suppression of an inflammatory response by multiple mechanisms. A central focus of this illustration is the Treg cell-mediated influence on cdT cells. cdT cells have the potential to cause either inflammatory or anti-inflammatory effects, but the influence of Treg cells while gut microbiota remain in steady state urges a reduced inflammatory state. Treg cells additionally produce IL-10, further suppressing Th1 and Th17 responses. In the presence of dysbiosis, interactions with dendritic cells and Treg cells result in a loss of inhibitory effects and, thus, the development of pro-inflammatory states

reducing inflammation, perhaps by modulating the population of gut microbes. More investigation of this promising topic, however, is required.

6.7

Summary

In this chapter, the extent of the gut–brain–microbiota axis has been explored within the context of brain pathology. This bidirectional communication between the gut and the brain is increasingly linked in various ways to disease states including ASD, PD, AD, MS, TBI, and stroke. Through these examples discussed

86

T. Tyler Patterson and R. Grandhi

Fig. 6.2 This figure demonstrates the proposed impact of dysbiotic gut microbiota on the integrity of gastrointestinal epithelial integrity and the central nervous system. When sampling dysbiotic flora, the local immune response results in inflammatory mediators including interferon (INF) and tumor necrosis factor (TNF) alpha. Systemic effects of this inflammatory process include a “leaky-gut” phenomenon at least partially due to tight junction downregulation. Synergistically, other products of dysbiotic flora such as short-chain fatty acids (SCFAs), lipopolysaccharide (LPS), and neurotransmitters may translocate. The pro-inflammatory signals and bacterial materials may then reach and affect the central nervous system

above, the communication pathways of the gut–brain–microbiota axis have been highlighted. It has been shown that various mechanisms underlie this relationship (Fig. 6.2). Neurologic disease has been consistently linked to gastrointestinal symptoms, which helped to uncover the frequent state of gut dysbiosis in these patients. The effects of altered gut microbial content can be traced to several effector functions on the CNS. One of the first consequences of dysbiosis seems to be an increase in intestinal permeability as demonstrated by increased serum levels of bacterial toxins and metabolites. These toxins and metabolites potentially have the capacity to directly influence the brain, but they also point to a further process. By increasing intestinal permeability, the gut is subjected to abnormal exposure to luminal contents. Because the gastrointestinal tract houses an immense part of the peripheral immune system, dysbiosis allows more communication between bacterial products and immune cells. Through a number of proposed mechanisms, a pro-inflammatory state becomes evident from circulating cytokines and translocated immune cells. Parallel findings have been noted in various disease states, stabilizing the foundations of these propositions.

6 Gut Microbiota and Neurologic Diseases and Injuries

87

Although more investigation is required to fully illuminate the role of the gut– brain–microbiota axis, the basic mechanisms at play have begun to reveal themselves. With more bacteria in the lumen of the gut than cells in the human body, it is no wonder that the gut–brain–microbiota axis retains the potential to contribute to disease states. Furthermore, eventual therapeutic targets may lie within the details of these mechanisms.

References 1. Wang Y, Kasper LH (2014) The role of microbiome in central nervous system disorders. Brain Behav Immun 38:1–12. https://doi.org/10.1016/j.bbi.2013.12.015 2. Sundman MH, Chen N-K, Subbian V, Chou Y-H (2017) The bidirectional gut-brain-microbiota axis as a potential nexus between traumatic brain injury, inflammation, and disease. Brain Behav Immun 66:31–44. https://doi.org/10.1016/j.bbi.2017.05.009 3. Flowers SA, Ellingrod VL (2015) The microbiome in mental health: potential contribution of gut microbiota in disease and pharmacotherapy management. Pharmacotherapy. 35(10):910– 916. https://doi.org/10.1002/phar.1640 4. Rhee SH, Pothoulakis C, Mayer EA (2009) Principles and clinical implications of the brain-gut-enteric microbiota axis. Nat Rev Gastroenterol Hepatol 6(5):306–314. https://doi. org/10.1038/nrgastro.2009.35 5. Ghaisas S, Maher J, Kanthasamy A (2016) Gut microbiome in health and disease: linking the microbiome-gut-brain axis and environmental factors in the pathogenesis of systemic and neurodegenerative diseases. Pharmacol Ther 158:52–62. https://doi.org/10.1016/j.pharmthera. 2015.11.012 6. Sharon G, Sampson TR, Geschwind DH, Mazmanian SK (2016) The central nervous system and the gut microbiome. Cell 167(4):915–932. https://doi.org/10.1016/j.cell.2016.10.027 7. Grenham S, Clarke G, Cryan JF, Dinan TG (2011) Brain–gut–microbe communication in health and disease. Front Physiol 2:94. https://doi.org/10.3389/fphys.2011.00094 8. Baxter AJ, Brugha TS, Erskine HE, Scheurer RW, Vos T, Scott JG (2015) The epidemiology and global burden of autism spectrum disorders. Psychol Med 45(3):601–613. https://doi.org/ 10.1017/S003329171400172X 9. Hallmayer J, Cleveland S, Torres A et al (2011) Genetic heritability and shared environmental factors among twin pairs with autism. Arch Gen Psychiatry 68(11):1095–1102. https://doi. org/10.1001/archgenpsychiatry.2011.76 10. Vuong HE, Hsiao EY (2017) Emerging roles for the gut microbiome in autism spectrum disorder. Biol Psychiatry 81(5):411–423. https://doi.org/10.1016/j.biopsych.2016.08.024 11. Grabrucker AM (2013) Environmental factors in autism. Front Psychiatry 3:118. https://doi. org/10.3389/fpsyt.2012.00118 12. Wang LW, Tancredi DJ, Thomas DW (2011) The prevalence of gastrointestinal problems in children across the United States with autism spectrum disorders from families with multiple affected members. J Dev Behav Pediatr 32(5):351–360. https://doi.org/10.1097/DBP. 0b013e31821bd06a 13. Song Y, Liu C, Finegold SM (2004) Real-time PCR quantitation of clostridia in feces of autistic children. Appl Environ Microbiol 70(11):6459–6465. https://doi.org/10.1128/AEM. 70.11.6459-6465.2004 14. Parracho HM, Bingham MO, Gibson GR, McCartney AL (2005) Differences between the gut microflora of children with autistic spectrum disorders and that of healthy children. J Med Microbiol 54(10):987–991. https://doi.org/10.1099/jmm.0.46101-0

88

T. Tyler Patterson and R. Grandhi

15. Mulle JG, Sharp WG, Cubells JF (2013) The gut microbiome: a new frontier in autism research. Curr Psychiatry Rep 15(2):337. https://doi.org/10.1007/s11920-012-0337-0 16. de Magistris L, Familiari V, Pascotto A et al (2010) Alterations of the intestinal barrier in patients with autism spectrum disorders and in their first-degree relatives. J Pediatr Gastroenterol Nutr 51(4):418–424. https://doi.org/10.1097/MPG.0b013e3181dcc4a5 17. Wang L, Christophersen CT, Sorich MJ, Gerber JP, Angley MT, Conlon MA (2012) Elevated fecal short chain fatty acid and ammonia concentrations in children with autism spectrum disorder. Dig Dis Sci 57(8):2096–2102. https://doi.org/10.1007/s10620-012-2167-7 18. Emanuele E, Orsi P, Boso M et al (2010) Low-grade endotoxemia in patients with severe autism. Neurosci Lett 471(3):162–165. https://doi.org/10.1016/j.neulet.2010.01.033 19. Onore CE, Nordahl CW, Young GS, Van de Water JA, Rogers SJ, Ashwood P (2012) Levels of soluble platelet endothelial cell adhesion molecule-1 and P-selectin are decreased in children with autism spectrum disorder. Biol Psychiatry 72(12):1020–1025. https://doi.org/10. 1016/j.biopsych.2012.05.004 20. Fiorentino M, Sapone A, Senger S et al (2016) Blood–brain barrier and intestinal epithelial barrier alterations in autism spectrum disorders. Mol Autism 7(1):49. https://doi.org/10.1186/ s13229-016-0110-z 21. MacFabe DF, Cain NE, Boon F, Ossenkopp K-P, Cain DP (2011) Effects of the enteric bacterial metabolic product propionic acid on object-directed behavior, social behavior, cognition, and neuroinflammation in adolescent rats: Relevance to autism spectrum disorder. Behav Brain Res 217(1):47–54. https://doi.org/10.1016/j.bbr.2010.10.005 22. MacFabe DF (2012) Short-chain fatty acid fermentation products of the gut microbiome: implications in autism spectrum disorders. Microb Ecol Health Dis 23(1):19260. https://doi. org/10.3402/mehd.v23i0.19260 23. Li Q, Han Y, Dy ABC, Hagerman RJ (2017) The gut microbiota and autism spectrum disorders. Front Cell Neurosci 11:120. https://doi.org/10.3389/fncel.2017.00120 24. Dauer W, Przedborski S (2003) Parkinson’s disease: mechanisms and models. Neuron 39 (6):889–909. https://doi.org/10.1016/S0896-6273(03)00568-3 25. Sampson TR, Debelius JW, Thron T et al (2016) Gut microbiota regulate motor deficits and neuroinflammation in a model of Parkinson’s disease. Cell 167(6):1469–1480.e12. https://doi. org/10.1016/j.cell.2016.11.018 26. Adler CH, Beach TG (2016) Neuropathological basis of non-motor manifestations of Parkinson’s disease. Mov Disord 31(8):1114–1119. https://doi.org/10.1002/mds.26605 27. Abbott RD, Ross GW, Petrovitch H et al (2007) Bowel movement frequency in late-life and incidental Lewy bodies. Mov Disord 22(11):1581–1586. https://doi.org/10.1002/mds.21560 28. Klingelhoefer L, Reichmann H (2015) Pathogenesis of Parkinson disease–the gut-brain axis and environmental factors. Nat Rev Neurol 11(11):625–636. https://doi.org/10.1038/nrneurol. 2015.197 29. Braak H, Tredici KD, Rüb U, de Vos RAI, Jansen Steur ENH, Braak E (2003) Staging of brain pathology related to sporadic Parkinson’s disease. Neurobiol Aging 24(2):197–211. https://doi.org/10.1016/S0197-4580(02)00065-9 30. Hill-Burns EM, Debelius JW, Morton JT et al (2017) Parkinson’s disease and PD medications have distinct signatures of the gut microbiome. Mov Disord 32(5):739–749. https://doi.org/ 10.1002/mds.26942 31. Scheperjans F, Aho V, Pereira PAB et al (2015) Gut microbiota are related to Parkinson’s disease and clinical phenotype. Mov Disord 30(3):350–358. https://doi.org/10.1002/mds. 26069 32. Unger MM, Spiegel J, Dillmann K-U et al (2016) Short chain fatty acids and gut microbiota differ between patients with Parkinson’s disease and age-matched controls. Parkinsonism Relat Disord 32:66–72. https://doi.org/10.1016/j.parkreldis.2016.08.019 33. Devos D, Lebouvier T, Lardeux B et al (2013) Colonic inflammation in Parkinson’s disease. Neurobiol Dis 50:42–48. https://doi.org/10.1016/j.nbd.2012.09.007

6 Gut Microbiota and Neurologic Diseases and Injuries

89

34. Forsyth CB, Shannon KM, Kordower JH et al (2011) Increased intestinal permeability correlates with sigmoid mucosa alpha-synuclein staining and endotoxin exposure markers in early Parkinson’s disease. PLoS One 6(12):e28032. https://doi.org/10.1371/journal.pone. 0028032 35. Lebouvier T, Neunlist M, Bruley des Varannes S, et al (2010) Colonic biopsies to assess the neuropathology of Parkinson’s disease and its relationship with symptoms. PLoS One 5(9): e12728. https://doi.org/10.1371/journal.pone.0012728 36. Moreira PI, Carvalho C, Zhu X, Smith MA, Perry G (2010) Mitochondrial dysfunction is a trigger of Alzheimer’s disease pathophysiology. Biochim Biophys Acta 1802(1):2–10. https:// doi.org/10.1016/j.bbadis.2009.10.006 37. Cummings JL, Vinters HV, Cole GM, Khachaturian ZS (1998) Alzheimer’s disease: etiologies, pathophysiology, cognitive reserve, and treatment opportunities. Neurology 51(1 Suppl 1):S2–S17; discussion S65–S67 38. Vogt NM, Kerby RL, Dill-McFarland KA et al (2017) Gut microbiome alterations in Alzheimer’s disease. Sci Rep 7(1):13537. https://doi.org/10.1038/s41598-017-13601-y 39. Minter MR, Zhang C, Leone V et al (2016) Antibiotic-induced perturbations in gut microbial diversity influences neuro-inflammation and amyloidosis in a murine model of Alzheimer’s disease. Sci Rep 6:30028. https://doi.org/10.1038/srep30028 40. Ascherio A, Munger K (2008) Epidemiology of multiple sclerosis: from risk factors to prevention. Semin Neurol 28(1):17–28. https://doi.org/10.1055/s-2007-1019126 41. Dendrou CA, Fugger L, Friese MA (2015) Immunopathology of multiple sclerosis. Nat Rev Immunol 15(9):545–558. https://doi.org/10.1038/nri3871 42. Jangi S, Gandhi R, Cox LM et al (2016) Alterations of the human gut microbiome in multiple sclerosis. Nat Commun 7:12015. https://doi.org/10.1038/ncomms12015 43. Miyake S, Kim S, Suda W et al (2015) Dysbiosis in the gut microbiota of patients with multiple sclerosis, with a striking depletion of species belonging to clostridia XIVa and IV clusters. PLoS One 10(9):e0137429. https://doi.org/10.1371/journal.pone.0137429 44. Jhangi S, Gandhi R, Glanz B, et al (2014) Increased Archaea species and changes with therapy in gut microbiome of multiple sclerosis subjects (S24.001). Neurology 82(10 Supplement):S24.001 45. McDermott AJ, Huffnagle GB (2014) The microbiome and regulation of mucosal immunity. Immunology 142(1):24–31. https://doi.org/10.1111/imm.12231 46. Berer K, Mues M, Koutrolos M et al (2011) Commensal microbiota and myelin autoantigen cooperate to trigger autoimmune demyelination. Nature 479(7374):538–541. https://doi.org/ 10.1038/nature10554 47. Mielcarz DW, Kasper LH (2015) The gut microbiome in multiple sclerosis. Curr Treat Options Neurol 17(4):18. https://doi.org/10.1007/s11940-015-0344-7 48. Lee YK, Menezes JS, Umesaki Y, Mazmanian SK (2011) Proinflammatory T-cell responses to gut microbiota promote experimental autoimmune encephalomyelitis. Proc Natl Acad Sci U S A 108(Suppl 1):4615–4622. https://doi.org/10.1073/pnas.1000082107 49. Constantinescu CS, Farooqi N, O’Brien K, Gran B (2011) Experimental autoimmune encephalomyelitis (EAE) as a model for multiple sclerosis (MS). Br J Pharmacol 164 (4):1079–1106. https://doi.org/10.1111/j.1476-5381.2011.01302.x 50. Ochoa-Repáraz J, Mielcarz DW, Ditrio LE, et al (2009) Role of gut commensal microflora in the development of experimental autoimmune encephalomyelitis. J Immunol 183(10):6041– 6050. https://doi.org/10.4049/jimmunol.0900747 51. Adams GC, Balbuena L, Meng X, Asmundson GJG (2016) When social anxiety and depression go together: a population study of comorbidity and associated consequences. J Affect Disord 206:48–54. https://doi.org/10.1016/j.jad.2016.07.031 52. Ressler KJ, Nemeroff CB (2000) Role of serotonergic and noradrenergic systems in the pathophysiology of depression and anxiety disorders. Depress Anxiety 12(S1):2–19. https:// doi.org/10.1002/1520-6394(2000)12:1+%3c2::aid-da2%3e3.0.co;2-4 53. Board F, Persky H, Hamburg DA (1956) Psychological stress and endocrine functions. Psychosom Med 18(4):324–333

90

T. Tyler Patterson and R. Grandhi

54. Smith SM, Vale WW (2006) The role of the hypothalamic-pituitary-adrenal axis in neuroendocrine responses to stress. Dialogues Clin Neurosci 8(4):383 55. Mayer EA, Craske M, Naliboff BD (2001) Depression, anxiety, and the gastrointestinal system. J Clin Psychiatry 62(Suppl 8):28–36; discussion 37 56. Park AJ, Collins J, Blennerhassett PA, et al (2013) Altered colonic function and microbiota profile in a mouse model of chronic depression. Neurogastroenterol Motil 25(9):733-e575. https://doi.org/10.1111/nmo.12153 57. Diaz Heijtz R, Wang S, Anuar F et al (2011) Normal gut microbiota modulates brain development and behavior. Proc Natl Acad Sci U S A 108(7):3047–3052. https://doi.org/10. 1073/pnas.1010529108 58. Neufeld KAM, Kang N, Bienenstock J, Foster JA (2011) Effects of intestinal microbiota on anxiety-like behavior. Commun Integr Biol 4(4):492–494. https://doi.org/10.4161/cib.4.4. 15702 59. Sudo N, Chida Y, Aiba Y et al (2004) Postnatal microbial colonization programs the hypothalamic–pituitary–adrenal system for stress response in mice. J Physiol 558(1):263–275. https://doi.org/10.1113/jphysiol.2004.063388 60. Desbonnet L, Garrett L, Clarke G, Kiely B, Cryan JF, Dinan TG (2010) Effects of the probiotic Bifidobacterium infantis in the maternal separation model of depression. Neuroscience 170(4):1179–1188. https://doi.org/10.1016/j.neuroscience.2010.08.005 61. Bravo JA, Forsythe P, Chew MV et al (2011) Ingestion of Lactobacillus strain regulates emotional behavior and central GABA receptor expression in a mouse via the vagus nerve. Proc Natl Acad Sci U S A 108(38):16050–16055. https://doi.org/10.1073/pnas.1102999108 62. Ait-Belgnaoui A, Durand H, Cartier C et al (2012) Prevention of gut leakiness by a probiotic treatment leads to attenuated HPA response to an acute psychological stress in rats. Psychoneuroendocrinology 37(11):1885–1895. https://doi.org/10.1016/j.psyneuen.2012.03. 024 63. Faul M, Xu L, Wald MM, Coronado VG (2010) Traumatic brain injury in the United States: emergency department visits, hospitalizations, and deaths 2002–2006. Atlanta: Centers for Disease Control and Prevention, National Center for Injury Prevention and Control 64. Hyder AA, Wunderlich CA, Puvanachandra P, Gururaj G, Kobusingye OC (2007) The impact of traumatic brain injuries: a global perspective. NeuroRehabilitation 22(5):341–353 65. Ovbiagele B, Nguyen-Huynh MN (2011) Stroke epidemiology: advancing our understanding of disease mechanism and therapy. Neurotherapeutics 8(3):319–329. https://doi.org/10.1007/ s13311-011-0053-1 66. Toklu HZ, Sakarya Y, Tümer N (2017) A proteomic evaluation of sympathetic activity biomarkers of the hypothalamus-pituitary-adrenal axis by western blotting technique following experimental traumatic brain injury. Methods Mol Biol 1598:313–325. https:// doi.org/10.1007/978-1-4939-6952-4_16 67. Kharrazian D (2015) Traumatic brain injury and the effect on the brain-gut axis. Altern Ther Health Med 21(Suppl 3):28–32 68. Kigerl KA, Hall JCE, Wang L, Mo X, Yu Z, Popovich PG (2016) Gut dysbiosis impairs recovery after spinal cord injury. J Exp Med 213(12):2603–2620. https://doi.org/10.1084/jem. 20151345 69. Nishio J, Honda K (2012) Immunoregulation by the gut microbiota. Cell Mol Life Sci 69 (21):3635–3650. https://doi.org/10.1007/s00018-012-0993-6 70. Hang C-H, Shi J-X, Li J-S, Li W-Q, Yin H-X (2005) Up-regulation of intestinal nuclear factor kappa B and intercellular adhesion molecule-1 following traumatic brain injury in rats. World J Gastroenterol 11(8):1149–1154. https://doi.org/10.3748/wjg.v11.i8.1149 71. Wen L, You W, Wang H, Meng Y, Feng J, Yang X (2018) Polarization of microglia to the M2 phenotype in a peroxisome proliferator-activated receptor gamma-dependent manner attenuates axonal injury induced by traumatic brain injury in mice. J Neurotrauma 35 (19):2330–2340. https://doi.org/10.1089/neu.2017.5540

6 Gut Microbiota and Neurologic Diseases and Injuries

91

72. Round JL, Mazmanian SK (2009) The gut microbiota shapes intestinal immune responses during health and disease. Nat Rev Immunol 9(5):313–323. https://doi.org/10.1038/nri2515 73. Kremlev SG, Palmer C (2005) Interleukin-10 inhibits endotoxin-induced pro-inflammatory cytokines in microglial cell cultures. J Neuroimmunol 162(1–2):71–80. https://doi.org/10. 1016/j.jneuroim.2005.01.010 74. Singh V, Roth S, Llovera G et al (2016) Microbiota dysbiosis controls the neuroinflammatory response after stroke. J Neurosci 36(28):7428–7440. https://doi.org/10.1523/JNEUROSCI. 1114-16.2016 75. Benakis C, Brea D, Caballero S et al (2016) Commensal microbiota affects ischemic stroke outcome by regulating intestinal cd T cells. Nat Med 22(5):516–523. https://doi.org/10.1038/ nm.4068 76. Sun G, Yang S, Cao G et al (2018) cd T cells provide the early source of IFN-c to aggravate lesions in spinal cord injury. J Exp Med 215(2):521–535. https://doi.org/10.1084/jem. 20170686 77. Chen G, Shi J, Jin W et al (2008) Progesterone administration modulates TLRs/NF-kappaB signaling pathway in rat brain after cortical contusion. Ann Clin Lab Sci 38(1):65–74

Chapter 7

Gut Microbiota and Renal Injury Lei Zhang, Wen Zhang, and Jing Nie

Abstract Renal injury, especially chronic kidney disease (CKD), is closely associated with gut microbiota. It is well known that renal injury development could cause enteric microbial compositional disruption. On the other hand, gut microbial composition, as well as their function, would directly influence the renal disease progression. Here, in the present chapter, we will summarize the crosstalk between intestinal microbiota and renal disease and discuss some potential therapeutic approaches based on this topic. Keywords Gut microbiota Metabolites

7.1

 Chronic kidney disease  Short-chain fatty acids 

Introduction

Chronic kidney disease (CKD) is a worldwide public health problem leading to poor outcomes and high financial burden. It is estimated that over 119 million Chinese people have CKD and further develop into end-stage renal failure (ESRD), a devastating condition that requires life-long dialysis or kidney transplantation [1]. Numerous epidemiological studies have indicated that the prevalence of CKD is increasing worldwide [2, 3]. As the population of patients with diabetes mellitus and obesity continues to grow, it is highly probable that this trend of increasing prevalence of CKD will only continue to worsen in the foreseeable future. In addition to decreased renal function, cardiovascular disease (CVD) is the most frequent cause of death in CKD patients. Despite the enormity of this problem, current therapeutic options for CKD in the clinical setting are scarce. New therapeutic interventions to delay the progression of CKD are critically needed. L. Zhang  W. Zhang  J. Nie (&) State Key Laboratory of Organ Failure Research, National Clinical Research Center of Kidney Disease, Key Laboratory of Organ Failure Research (Ministry of Education), Division of Nephrology, Nanfang Hospital, Southern Medical University, No. 1838 Guangzhou Ave., Guangzhou 510515, China e-mail: [email protected] © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_7

93

94

L. Zhang et al.

In the past decade, accumulating studies have shown that, compared with healthy people, the composition of intestinal flora in CKD patients has significantly changed [2, 3]. CKD-related dysbiosis and changes in the intestinal barrier encourage the increased translocation of living bacteria or bacterial products from the intestinal lumen into circulation, a process that could account for the persistent systemic inflammation in CKD/ESRD. Despite technical innovations, systemic inflammation caused by bacterial migration remains a serious problem to be solved. Therefore, studies have increasingly focused on the mechanisms of the kidney– intestinal axis, aiming to delay the progression of CKD and prevent ESRD complications through regulation of gut microbiota [4–8]. This chapter focuses on the potential roles of intestinal microbiota in the context of chronic kidney disease.

7.2

Impact of CKD/ESRD on the Intestinal Microbiota

In CKD patients, the homeostasis of the original intestinal flora is broken, which manifests as a decrease in probiotics and increase in toxigenic flora. Changes in the gut microbiome are thought to be different in different stages of CKD. Vaziri et al. [9] reported that the proportions of Cryptotermes brevis, Enterobacteriaceae, Halomonas, Moraxella, Polycystiaceae, Pseudomonas and Thiothrix were significantly increased, while those of Lactobacillaceae and Prevottiaceae were significantly decreased in stool samples of ESRD patients compared with healthy controls. In patients receiving hemodialysis, the number of intestinal anaerobic bacteria, especially enterobacteria and enterococci, is 100 times higher than in healthy people [10, 11]. The ratio of cellulose to protein is decreased in the intestinal tract of CKD patients, resulting in an increase in the proportion of Clostridium difficile and Bacteroides bacteria, which tend to produce toxic metabolites such as ammonia, amines, thiols and phenol, the main components of uremia toxins [8]. The intestinal flora imbalance in CKD patients is mainly due to the following reasons: (1) CKD patients usually have body fluid retention and urea accumulation; the microzymes of the intestinal flora hydrolyze the urea to produce ammonia, resulting in the uremic enterocolitis [12, 13]. (2) The intestinal epitheliums of CKD patients actively secrete a large amount of uric acid and oxalic acid into the intestinal cavity, leading to changes in the energy metabolism of the flora [14]. (3) Due to restrictions on potassium ion intake, CKD patients are limited in the amount of fruit and fresh vegetables they can consume, resulting in decreased cellulose intake. (4) In addition, the use of antibiotics and phosphorus-binding agents impairs the structure of intestinal flora.

7 Gut Microbiota and Renal Injury

7.3

95

Impairment of the Gut Barrier in CKD

The intestinal epithelium is a major component of the gut barrier for separating microbial products in the gut lumen from the host circulatory system. The tight junction proteins between epithelial cells can prevent non-selective transport of substances across the barrier [15]. Intestinal Bifidobacteria promote the expression of tight junction proteins such as ZO-1, claudins, occludin and E-cadherin, to maintain the integrity of intestinal barrier [16]. In CKD patients, however, an imbalance in intestinal flora causes an increased proportion of pathogens and disruption of the epithelial layer ecosystem, which lead to increased intestinal permeability and impaired nutrient absorption [17]. Meanwhile, uremic toxins such as urea and ammonia are capable of directly damaging the gut barrier. In addition, intestinal wall edema, ischemia, electrolyte disorder and hypoalbuminemia can also lead to intestinal barrier dysfunction in CKD patients. Vaziri et al. demonstrated chronic inflammation extending throughout the gastrointestinal tract in hemodialysis patients [18]. Subsequent studies have found that CKD rats showed systemic oxidative stress and marked depletion of epithelial tight junction proteins such as claudin-1, occludin, and ZO-1 [19]. Systemic inflammation marked by activation of circulating leukocytes and elevation of plasma proinflammatory cytokines persists in CKD and has been suggested as a major catalyst for the progression of CKD as well as CVD in CKD [20, 21]. The major causes of systemic inflammation in CKD/ESRD patients include direct induction by uremic toxins, oxidative stress, catheter-related infections and biocompatibility between dialysis solution and membranes. Another cause not to be ignored is the translocation of living bacteria and endotoxins into systemic circulation in ESRD patients due to gut barrier impairment [22, 23]. Bacterial endotoxins are a powerful stimulus of inflammation activation via the Toll-like receptor 4 (TLR4) and NOD-like receptor protein 3 (NLRP3) inflammasome pathways. Endotoxins may also activate pattern recognition receptors (PRRs) in immune cells, triggering an inflammation cascade. Therefore, impairment of the intestinal barrier permits exposure of the mucosal immune system to pathogens, toxins and allergens, which leads to activation of inflammation pathways and release of proinflammatory cytokines. Persistent chronic inflammation and inflammatory cytokines could promote epithelial transition of tubular epithelial cells and fibroblast activation, and promote the progression of renal failure. Circulatory LPS can also activate vascular endothelial cells to release soluble TNF receptors and promote atherosclerosis [22– 24]. Continued inflammatory activation does not imply a stronger immunity to pathogens. In contrast, sustained immune activation induces immune tolerance, leading to a reduced immune response to pathogens and increased risk of severe infection in CKD patients. Therefore, the persistent chronic inflammatory state caused by intestinal barrier injury and intestinal flora migration is a factor for accelerated development of CKD; this calls for new therapeutic methods.

96

7.4

L. Zhang et al.

SCFA, a Beneficial Product of Intestinal Flora

Short-chain fatty acids (SCFAs) are important products of intestinal flora and exert anti-inflammatory effects. The most abundant SCFAs are acetate, propionate and butyrate, produced by enteric bacterially mediated catabolism of dietary fiber. SCFAs are the preferred energy source for the colonic epithelium [25]. In addition, butyrate has been shown to promote reassembly of intestinal tight junction proteins and improve gut barrier function [26]. However, the abundance of SCFA-producing bacteria as well as serum levels of SCFA are significantly reduced in ESRD patients, and there is an inverse correlation between butyrate level and renal function [27]. It has been demonstrated that SCFA can exert anti-inflammatory effects through a variety of pathways. So far, researchers have suggested that SCFAs may operate in two manners: by binding G-protein membrane receptors (GPR41 and GPR43) to decrease cAMP production by inhibiting adenylate cyclase pathway, or by entering cells directly through transporter channels in the cellular membrane [28, 29] and working as histone deacetylase (HDAC) inhibitors to modulate epigenetic processes [30]. In acute renal injury, SCFA supplementation reduces cellular oxidative stress and promotes autophagy, thereby attenuating renal injury [31]. Gonzalez et al. reported that butyrate treatment increases AMPK phosphorylation in the intestines of 5/6th nephrectomy rats and promotes colonic mucin and tight junction protein expression to decrease LPS leakage and ameliorate inflammation [26]. In addition, SCFAs can act as a non-competitive inhibitor of histone deacetylases (HDACs) to modulate gene expression in immune cells and achieve immune tolerance [32]. Butyrate has been shown to inhibit histone deacetylases (HDACs) [33] and up-regulate FOXP3, which further promotes induction of peripheral Treg cells and reduction of inflammation [34]. Given the critical role of local and systemic inflammation in deterioration of CKD and arteriosclerosis, the anti-inflammatory effects of SCFAs are expected to delay the progression of CKD and reduce the risk of cardiovascular diseases in CKD patients.

7.5

Gut-Derived Uremic Toxins

CKD, especially ESRD, is characterized by the progressive accumulation of uremic toxins due to a decline in renal excretion. The precursors of uremic toxins are formed in the GI tract during protein fermentation by the microbiota [21]. Most uremic toxins are sulfidized or glycosylated and bound to albumin [35]. In a normally functioning kidney, the protein-bound solutes are removed through the active process of tubular secretion through the organic anion transport system. However, in patients receiving maintenance dialysis, dialysis does not adequately remove uremia toxins. As a result, the levels of these compounds in dialysis patients are extremely higher than those in normal people [36, 37].

7 Gut Microbiota and Renal Injury

97

Among protein-bound solutes, indoxyl sulfate (IS) and p-cresyl sulfate (PCS) are the most well studied (Fig. 7.1). Indole is generated by bacterial metabolism of tryptophan in the colon, which is further oxidized and sulfurized into IS in the liver. Gut flora can also convert tyrosine into 4-hydroxyphenylacetic acid, which is turned into p-methylphenol, mostly by decarboxylase of C. difficile; finally, p-methylphenol is converted into PCS [36, 37]. Serum levels of IS and PCS are reported to increase parallel to GFR declination. Their levels are in positive correlation with all-cause mortality and increased risk for cardiovascular events in CKD patients [38–40]. Mechanistically, IS can induce pro-inflammatory responses in endothelial cells through inducing oxidative stress and inhibiting their self-repairing processes [41, 42]. In vitro experiments showed that IS could induce NADPH oxidase and activate the MAPK/NF-jB pathway in vascular smooth muscle cells [43]. It has been reported that PCS stimulated endothelial microparticle release, a marker of endothelial damage. PCS could magnify oxidative stress in human vascular smooth muscle cells (HVSMCs) as well as in cardiomyocytes by up-regulating NADPH oxidase [44, 45]. In addition to vascular inflammation, IS and PCS are also involved in renal fibrosis processes. IS accelerates glomerular sclerosis and renal interstitial fibrosis by activating the RAS system and TGF-b/Smad3 pathway [46, 47]. PCS aggravates renal interstitial fibrosis by increasing inflammatory cytokine production and inducing epithelial to mesenchymal transition (EMT) of renal tubular epithelial cells. Watanabe et al. demonstrated that PCS promotes the production of reactive oxygen species (ROS) in renal tubular cells by enhancing NADPH oxidase activity, which further up-regulates the production of inflammatory cytokines and TGF-b1 [48]. Among the gut microbiota-produced metabolites, trimethylamine-N-oxide (TMAO) has entered the spotlight recently because of its tight connection with CVD [49–51]. A diet enriched in red meat, egg yolk, cheese or seafood contains high choline, lecithin, betaine and carnitine content. They can be metabolized by intestinal microbiota to produce trimethylamine (TMA), which in turn is oxidized to

Fig. 7.1 Crosstalk between kidney and gut microbiota

98

L. Zhang et al.

TMAO by hepatic flavin-containing monooxygenases (FMOs) [52]. TMAO is predominantly excreted by the kidney, and its serum level therefore increases as eGFR declines. Concentrations of TMAO in dialysis patients are more than 5–10 fold higher than in non-dialysis CKD patients, and 30-fold higher than in healthy people. Accumulating clinical studies have demonstrated that elevated TMAO levels predict an increased risk of major adverse cardiovascular events in CKD patients [53]. A study of hemodialysis patients showed higher TMAO levels were associated with increased risks of cardiovascular events and all-cause death [54]. Seldin et al. found that TMAO could induce inflammatory marker expression in vascular endothelial cells, as well as activate mitogen-activated protein kinase (MAPK) and NF-jB signaling cascade [55]. Translocation of the NF-jB p65 subunit to the nucleus turns on the transcription of many target genes, such as inflammatory cytokines and the NLRP3 inflammasome [56, 57]. A previous study demonstrated that elevated TMAO levels portend poorer prognosis of kidney within non-CKD patients. Experiments in animal models showed that dietary choline or TMAO directly led to progressive renal tubulointerstitial fibrosis and dysfunction by up-regulating TGF-b production [58]. However, further mechanism studies are needed to determine how TMAO directly accelerates the progression of chronic kidney disease.

7.6

Modulations of Intestinal Flora Disorder in CKD/ ESRD

Since gut microbiota imbalance is associated with increased circulating levels of gut-derived uremic toxins, inflammation and oxidative stress, which are linked to cardiovascular disease and increased morbimortality of CKD patients, various nutritional strategies have been proposed to improve dysbiosis in CKD.

7.6.1

Dietary Interventions

The first approach is to optimize dietary components. Some research has focused on the use of dietary fibers. It has been shown that 56.4% of CKD patients had insufficient cellulose intake, and decreased soluble cellulose intake could lead to an increase in all-cause mortality in CKD patients [59]. Rossi et al. conducted a cohort study on CKD stage 4 and 5 patients, and found that the protein/cellulose ratio in patient diet was positively correlated with serum levels of IS and PCS [60]. Increasing the intake of soluble cellulose in diet is beneficial to the growth of Bacteroides, which promote the proliferation of glycolytic bacteria such as Bifidobacillus and Lactobacillus. The production of SCFAs by these bacteria can also play a beneficial role in reducing systemic inflammation [61]. Another animal

7 Gut Microbiota and Renal Injury

99

study also showed that supplementation of fermentable dietary fibers could increase the expression of colonic tight junction proteins including ZO-1, occludin and claudin 7, as well as cecal SCFA concentration [62]. Therefore, in dietary guidance, CKD patients should be taught to eat coarse grains and cellulose-rich food in high-quality and low-protein diet. As mentioned above, TMAO mainly derives from the intestinal flora metabolites of red meat diet, and serum TMAO levels of CKD patients increase with the decline of renal function. Recent clinical studies revealed that TMA and TMAO production was substantially reduced in vegans/vegetarians [63]. Wang et al. demonstrated a significant correlation between serum TMAO level and red meat intake, and that discontinuation of dietary red meat reduced plasma TMAO within 4 weeks [63]. The PREDIMED study showed that a Mediterranean diet supplemented with extra-virgin olive oil or nuts could decrease the incidence of major cardiovascular events in a population at high cardiovascular risk [64]. Reducing the conversion of TMA to TMAO by inhibiting human FMO3 is not ideal, because direct FMO3 inhibition is usually accompanied by severe hepatic inflammation. Wang et al. treated mice with 3,3-dimethyl-1-butanol (DMB), an analog of choline, to feedback inhibit microbial TMA lyase, and thus raised a new approach to inhibit TMAO production [65].

7.6.2

Probiotics, Prebiotics and Synbiotics

Probiotics are defined as live microorganisms that maintain health and prevent disease of the host when administered in adequate amounts [66]. Bifidobacterium and Lactobacillus represent well-studied species that probably have some general benefits [67]. Several studies demonstrate that the administration of probiotics could significantly reduce the serum level of uremia toxins such as PCS, IS or dimethylamine nitrosamine in CKD patients [67–69]. In addition, probiotic supplementation could also decrease serum levels of endotoxin and pro-inflammatory cytokine IL-6 in peritoneal dialysis patients [70]. Mechanisms of the probiotics include competition with pathogenic bacteria for nutrients as well as inhibiting their adhesion, maintaining the integrity of the gut barrier, and modulation of immune response [70, 71]. Another group of microflora regulators are called prebiotics, mainly including inulin, fructooligosaccharides (FOS) and galactooligosaccharides (GOS), which are also found in some fruits, cereals, vegetables and in human milk [72]. Prebiotics serve as nutrients for microbial metabolism, which selectively promotes beneficial microorganisms [73]. Meijers et al. reported that prebiotic oligofructose inulin significantly reduced serum PCS concentrations in hemodialysis patients [74]. Several studies showed that, in animal models or CKD patients, a high amylose-resistant starch diet could increase amounts of Bifidobacteria and Lactobacillus, and decrease levels of PCS and IS [74]. Recently, an RCT study showed that supplementation of amylose-resistant starch, HAM-RS2, in CKD

100

L. Zhang et al.

patients could increase the proportion of Faecalibacterium in the fecal microbiota of CKD patients, decrease serum urea, IL-6, TNFa and malondialdehyde, indicating an anti-inflammatory effect of prebiotic fibers on CKD patients [75]. However, the effect of probiotics or prebiotics alone on CKD patients remains controversial. A combination of prebiotics and probiotics called synbiotics is now considered necessary. The goal of this treatment is to selectively promote the growth of probiotic microorganisms. The combination of Bifidobacterium or Lactobacillus with fructooligosaccharides (FOS) has been commonly used in clinical studies [76]. Many studies report a decrease in serum levels of PCS, and some had amelioration of gastrointestinal symptom in hemodialysis or non-dialysis CKD patients [77, 78].

7.7

Therapeutic Drugs for Microbiota Regulation

At present, several drugs have been studied to correct dysbiosis-induced chronic inflammation by adsorbing intestinal toxins, promoting intestinal emptying or regulating TLR receptors. Phosphate binders are commonly used in ESRD patients. In addition to treating secondary hyperparathyroidism caused by hyperphosphatemia, several phosphate binders are believed to regulate intestinal flora of ESRD patients and promote the growth of probiotics. Ferric citrate complex is used as a dietary phosphate binder as well as an iron supplement. Lau et al. found that ferric citrate supplementation significantly changed gut microbiota in CKD patients, together with improved anemia, residual renal function and hypertension [79]. Sevelamer is an anion exchange resin with a backbone of multiple amines. Sevelamer has been reported to bind molecules such as LPS, indoles and phenols in vitro, as well as decrease serum levels of indoxyl sulfate and p-cresyl sulfate in hemodialysis patients [80]. In patients receiving peritoneal dialysis with type 2 diabetes mellitus, sevelamer treatment decreased serum PAI-1, CRP and IL-6 compared with the calcium carbonate treatment group [81]. Another randomized controlled trial showed that sevelamer treatment for 3 months significantly decreased serum levels of p-cresol in non-dialysis CKD patients [82]. As mentioned above, SCFAs are a product of intestinal flora with a protective effect on CVD events. Because sevelamer can also bind and remove SCFAs from the colon, the overall effect of this drug on intestinal flora remains to be explored. It should also be noted that sevelamer hydrochloride can lead to metabolic acidosis, so sevelamer carbonate is currently recommended [83]. AST-120 is a microsphere composed of porous carbon material, which can effectively absorb small molecules of intestinal toxins [84]. ESRD patients treated with AST-120 showed decreased levels of serum IS and PCS. An animal study confirmed that AST-120 not only directly absorbs toxins but also stabilizes the tight junction proteins of the intestinal epithelium, reduces endotoxin levels and relieves oxidative stress injury [85–87]. Some researches supported that AST-120 could ameliorates tubular injury in CKD patients by reducing proteinuria and oxidative

7 Gut Microbiota and Renal Injury

101

stress generation, as well as effectively reducing the serum concentration of IS [88]. Adesso et al. revealed that AST-120 could decrease the inflammation and oxidative stress in the primary central nervous system induced by IS, and therefore ameliorates neurodegeneration in CKD patients [89]. AST-120 has been approved in several countries for improving uremia symptoms and delaying the initiation of dialysis in CKD patients. However, in a multinational RCT study, the EPPIC trials, AST-120 didn’t show significant advantages with regards to protecting residual renal function when compared to placebo in patients with moderate to severe CKD. Notably, the median durations of treatment for the EPPIC trials were 102.1 and 96.3 weeks, respectively, while the median time to primary end points in EPPIC-1 and EPPIC-2 were 189.0 and 170.3 weeks. Therefore, a longer observation period might be required to determine whether there is truly a significant difference between AST-120 and the placebo [90]. Intestinal flora disorders in CKD patients directly affect fecal formation and intestinal emptying, leading to increased absorption of harmful metabolites and accumulation of uremia toxins. Rubiprostone, a ClC-2 chloride channel agonist, can promote intestinal water secretion and fecal formation, and is used as laxative. Mishima et al. treated CKD rats with rubiprostone orally and found that fecal formation and intestinal emptying of these rats were improved, and renal interstitial fibrosis was decreased compared with the control group. They also found that rubiprostone could elevate the proportions of Lactobacillus and Prevotella in the intestinal flora, and decrease serum levels of IS and anti-aconitic acid [91].

7.8

Summary and Perspectives

There are significant changes in both the number and composition of intestinal flora in CKD patients. Proteins and other nitrogenous substances are metabolized by intestinal flora, generating toxic products, which become important components of uremic toxins in ESRD patients. Intestinal barrier dysfunction and increased permeability lead to increased amounts of toxins entering circulation to cause endotoxemia and aggravates systemic inflammation, which promotes renal function decline, insulin resistance, immune dysfunction, atherosclerosis and cardiovascular events in CKD patients [92]. Several clinical trials have proved that therapeutic approaches based on intestinal flora regulation, such as dietary guidance, and prebiotic and probiotic supplementation, could effectively delay CKD progression. Although significant challenges remain in terms of figuring out which patterns among the enormous diversity of microorganisms are linked to delayed progression of CKD, it is expected that microbiota modulation will become a promising therapeutic strategy in complement to traditional drugs and dialysis therapy for CKD patients.

102

L. Zhang et al.

Acknowledgment This chapter was modified from a paper reported by our group in Journal of Clinical Nephrology (Lin Chuang Shen Zang Bing Za Zhi) (Zhang et al. 2016). The related contents are reused with permission.

References 1. Zhang L, Wang F, Wang L et al (2012) Prevalence of chronic kidney disease in China: a cross-sectional survey. Lancet 379(9818):815–822 2. Mills KT, Xu Y, Zhang W et al (2015) A systematic analysis of worldwide population-based data on the global burden of chronic kidney disease in 2010. Kidney Int 88(5):950–957 3. Meguid ENA, Bello AK (2005) Chronic kidney disease: the global challenge. Lancet 365 (9456):331–340 4. Nallu A, Sharma S, Ramezani A, Muralidharan J, Raj D (2017) Gut microbiome in chronic kidney disease: challenges and opportunities. Transl Res 179:24–37 5. Chen YY, Chen DQ, Chen L et al (2019) Microbiome-metabolome reveals the contribution of gut-kidney axis on kidney disease. J Transl Med 17(1):5 6. Li DY, Tang W (2018) Contributory role of gut microbiota and their metabolites toward cardiovascular complications in chronic kidney disease. Semin Nephrol 38(2):193–205 7. Rooks MG, Garrett WS (2016) Gut microbiota, metabolites and host immunity. Nat Rev Immunol 16(6):341–352 8. Sabatino A, Regolisti G, Brusasco I, Cabassi A, Morabito S, Fiaccadori E (2015) Alterations of intestinal barrier and microbiota in chronic kidney disease. Nephrol Dial Transplant 30 (6):924–933 9. Vaziri ND, Wong J, Pahl M et al (2013) Chronic kidney disease alters intestinal microbial flora. Kidney Int 83(2):308–315 10. Jiang S, Xie S, Lv D et al (2017) Alteration of the gut microbiota in Chinese population with chronic kidney disease. Sci Rep 7(1):2870 11. Hida M, Aiba Y, Sawamura S, Suzuki N, Satoh T, Koga Y (1996) Inhibition of the accumulation of uremic toxins in the blood and their precursors in the feces after oral administration of Lebenin, a lactic acid bacteria preparation, to uremic patients undergoing hemodialysis. Nephron 74(2):349–355 12. Sud K, Sakhuja V (1997) The gastrointestinal tract in uremia. J Assoc Physicians India 45 (11):833–834 13. Kang JY (1993) The gastrointestinal tract in uremia. Dig Dis Sci 38(2):257–268 14. Vaziri ND, Freel RW, Hatch M (1995) Effect of chronic experimental renal insufficiency on urate metabolism. J Am Soc Nephrol 6(4):1313–1317 15. Turner JR (2009) Intestinal mucosal barrier function in health and disease. Nat Rev Immunol 9(11):799–809 16. Ewaschuk JB, Diaz H, Meddings L et al (2008) Secreted bioactive factors from Bifidobacterium infantis enhance epithelial cell barrier function. Am J Physiol Gastrointest Liver Physiol 295(5):G1025–G1034 17. Nakashima K, Kimura S, Ogawa Y et al (2018) Chitin-based barrier immunity and its loss predated mucus-colonization by indigenous gut microbiota. Nat Commun 9(1):3402 18. Vaziri ND, Dure-Smith B, Miller R, Mirahmadi MK (1985) Pathology of gastrointestinal tract in chronic hemodialysis patients: an autopsy study of 78 cases. Am J Gastroenterol 80 (8):608–611 19. Vaziri ND, Yuan J, Nazertehrani S, Ni Z, Liu S (2013) Chronic kidney disease causes disruption of gastric and small intestinal epithelial tight junction. Am J Nephrol 38(2):99–103 20. Szeto CC, Kwan BC, Chow KM et al (2008) Endotoxemia is related to systemic inflammation and atherosclerosis in peritoneal dialysis patients. Clin J Am Soc Nephrol 3(2):431–436

7 Gut Microbiota and Renal Injury

103

21. McIntyre CW, Harrison LE, Eldehni MT et al (2011) Circulating endotoxemia: a novel factor in systemic inflammation and cardiovascular disease in chronic kidney disease. Clin J Am Soc Nephrol 6(1):133–141 22. Vaziri ND, Yuan J, Rahimi A, Ni Z, Said H, Subramanian VS (2012) Disintegration of colonic epithelial tight junction in uremia: a likely cause of CKD-associated inflammation. Nephrol Dial Transplant 27(7):2686–2693 23. Andersen K, Kesper MS, Marschner JA et al (2017) Intestinal dysbiosis, barrier dysfunction, and bacterial translocation account for CKD-related systemic inflammation. J Am Soc Nephrol 28(1):76–83 24. Carlsson AC, Larsson TE, Helmersson-Karlqvist J, Larsson A, Lind L, Arnlov J (2014) Soluble TNF receptors and kidney dysfunction in the elderly. J Am Soc Nephrol 25(6):1313– 1320 25. Vinolo MA, Rodrigues HG, Nachbar RT, Curi R (2011) Regulation of inflammation by short chain fatty acids. Nutrients 3(10):858–876 26. Gonzalez A, Krieg R, Massey HD et al (2019) Sodium butyrate ameliorates insulin resistance and renal failure in CKD rats by modulating intestinal permeability and mucin expression. Nephrol Dial Transplant 34(5):783–794 27. Wang S, Lv D, Jiang S, Jiang J, Liang M, Hou F, Chen Y (2019) Quantitative reduction in short-chain fatty acids, especially butyrate, contributes to the progression of chronic kidney disease. Clin Sci (Lond) 133(17):1857–1870 28. Maslowski KM, Vieira AT, Ng A et al (2009) Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature 461(7268):1282–1286 29. Yang G, Chen S, Deng B, Tan C, Deng J, Zhu G, Yin Y, Ren W (2018) Implication of G protein-coupled receptor 43 in intestinal inflammation: a mini-review. Front Immunol 9:1434 30. Waldecker M, Kautenburger T, Daumann H, Busch C, Schrenk D (2008) Inhibition of histone-deacetylase activity by short-chain fatty acids and some polyphenol metabolites formed in the colon. J Nutr Biochem 19(9):587–593 31. Andrade-Oliveira V, Amano MT, Correa-Costa M et al (2015) Gut bacteria products prevent AKI induced by ischemia-reperfusion. J Am Soc Nephrol 26(8):1877–1888 32. Candido EP, Reeves R, Davie JR (1978) Sodium butyrate inhibits histone deacetylation in cultured cells. Cell 14(1):105–113 33. Park J, Kim M, Kang SG et al (2015) Short-chain fatty acids induce both effector and regulatory T cells by suppression of histone deacetylases and regulation of the mTOR-S6K pathway. Mucosal Immunol 8(1):80–93 34. Arpaia N, Campbell C, Fan X et al (2013) Metabolites produced by commensal bacteria promote peripheral regulatory T-cell generation. Nature 504(7480):451–455 35. Viaene L, Annaert P, de Loor H, Poesen R, Evenepoel P, Meijers B (2013) Albumin is the main plasma binding protein for indoxyl sulfate and p-cresyl sulfate. Biopharm Drug Dispos 34(3):165–175 36. Cornelis T, van der Sande FM, Eloot S et al (2014) Acute hemodynamic response and uremic toxin removal in conventional and extended hemodialysis and hemodiafiltration: a randomized crossover study. Am J Kidney Dis 64(2):247–256 37. Eloot S, Van Biesen W, Glorieux G, Neirynck N, Dhondt A, Vanholder R (2013) Does the adequacy parameter Kt/V(urea) reflect uremic toxin concentrations in hemodialysis patients? PLoS ONE 8(11):e76838 38. Meyer TW, Hostetter TH (2012) Uremic solutes from colon microbes. Kidney Int 81 (10):949–954 39. Wu IW, Hsu KH, Lee CC et al (2011) p-Cresyl sulphate and indoxyl sulphate predict progression of chronic kidney disease. Nephrol Dial Transplant 26(3):938–947 40. Meijers BK, Claes K, Bammens B et al (2010) p-Cresol and cardiovascular risk in mild-to-moderate kidney disease. Clin J Am Soc Nephrol 5(7):1182–1189

104

L. Zhang et al.

41. Ito S, Osaka M, Higuchi Y, Nishijima F, Ishii H, Yoshida M (2010) Indoxyl sulfate induces leukocyte-endothelial interactions through up-regulation of E-selectin. J Biol Chem 285 (50):38869–38875 42. Matsuo K, Yamamoto S, Wakamatsu T et al (2015) Increased proinflammatory cytokine production and decreased cholesterol efflux due to downregulation of ABCG1 in macrophages exposed to indoxyl sulfate. Toxins (Basel) 7(8):3155–3166 43. Masai N, Tatebe J, Yoshino G, Morita T (2010) Indoxyl sulfate stimulates monocyte chemoattractant protein-1 expression in human umbilical vein endothelial cells by inducing oxidative stress through activation of the NADPH oxidase-nuclear factor-kappaB pathway. Circ J 74(10):2216–2224 44. Gross P, Massy ZA, Henaut L et al (2015) Para-cresyl sulfate acutely impairs vascular reactivity and induces vascular remodeling. J Cell Physiol 230(12):2927–2935 45. Han H, Zhu J, Zhu Z et al (2015) p-Cresyl sulfate aggravates cardiac dysfunction associated with chronic kidney disease by enhancing apoptosis of cardiomyocytes. J Am Heart Assoc 4 (6):e1852 46. Sun CY, Chang SC, Wu MS (2012) Uremic toxins induce kidney fibrosis by activating intrarenal renin-angiotensin-aldosterone system associated epithelial-to-mesenchymal transition. PLoS ONE 7(3):e34026 47. Lekawanvijit S, Adrahtas A, Kelly DJ, Kompa AR, Wang BH, Krum H (2010) Does indoxyl sulfate, a uraemic toxin, have direct effects on cardiac fibroblasts and myocytes? Eur Heart J 31(14):1771–1779 48. Watanabe H, Miyamoto Y, Honda D et al (2013) p-Cresyl sulfate causes renal tubular cell damage by inducing oxidative stress by activation of NADPH oxidase. Kidney Int 83(4):582– 592 49. Wang Z, Klipfell E, Bennett BJ et al (2011) Gut flora metabolism of phosphatidylcholine promotes cardiovascular disease. Nature 472(7341):57–63 50. Tang WH, Wang Z, Levison BS et al (2013) Intestinal microbial metabolism of phosphatidylcholine and cardiovascular risk. N Engl J Med 368(17):1575–1584 51. Tang WH, Hazen SL (2014) The contributory role of gut microbiota in cardiovascular disease. J Clin Invest 124(10):4204–4211 52. Koeth RA, Wang Z, Levison BS et al (2013) Intestinal microbiota metabolism of L-carnitine, a nutrient in red meat, promotes atherosclerosis. Nat Med 19(5):576–585 53. Tomlinson J, Wheeler DC (2017) The role of trimethylamine N-oxide as a mediator of cardiovascular complications in chronic kidney disease. Kidney Int 92(4):809–815 54. Shafi T, Powe NR, Meyer TW et al (2017) Trimethylamine N-oxide and cardiovascular events in hemodialysis patients. J Am Soc Nephrol 28(1):321–331 55. Seldin MM, Meng Y, Qi H et al (2016) Trimethylamine N-oxide promotes vascular inflammation through signaling of mitogen-activated protein kinase and nuclear factor-kappaB. J Am Heart Assoc 5(2) 56. El-Deeb OS, Atef MM, Hafez YM (2019) The interplay between microbiota-dependent metabolite trimethylamine N-oxide, Transforming growth factor beta/SMAD signaling and inflammasome activation in chronic kidney disease patients: a new mechanistic perspective. J Cell Biochem 120(9):14476–14485 57. Chen ML, Zhu XH, Ran L, Lang HD, Yi L, Mi MT (2017) Trimethylamine-N-oxide induces vascular inflammation by activating the NLRP3 inflammasome through the SIRT3-SOD2-mtROS signaling pathway. J Am Heart Assoc 6(9) 58. Tang WH, Wang Z, Kennedy DJ et al (2015) Gut microbiota-dependent trimethylamine N-oxide (TMAO) pathway contributes to both development of renal insufficiency and mortality risk in chronic kidney disease. Circ Res 116(3):448–455 59. Krishnamurthy VM, Wei G, Baird BC et al (2012) High dietary fiber intake is associated with decreased inflammation and all-cause mortality in patients with chronic kidney disease. Kidney Int 81(3):300–306

7 Gut Microbiota and Renal Injury

105

60. Rossi M, Johnson DW, Xu H et al (2015) Dietary protein-fiber ratio associates with circulating levels of indoxyl sulfate and p-cresyl sulfate in chronic kidney disease patients. Nutr Metab Cardiovasc Dis 25(9):860–865 61. Trompette A, Gollwitzer ES, Yadava K et al (2014) Gut microbiota metabolism of dietary fiber influences allergic airway disease and hematopoiesis. Nat Med 20(2):159–166 62. Hung TV, Suzuki T (2018) Dietary fermentable fibers attenuate chronic kidney disease in mice by protecting the intestinal barrier. J Nutr 148(4):552–561 63. Conlon MA, Bird AR (2014) The impact of diet and lifestyle on gut microbiota and human health. Nutrients 7(1):17–44 64. Stewart R (2018) Primary prevention of cardiovascular disease with a Mediterranean diet supplemented with extra-virgin olive oil or nuts. N Engl J Med 379(14):1388 65. Wang Z, Roberts AB, Buffa JA et al (2015) Non-lethal inhibition of gut microbial trimethylamine production for the treatment of atherosclerosis. Cell 163(7):1585–1595 66. Hill C, Guarner F, Reid G et al (2014) Expert consensus document. The International Scientific Association for Probiotics and Prebiotics consensus statement on the scope and appropriate use of the term probiotic. Nat Rev Gastroenterol Hepatol 11(8):506–514 67. Guarner F, Khan AG, Garisch J et al (2012) World Gastroenterology Organisation Global Guidelines: probiotics and prebiotics October 2011. J Clin Gastroenterol 46(6):468–481 68. Simenhoff ML, Dunn SR, Zollner GP et al (1996) Biomodulation of the toxic and nutritional effects of small bowel bacterial overgrowth in end-stage kidney disease using freeze-dried Lactobacillus acidophilus. Miner Electrolyte Metab 22(1–3):92–96 69. Guida B, Germano R, Trio R et al (2014) Effect of short-term synbiotic treatment on plasma p-cresol levels in patients with chronic renal failure: a randomized clinical trial. Nutr Metab Cardiovasc Dis 24(9):1043–1049 70. Wang IK, Wu YY, Yang YF et al (2015) The effect of probiotics on serum levels of cytokine and endotoxin in peritoneal dialysis patients: a randomised, double-blind, placebo-controlled trial. Benef Microbes 6(4):423–430 71. van Baarlen P, Troost F, van der Meer C et al (2011) Human mucosal in vivo transcriptome responses to three lactobacilli indicate how probiotics may modulate human cellular pathways. Proc Natl Acad Sci U S A 108(Suppl 1):4562–4569 72. Slavin J (2013) Fiber and prebiotics: mechanisms and health benefits. Nutrients 5(4):1417– 1435 73. Simpson HL, Campbell BJ (2015) Review article: dietary fibre-microbiota interactions. Aliment Pharmacol Ther 42(2):158–179 74. Meijers BK, De Preter V, Verbeke K, Vanrenterghem Y, Evenepoel P (2010) p-Cresyl sulfate serum concentrations in haemodialysis patients are reduced by the prebiotic oligofructose-enriched inulin. Nephrol Dial Transplant 25(1):219–224 75. Laffin MR, Tayebi KH, Park H et al (2019) Amylose resistant starch (HAM-RS2) supplementation increases the proportion of Faecalibacterium bacteria in end-stage renal disease patients: microbial analysis from a randomized placebo-controlled trial. Hemodial Int 23(3):343–347 76. Mafra D, Borges N, Alvarenga L et al (2019) Dietary components that may influence the disturbed gut microbiota in chronic kidney disease. Nutrients 11(3) 77. Nakabayashi I, Nakamura M, Kawakami K et al (2011) Effects of synbiotic treatment on serum level of p-cresol in haemodialysis patients: a preliminary study. Nephrol Dial Transplant 26(3):1094–1098 78. Cruz-Mora J, Martinez-Hernandez NE, Martin DCF et al (2014) Effects of a symbiotic on gut microbiota in Mexican patients with end-stage renal disease. J Ren Nutr 24(5):330–335 79. Lau WL, Vaziri ND, Nunes A et al (2018) The phosphate binder ferric citrate alters the gut microbiome in rats with chronic kidney disease. J Pharmacol Exp Ther 367(3):452–460 80. Lin CJ, Pan CF, Chuang CK et al (2017) Effects of sevelamer hydrochloride on uremic toxins serum indoxyl sulfate and P-cresyl sulfate in hemodialysis patients. J Clin Med Res 9(9):765– 770

106

L. Zhang et al.

81. Chennasamudram SP, Noor T, Vasylyeva TL (2013) Comparison of sevelamer and calcium carbonate on endothelial function and inflammation in patients on peritoneal dialysis. J Ren Care 39(2):82–89 82. Riccio E, Sabbatini M, Bruzzese D et al (2018) Plasma p-cresol lowering effect of sevelamer in non-dialysis CKD patients: evidence from a randomized controlled trial. Clin Exp Nephrol 22(3):529–538 83. Oka Y, Miyazaki M, Matsuda H et al (2014) Sevelamer hydrochloride dose-dependent increase in prevalence of severe acidosis in hemodialysis patients: analysis of nationwide statistical survey in Japan. Ther Apher Dial 18(1):37–43 84. Bolati D, Shimizu H, Niwa T (2012) AST-120 ameliorates epithelial-to-mesenchymal transition and interstitial fibrosis in the kidneys of chronic kidney disease rats. J Ren Nutr 22 (1):176–180 85. Vaziri ND, Yuan J, Khazaeli M, Masuda Y, Ichii H, Liu S (2013) Oral activated charcoal adsorbent (AST-120) ameliorates chronic kidney disease-induced intestinal epithelial barrier disruption. Am J Nephrol 37(6):518–525 86. Nakamura T, Sato E, Fujiwara N et al (2011) Oral adsorbent AST-120 ameliorates tubular injury in chronic renal failure patients by reducing proteinuria and oxidative stress generation. Metabolism 60(2):260–264 87. Shibahara H, Shibahara N (2010) Cardiorenal protective effect of the oral uremic toxin absorbent AST-120 in chronic heart disease patients with moderate CKD. J Nephrol 23 (5):535–540 88. Yoshifuji A, Wakino S, Irie J et al (2018) Oral adsorbent AST-120 ameliorates gut environment and protects against the progression of renal impairment in CKD rats. Clin Exp Nephrol 22(5):1069–1078 89. Adesso S, Paterniti I, Cuzzocrea S et al (2018) AST-120 reduces neuroinflammation induced by indoxyl sulfate in glial cells. J Clin Med 7(10) 90. Schulman G, Berl T, Beck GJ et al (2015) Randomized placebo-controlled EPPIC trials of AST-120 in CKD. J Am Soc Nephrol 26(7):1732–1746 91. Mishima E, Fukuda S, Shima H et al (2015) Alteration of the intestinal environment by lubiprostone is associated with amelioration of adenine-induced CKD. J Am Soc Nephrol 26 (8):1787–1794 92. Meijers B, Evenepoel P, Anders HJ (2019) Intestinal microbiome and fitness in kidney disease. Nat Rev Nephrol 15(9):531–545

Chapter 8

Gut Microbiota and Heart, Vascular Injury Cheng Zeng and Hongmei Tan

Abstract The gut microbiota plays an important role in maintaining human health. Accumulating evidence has indicated an intimate relationship between gut microbiota and cardiovascular diseases (CVD) which has become the leading cause of death worldwide. The alteration of gut microbial composition (gut dysbiosis) has been proven to contribute to atherosclerosis, the basic pathological process of CVD. In addition, the metabolites of gut microbiota have been found to be closely related to the development of CVD. For example, short-chain fatty acids are widely acclaimed beneficial effect against CVD, whereas trimethylamine-N-oxide is considered as a contributing factor in the development of CVD. In this chapter, we mainly discuss the gut microbial metabolite-involved mechanisms of CVD focusing on atherosclerosis, hypertension, diabetes, obesity, and heart failure. Targeting gut microbiota and related metabolites are novel and promising strategies for the treatment of CVD.





Keywords Gut microbiota Gut microbial metabolites Cardiovascular diseases Atherosclerosis Hypertension Diabetes mellitus Obesity



8.1







Introduction

The human gut hosts a vast number of commensal microorganisms including bacteria, viruses, protozoa, achaea, and fungi. The complexity of this ecosystem is still far from being completely understood and analyzed, although important steps have been made in this direction. It is estimated that the cumulative genome of all these microorganisms consists of up to 100 trillion microbes and contains more than 150 times as many genes as the whole human genome [1]. Gut microbiota is dominated by two bacterial phyla, Bacteroidetes or Firmicutes, which constitute C. Zeng  H. Tan (&) Department of Pathophysiology, Zhongshan School of Medicine, Sun Yat-sen University, Guangzhou 510080, China e-mail: [email protected] © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_8

107

108

C. Zeng and H. Tan

>90% of the taxa present in the human gut, and to a lesser extent by Actinobacteria, Cyanobacteria, Fusobacteria, Proteobacteria, and Verrucomicrobia [2], and these constantly adapt to lifestyle modifications [3]. Microorganisms benefit from a constant supply of human substrates. The human hosts benefit from microbial activities such as host metabolism, neurological development, energy homeostasis, instructing immune cell differentiation [4], vitamin synthesis, and digestion of carbohydrates as well as production of signaling molecules including short-chain fatty acids (SCFAs), bile acids (BAs), and so on [5]. The microbiota is capable of secreting or altering the production of molecules that affect host physiology and the microbial composition [6]. The homeostasis of gut microbiota is critical for maintaining human health. Over the past decade, studies have demonstrated that composition and extent of alterations of the gut microbiota (so-called “dysbiosis”) contributed to pathogenesis of various diseases including cardiovascular disease and cancers [7, 8]. Cardiovascular disease (CVD) is still the most common cause of death causing almost two times as many deaths as cancer across the worldwide [9]. In addition to genetic factors, other environmental factors such as nutrition and gut microbiota have also been recognized as the main contributor for developing CVD. Moreover, gut dysbiosis has been implicated as a risk factor for the development of diabetes and obesity, two major risk factors for CVD [10, 11]. Mountains of studies demonstrated that the gut microbiota affects the host via various distinct pathways, such as production of gut microbial metabolites. In this review, we discussed the roles of gut microbiota implicated in the development of CVD, mainly focusing on atherosclerosis, hypertension and the associated metabolic diseases, diabetes, and obesity. We further systematically overview the interplay between gut microbial metabolites and CVD, and summarize the recent advances of gut microbiota-targeted therapies for CVD.

8.2

The Interaction Between Host Metabolism and Gut Microbial Metabolites

Microbial metabolites are indispensable for the majority of the biological effects of the gut microbiota. Under physiological conditions, indigestible carbohydrates (for example, dietary fiber) and peptides/proteins can be actively fermented by commensal microbiota in the intestine. Gut microbiome metabolites that are associated with the onset of CVD include trimethylamine-N-oxide (TMAO), BAs, SCFAs, B vitamins, protocatechuic acid (PCA), 4-ethyl phenol sulfate (4-EPS), D-glycerob-D-manno-heptose-1,7-biphosphate (HBP), and other metabolites, such as phenylacetylglutamine, p-cresyl sulfate, indoxyl sulfate (IS), enterolactone, and hydrogen sulfide (H2S). Among these metabolites, TMAO, BAs, SCFAs, and H2S have been well studied in CVD.

8 Gut Microbiota and Heart, Vascular Injury

8.2.1

109

TMAO

Intestinal microbiota cleave trimethylamine containing compounds to produce trimethylamine (TMA), which is translocated to liver, oxidized by flavin monooxygenase 3 (FMO3) and further metabolized to TMAO. The precursors for gut microbiota to produce TMA include TMAO, choline, phosphatidylcholine, carnitine, c-butyrobetaine, betaine, crotonobetaine, and glycerophosphocholine, all of which are rich in red meat and fat [12, 13].

8.2.2

BAs

Primary BAs synthesized from cholesterol in the liver are chenodeoxycholic acid (CDCA) and cholic acid (CA) which may be conjugated with glycine or taurine, stored, and concentrated in gallbladder. The primary bile acids are produced in the liver via two pathways: the classical and the alternative pathway (80% vs. 20% of the bile acid pool). The classical pathway generates primary bile acids CDCA and CA, while the alternative pathway mainly generates CDCA; 12a-hydroxylase (CYP8B1) determined the ratio between them. The CYP7A1 gene encodes the enzyme cholesterol 7a-hydroxylase, which is the first and rate-limiting enzyme in the classical pathway of total bile-acid synthesis [14]. In rodents, there are additional primary bile acids UDCA and a/b-MCA. CDCA and UDCA are 6b-hydroxylated to product a- and b-MCAs by cytochrome P450 2C70 (CYP2C70) [15]. Primary BAs are metabolized by the gut microbiota-mediated deconjugation, dehydroxylation, dehydrogenation, and epimerization into secondary BAs, which are stored in the gallbladder and released into the duodenum to digest and enhance absorption of triglycerides, cholesterol, and lipid-soluble vitamins [16]. Secondary BAs include deoxycholic acid (DCA), hyocholic acid (HCA), hyodeoxycholic acid (HDCA), lithocholic acid (LCA), x-muricholic acid (xMCA), and murideoxycholic acid (MDCA) [15]. In the ileum, these bile acids are then reabsorbed and carried in the portal blood to the liver. This process is called enterohepatic circulation which preserves about 95% of the bile acid pool [17].

8.2.3

SCFAs

Dietary fibers, proteins and peptides which escape digestion by host enzymes in the upper gut, are metabolized by the microbiota in the cecum and colon [18] to produce three major SCFAs: acetic acid, propionic acid, and butyric acid; and two less abundant valeric acid and hexanoic acid [19]. The proposed biosynthesis of SCFAs in bacteria is from glycolysis of glucose to pyruvate, then to acetylcoA, and eventually to acetic acid, propionic acid, and butyric acid. These SCFAs either are

110

C. Zeng and H. Tan

used locally as fuel for colonic mucosal epithelial cells or enter the portal bloodstream to affect host health. Among the three major SCFAs, acetic acid, the most abundant SCFA in the colon comprising more than half of the total SCFA, can be generated by carbohydrate fermentation. Interestingly, major acetate-producing bacteria include Streptococcus spp., Prevotella spp., Clostridium spp. and Bifidobacterium spp., A. muciniphila, etc. [20]. The main propionate-producing bacteria are Bacteroides spp., Salmonella spp., Dialister spp., Veillonella spp., Roseburia inulinivorans, Coprococcus Catus, Blautia obeum, and so on [21]. And the main butyric acid-producing bacteria refer to Lachnospiraceae, Ruminococcaceae, and Acidaminococcaceae families [22]. More interestingly, dietary fiber can selectively increase abundance of SCFA-producing bacteria [23].

8.2.4

H2S

H2S is a colorless, flammable, water-soluble gas with the characteristic smell of rotten eggs [24]. Until recently, H2S is mostly associated with its toxic effects. However, research conducted over the last decade revealed that H2S serves biological signaling and function in numerous biological systems. Gut microbiota metabolizes carbohydrates, proteins, fat, and many other compounds from food uptake to produce H2S. Sulfate-reducing bacteria are the main H2S-generating bacteria that are widely distributed in mammalian colon [25]. The major genera for H2S production are Desulfovibrio, Desulfobacter, Desulfobulbus, and Desulfotomaculum [25]. These sulfate-reducing gut bacteria transfer a sulfate and an electron donor for the sulfate reduction to generate H2S. Therefore, a sulfate-rich diet may result in H2S production in the gut of humans and mice. Second source is an enzymatic reaction performed by either gut bacteria or colonic tissues. Several anaerobic bacterial strains Escherichia coli, Salmonella enterica, Clostridia, and Enterobacter aerogenes could catalyze cysteine to produce H2S by cysteine desulfhydrase. Moreover, sulfite reductase presented in some bacterial strains was also involved in generation of H2S. Finally, H2S is synthesized by mammalian tissues via cystathionine beta-synthase (CBS), cystathionine gamma-lyase (CSE), and 3-mercaptopyruvate sulfurtransferase (3-MST) responsible for metabolism of L-cysteine [25]. CSE seems to be a major source of the gut H2S generation [26].

8.3

Gut Microbiota and Atherosclerosis

Atherosclerosis is the main pathological basis for CVD, which is characterized by accumulation of cholesterol and recruitment of macrophages into artery walls and the consequent formation of atherosclerotic plaques. Emerging evidence has suggested an important role of gut microbiota in the progression of atherosclerosis [27]. By using shotgun sequencing of the gut metagenome in patients with or without

8 Gut Microbiota and Heart, Vascular Injury

111

symptomatic atherosclerosis, researchers have found that relative abundance of the genus Collinsella was higher in patients with symptomatic atherosclerosis, defined as stenotic atherosclerotic plaques in the carotid artery, leading to cerebrovascular disorders. In contrast, Roseburia and Eubacterium were enriched in healthy controls [28]. Moreover, Akkermansia muciniphila, a mucin-degrading bacterium belonging to a genus of Verrucomicrobia, has recently emerged as an important component of the gut microbial ecosystem and reversed western diet-induced increases in gut permeability to exert protective effects against atherosclerosis [29]. More importantly, bacterial DNA has been detected in atherosclerotic plaques, and the pyrosequencing result, suggesting that the bacteria in lesions may be derived from the gut and oral cavity that might thus serve as reservoirs of these potentially pathogenic microorganisms. On the other hand, germ-free (GF) conditions accelerated the atherosclerosis in ApoE−/− mice-fed standard low-cholesterol diet [30], and antibiotic therapy aimed at killing the gut bacteria also had no beneficial effect on cardiovascular events in human trials [31]. Thus, it is possible that modulation rather than clearance of the gut microbiota is a promising therapeutic strategy for atherosclerosis.

8.3.1

Gut Dysbiosis and Atherosclerosis

A large body of evidence suggests that infectious agents may contribute to atherosclerotic processes. This could occur by a direct infection of vessel wall cells and/or a distant infection by induction of cytokine and acute phase reactant proteins by infection at other sites. This notion is supported by evidence that bacterial DNA has been detected in atherosclerotic plaques. The pyrosequencing result revealed that the bacteria in lesions were predominantly derived from gut and oral cavity. A clinical study demonstrated that patients with symptomatic atherosclerosis (myocardial infarction or stroke) had a higher abundance of Anaeroglobus compared with healthy controls in the oral cavity. Several oral bacteria, including Porphyromonas gingivalis and Aggregatibacter actinomycetemcomitans, have also been reported to be a pathogenic bacterium as increased lesions were observed in ApoE−/− and apolipoprotein E-deficient spontaneously hyperlipidemic (Apoe(shl)) mice by oral or intravenous infection [32, 33]. These results implicated that bacteria from the oral cavity might translocate to the vessel, where they might induce rupture of the atherosclerotic plaque. The gut is another source of microorganisms that could influence the development of atherosclerosis. Gut dysbiosis increased intestinal permeability by inhibiting the expression of tight junction proteins and subsequently resulted in the translocation of LPS, disposition of gut microbiota, into the blood [34]. LPS, as part of the outer membrane of Gram-negative bacteria, could recruit adaptor proteins such as myeloid differentiation of primary response protein MYD88 to the cytoplasmic domain of TLRs which further activated downstream signaling and resulted in production of proinflammatory cytokines and chemokines [35]. As such, LPS

112

C. Zeng and H. Tan

induces low-grade inflammation and promotes the progression of atherosclerosis [36]. In contrast, gavage with live Bacteroides vulgatus and Bacteroides dorei inhibited atherosclerotic plague formation in atherosclerosis-prone mice by strengthening tight junction formation and decreasing gut microbial LPS production [37], suggesting a novel therapeutic strategy for atherosclerosis by decreasing gut microbiota-derived LPS.

8.3.2

BAs and Atherosclerosis

BAs, as signaling molecules, not only regulate their own biosynthesis but also affect metabolic pathways involved in lipoprotein, glucose, drug, and energy metabolism by activation of nuclear receptors [38, 39] (Fig. 8.1). BA receptor, also known as farnesoid X receptor (FXR), is a dedicated nuclear receptor for bile acids. In liver, activation of FXR mediated triglyceride metabolism involving the production of VLDL and de novo lipogenesis and has also regulated steatosis and obesity in the small intestine [40].

Secondary BAs

TGR5 GTP

GTP AC

cAMP cAMP

eNOS AKT

NO

Ca2+

FXR PXR

eNOS NO

BAs and triglyceride metabolism

Inflammation response

Fig. 8.1 BAs and possible molecular pathways linked to CVD. Gut bacteria are important in modifying BAs, which function as signaling molecules through the FXR, PXR, and TGR5. Activation of FXR and PXR not only promotes eNOS expression which induces NO release in endothelial cells but also regulates BAs and lipid metabolism in hepatic cells. In addition, TGR5 activation induces NO production via Akt activation and [Ca2+] increases in endothelial cells while inhibits inflammatory response via cAMP-mediated NF-jB inhibition in leukocytes

8 Gut Microbiota and Heart, Vascular Injury

113

However, conflicting data exist on the role of FXR on atherosclerosis progression. Compared with wild-type, FXR-deficient mice exhibit reduced expression of hepatic genes involved in reverse cholesterol transport (most notably for scavenger receptor B1), which lead to markedly reduction in plasma high-density lipoprotein (HDL) cholesterol ester clearance and profound hypercholesterolemia, suggesting that FXR is a critical regulator of cholesterol metabolism [41]. Conversely, activation of FXR reduced atherosclerotic plaque formation in apolipoprotein E (ApoE) or low-density lipoprotein receptor (LDLR)-deficient mice [42–44], while FXR knockout resulted in worse plasma lipid profile and larger atherosclerotic lesions in ApoE−/− mice [45]. These results suggest that FXR has protective role on atherosclerosis progression. In contrast, Guo and Zhang et al. reported that the FXR knockout decreased atherosclerotic lesion area in the aorta in ApoE−/− or LDLR−/− mice [46, 47], which suggests a nonprotective role for FXR. Therefore, the role of FXR on atherosclerosis need to be further understood in vivo and in vitro. The dysfunction of aortic vascular smooth muscle cells (VSMCs) and endothelial cells has been reported to promote plaque formation in atherosclerosis by inducing inflammatory response. Bacteria have the potential to directly regulate bile acid signaling in vasculature as VSMCs and endothelial cells also expressed FXR [47, 48]. Activation of FXR in VSMCs was shown to inhibit the inflammatory response and the migration of VSMCs. However, the role of endothelial-specific FXR on atherosclerosis progression seems to be controversial. A study in rabbits showed that FXR activation impaired endothelium-dependent relaxation due to a reduced sensitivity of smooth muscle to nitric oxide (NO) [49]. However, GW4064, an agonist of FXR, increased eNOS mRNA and protein expression and NO production in isolated endothelial cell, which induces vasodilatation [50]. Given that the role of microbial regulation of BAs and FXR in atherogenesis is not fully understood, additional investigation is warranted. Furthermore, most current data are based on animal models which may limit the translation ability to human subjects. G protein-coupled bile acid receptor 1 (GPBAR1; TGR5), another BA receptor, is responsive to BAs. Recent study has indicated that activation of TGR5 inhibited atherosclerosis, decreased intraplaque inflammation, and reduced plaque macrophage content which is a consequence of TGR5-induced cAMP signaling activation and the subsequent inhibition of NF-jB in macrophages [51]. In addition, taurolithocholic acid has a high affinity to TGR5 which significantly increased NO production and reduced inflammatory responses via activation of intracellular Akt and Ca2+ signaling in human umbilical vein endothelial cells [52]. Furthermore, pregnane X receptor (PXR) was also activated by BAs [53] and has been shown to accelerate development of atherosclerosis in PXR and ApoE double knockout mice [54]. This pro-atherosclerosis role of PXR attributes to the increase in CD36 expression and lipid uptake of macrophages [54] as well as the higher levels of atherogenic lipoproteins VLDL and LDL [55].

114

8.3.3

C. Zeng and H. Tan

TMA/TMAO and Atherosclerosis

Clinical studies indicated that TMAO level was associated with increased carotid intima-media thickness, risk of cardiovascular events, poor prognosis, death independent of traditional cardiac risk factors, lipid parameters, C-reactive protein, and even renal function [56]. In line with role of TMAO in CVD, circulating choline, betaine, and carnitine levels also have been shown associated with prevalence of CVD and can predict incident risk for major adverse cardiac events [13]. Zhu et al. showed that gut microbe-generating TMAO also directly contributes to platelet hyperreactivity, which is associated with cardiometabolic diseases and enhanced potential for thrombotic events [57]. In an animal study, dietary supplementation of mice with choline or TMAO promoted atherosclerosis formation, which was reduced by short-term broad-spectrum antibiotic, and in germ-free mice [58]. These findings suggest that circulating levels of microbiota-producing TMAO are important risk factors for the pathogenesis of CVD. Reversely, silencing of TMAO-producing enzyme FMO3 decreased circulating TMAO levels and attenuated atherosclerosis through enhancing basal metabolism and activating macrophage reverse cholesterol transport [59, 60]. It is reported that resveratrol, a natural phytoalexin, could attenuate TMAO-induced atherosclerosis by remodeling microbiota in ApoE−/− mice [58]. Moreover, a structural analog of choline, 3,3-dimethyl-1-butanol (DMB), and probiotic strains Lactobacillus plantarum ZDY04 are reported to inhibit TMAO production, and thus suppress atherosclerotic lesion development in mice fed a high-choline or L-carnitine diet [61, 62]. These studies indicated that pharmacological elimination or reduction of plasma TMAO is a potential therapeutic approach to reduce atherosclerotic cardiovascular events. Given the significant roles of TMAO-promoting CVD, a number of studies have been attempted to determine the underlying mechanisms (Fig. 8.2). Increasing intake of a TMAO precursor, choline, also accelerated the development of atherosclerosis. In choline-fed ApoE−/− mice, aortic macrophage number and positive scavenger receptor CD36 area as well as foam cell formation within aortic lesions were markedly increased, but not when intestinal microflora and TMAO production were suppressed by antibiotic treatment [8]. These results further support that gut microbial-generated TMAO accelerates atherosclerosis via inducing foam cell formation. However, in vitro study showed that different concentrations of TMAO treatment had no effect on acetylated LDL cholesterol loading or cholesterol efflux, and thus did not impact foam cell formation [63]. Therefore, the effect of gut microbial-producing TMAO on foam cell formation needs to be further investigated. TMAO can also contribute to atherosclerosis by inhibiting reverse cholesterol transport [12]. This mechanism may refer to changes in activity of cholesterol transporters in macrophages [12]. In addition, mRNA levels of key enzymes involved in BAs synthesis and transport in liver tissues were significantly reduced in mice supplemented with dietary TMAO [12], indicating that the pro-atherosclerotic role of TMAO is also associated with the alterations in BAs metabolism. TMAO was also shown to induce several atherosclerosis-promoting

8 Gut Microbiota and Heart, Vascular Injury

115

inflammatory responses in primary human aortic endothelial cells and vascular smooth muscle cells through activation of NF-jB or TXNIP/NLRP3 inflammasome [64–66]. These findings suggested that TMAO promoted inflammation-mediated atherosclerosis by inducing endothelial and monocyte dysfunction. Additionally, TMAO also increased stimulus-dependent platelet activation via inducing Ca2+ releases, resulting in the increased risks of thrombosis and plaque instability [57]. Although a number of studies have demonstrated an association between plasma levels of choline, TMAO, and atherosclerosis, the available data are still inconsistent. For example, increased TMAO level in circulation due to L-carnitine administration was surprisingly negatively correlated with aortic lesion size in ApoE−/− mice expressing human cholesteryl ester transfer protein [63]. Similarly, data show that digestion of dietary choline and betaine, as primary sources of TMAO, was not related with the progression of CVD [67, 68]. Therefore, the pro-atherosclerotic effects of TMAO as well as the validation of its therapeutic potential by targeting TMAO-producing bacteria or enzymes need to be further investigated in both animal and human models.

Dietary phosphatidylcholine /L-carnitine

Gut microbiota TMA

CD36

TMAO

Reversed cholesterol transporters

Macrophage

Foam cell formation Fig. 8.2 TMAO and possible molecular pathways linked to CVD. Direct provision of TMA within the diet is phosphatidylcholine, choline, and L-carnitine. Activation of FXR induces FMO3 in the liver; this enzyme converts TMA to TMAO. TMAO induces foam cell formation via CD36 and suppresses reversed cholesterol transport by modulating the activity of cholesterol transporters in macrophages

116

8.4

C. Zeng and H. Tan

Gut Microbiota and Hypertension

Hypertension has been documented as the most important single risk factor for CVD [69]. Hypertension is associated with both genetic susceptibility and environmental factors. However, the precise cause of hypertension has not been elucidated to date [70]. As the role of gut microbiota in pathophysiology of a variety of disorders has been gradually recognized, many studies paid more attention to investigating relationships between gut microbiota and hypertension. It has been revealed that the germ-free (GF) mice, in which the intestinal bacteria are completely absent, showed relatively lower blood pressure (BP) compared to conventionally raised (CONVR) mice [71]. In addition, investigators have demonstrated that probiotics treatment can improve BP by meta-analyses in human clinical trials [72]. Especially, a recent study based on metagenomic analyses of the fecal samples of 41 healthy controls, 56 human subjects with pre-hypertension, and 99 patients with primary hypertension revealed a novel causal role of aberrant gut microbiota in contributing to the pathogenesis of hypertension. Elevated BP was observed to be transferrable through fecal transplantation from hypertensive human donors to germ-free mice, pointing out the significance of gut microbiota in pre-hypertension [73]. Further studies revealed that overgrowth of bacteria such as Prevotella, Klebsiella, and Streptococcus were abundantly distributed in hypertensive patients compared to healthy controls [73, 74]. In both spontaneously hypertensive rats and the chronic angiotensin II (AngII) infusion rat model, Yang et al. observed a significant decrease in microbial richness and diversity, while an increase in the ratio of Firmicutes/Bacteroidetes compared with controls [75]. Likewise, in AngII-induced hypertensive rats, compared with CONVR mice, GF mice have attenuated the BP increase and inhibited vascular dysfunction. Besides, transplant of dysbiotic cecal contents from obstructive sleep apnea (OSA)-induced hypertensive rats on high-fat diet into OSA recipient rats on normal chow diet resulted in hypertension [76]. These results may implicate a causal relationship between gut microbiota and hypertension, and suggest that manipulation of the microbiota may be a viable treatment for hypertension although the underlying mechanism needs to be further characterized.

8.4.1

SCFAs and Hypertension

A previous study of human populations living in China, Japan, and United Kingdom showed a direct association between urinary formate, a SCFA generated by microbial fermentation of dietary polysaccharides, and BP [77]. Moreover, by investigating the gut microbial composition of 60 hypertensive patients and 60 healthy controls, Yan et al. observed that F. prausnitzii and Roseburia (both R. intestinalis and R. hominis), as major SCFA producers in human colon, were abundantly distributed in healthy controls compared to hypertensive patients [74].

8 Gut Microbiota and Heart, Vascular Injury

117

A recent study found that high-fiber diet and acetate supplementation significantly decreased diastolic BP, cardiac fibrosis, and left ventricular hypertrophy compared with mineralocorticoid-excess mice fed a control diet [78]. These data demonstrated that gut microbiota-producing SCFAs in circulation played important role in hypertension. Fiber fermentation by gut microbiota generates SCFAs that are either absorbed by the gut or excreted in feces. In a study of 441 community-dwelling adults, investigators found that higher SCFA excretion was associated with evidence of gut dysbiosis and excess adiposity, and cardiometabolic risk factors [79]. These studies revealed that cardiometabolic dysregulation is due to less efficient SCFA absorption. The G protein-coupled receptors (GPCRs) pathways play a major role in SCFAs absorption and have been well documented to be associated with the progression of hypertension. At least four GPCRs including GPR41, GPR43, GPR109A, and Olfactory receptor 78 (Olfr78) have been reported to be regulated by SCFA [80]. Except GPR41 and Olfr78, other receptors in response to signals generated via gut microbes have not been characterized in hypertension [80]. Propionate produced an acute hypotensive response in mice through disruption of Olfr78 and GPR41 expression [81]. Interestingly, GPR41 contributed to the hypotensive effects of propionate, whereas Olfr78-mediated renin release to raise BP and to antagonize the hypotensive effects of propionate [81]. The fact that these receptors showed their apparently opposing effects on BP in response to SCFA stimulation indicated that the net effects of SCFAs may be complex as multiple pathways are involved in the mediation of SCFA signaling. Nevertheless, these findings revealed that SCFAs exert multiple beneficial effects on cardiovascular diseases (Fig. 8.3).

Fig. 8.3 SCFAs and possible molecular pathways linked to CVD. The dietary fiber is metabolized to SCFAs, such as acetate, butyrate, and propionate. These SCFAs activate GPR41 and Olfr78, and then inhibits CVD by decreasing blood pressure or stimulating hormone PYY and GLP-1 release

Fiber diet

Glycolysis Acetate

Propionic acid

Butyric acid

SCFAs

GPR41

Blood pressure GLP-1

PYY

CVD

118

8.4.2

C. Zeng and H. Tan

H2S and Hypertension

H2S modulates a variety of physiological processes including cell protection, blood vessel relaxation, angiogenesis, lowing of BP, and heart rate [82, 83]. Investigators found that the H2S donor increased portal blood levels of thiosulfate and sulfane sulfur, and decreased mean arterial blood pressure (MABP) in a dose-dependent manner in spontaneously hypertensive rats, which lasted much longer than 90 min, and action could be inhibited by an antibiotic, neomycin [84]. This study implicated that gut-producing H2S may contribute to the control of BP and may be one of the links between gut microbiota and hypertension. Although the number of studies has reported that H2S has capacity to improve hypertension by multiple mechanisms including suppression of oxidative stress (OS) or inflammation, interaction with NO and CO, and ion channel-mediated relaxation of vascular smooth muscle [85], there are no direct evidence revealing that the H2S derived from gut microtia influences hypertension.

8.5

Gut Microbiota and Diabetes

Diabetes mellitus (DM) is a group of chronic diseases characterized by hyperglycemia and major risk factor for the development of CVD with a higher incidence of myocardial infarction in patients with DM than those without [86]. One of the injuries arising from diabetes is injury to vasculature, which is classified as either small vascular injury (microvascular disease) or injury to the large blood vessels of the body (macrovascular disease). Diabetic retinopathy may be the most common microvascular complication of diabetes. Among the diabetic patients, aerobic bacterial conjunctival flora were detected more frequently in those with diabetic retinopathy compared to those without retinopathy [87]. Further study demonstrated that predominant conjunctival organisms in type 2 diabetes (T2D) patients were Staphylococcus epidermidis and Staphylococcus aureus [88]. These works suggest that microbiota dysbiosis was associated with diabetic retinopathy. Recently, Beli et al. reported that intermittent fasting prevents diabetic retinopathy by restructuring the gut microbiota toward species producing tauroursodeoxycholic acid and subsequent retinal protection by TGR5 activation [89]. Although this study suggests that gut-derived microbiota was associated with progression of diabetic retinopathy in mice, more evidences that gut microbiota contributed to diabetic retinopathy need to be provided. Recently, studies revealed the gut microbiota functions as an important environmental factor in the development of diabetes including type 1 diabetes (T1D) and T2D. T2D is characterized by either insufficient production of insulin required to maintain normoglycemia in the face of insulin resistance or an inability of the body to utilize insulin produced. In 2010, Larsen et al. reported that the ratios of Bacteroidetes to Firmicutes as well as the ratios of Bacteroides-Prevotella group to

8 Gut Microbiota and Heart, Vascular Injury

119

C. coccoides-E. rectale group correlated positively and significantly with plasma glucose concentration in patients with T2D [90]. In addition, by analyzing the gut microbial DNA from 345 Chinese individuals by metagenome-wide association study (MGWAS) and a two-stage MGWAS based on deep shotgun sequencing, researchers found that patients with T2D were characterized as a decrease in the abundance of some universal butyrate-producing bacteria (Clostridiales sp. SS3/4, Eubacterium rectale, Faecalibacterium prausnitzii, etc.) and an increase in various opportunistic pathogens, as well as an enrichment of other microbial functions conferring sulfate reduction and oxidative stress resistance, further emphasizing that changes of gut microbiotas might be useful markers for classifying T2D [91]. Therapy of treatment-naïve T2D with metformin was recently found to alter the gut microbiome, which contributed to the therapeutic effects of this drug [92]. Moreover, researchers used shotgun sequencing to identify the fecal metagenome of 145 European women with normal, impaired, or diabetic glucose control and revealed that the abundances of four Lactobacillus species was higher, while five Clostridium species were lower in gut of T2D patient individuals compared to healthy controls [93]. In an cohort of Japanese T2D patients, the counts of Clostridium coccoides, Atopobium, and Prevotella were significantly reduced, while total Lactobacillus were increased in fecal samples of diabetic patients [94]. In addition to fecal samples, this study further revealed that gut bacteria were also detected in blood with significantly higher rate in diabetic patients than in control subjects, and most of these bacteria were Gram-positive [94]. Although T1D is an autoimmune disease that targets pancreatic islet beta cells, epigenetic and environmental factors have been shown to play an important role in this disease [95]. Patients with T1D showed a less diverse and less stable gut microbiome compared to healthy controls and gut dysbiosis have been observed in the patients [96, 97]. Abundance of butyrate-producing bacterium was decreased in children with T1D [98]. These evidences show that gut microbiota plays a significant role in the development for diabetes and also hinted that gut microbiota was related to complications of diabetes. More experiments need to be conducted whether gut microbiota contributes to microvascular and macrovascular complications of diabetes such as diabetic cardiomyopathy.

8.5.1

BAs and Diabetes

Over the past decades, a few studies have described alterations of the bile acid pool in T2D patients and animal models. One study showed an increased BA pool size in patients with uncontrolled T2D [99]. Subsequent studies demonstrated that the pool of the secondary bile acid DCA was elevated, whereas the CDCA pool decreased [100]. In addition, upregulation of BAs pool size by CYP7A1 overexpression improved insulin resistance and hepatic steatosis in high-fat-diet (HFD)-induced obesity mice [101]. These results indicate that modulation of BAs pools might be a potential therapeutic strategy for treating metabolic disorders.

120

C. Zeng and H. Tan

Indeed, BA sequestrants have been used in the treatment of T2D patients. In patients not adequately controlled by common antidiabetic therapeutics, the administration of colesevelam, a BA sequestrant, increased BA synthesis, particularly of CA, and decreased plasma glucose and hemoglobin by 0.7% in diabetics [102]. Schwartz et al. demonstrated that BA sequestrant may improve glucose control by improving whole-body insulin sensitivity in T2D patients [103]. Similar effect of BA sequestrant was confirmed in diet-induced obese mice [104]. In addition, Shang et al. reported that colesevelam induced the release of GLP-1 and thus improved insulin release in a diet-induced obesity rat model [105]. In addition to BA sequestrant, treatment with the semisynthetic BA derivative obeticholic acid (OCA), a high-affinity ligand of FXR, was able to improve insulin sensitivity in patients with nonalcoholic steatohepatitis (NASH) and T2D [106]. However, a multicentre, double-blind, randomized, placebo-controlled, phase IIIa study with OCA (FLINT) in 283 patients with NASH, with and without T2D, showed that insulin sensitivity and lipid profiles were deteriorated [107]. Therefore, the definite role of BAs in diabetes needs to be further investigated in clinical trials. A recent study further found that higher plasma bile acid levels improved glucose tolerance and high cholesterol levels after ileal interposition surgery in diet-induced obese rats [108]. Similar results that increased plasma bile acid levels were inversely correlated with 2-h post-meal glucose, and peak glucagon-like peptide-1 (GLP-1) was observed in patients after gastric bypass surgery [109].

8.5.2

SCFAs and Diabetes

In The Environmental Determinants of Diabetes in the Young (TEDDY) study, by analyzing microbiomes in children who were from 3 months to up to 5 years of age, the expression of microbial genes involved in the biosynthesis of SCFAs was lower in children who developed T1D than in those who did not [110]. In addition to T1D, Zhao et al. reported that giving high-fiber diets to patients with T2D for 12 weeks significantly increased acetate- and butyrate-producing microbial subpopulations and decreased hemoglobin A1c levels and improved glucose tolerance [23]. Sanna et al. further expanded this field by using bidirectional mendelian randomization (MR) analyses to assess causality, that is, the relationship between host genetics on microbial expression of SCFAs in the gut and its subsequent impact on glycemic indices. They found that host-genetic-driven increase in gut production of the SCFA butyrate was associated with improved insulin response by an oral glucose tolerance test, whereas abnormalities in the production or absorption of another SCFA, propionate, were causally related to an increased risk of T2D [111]. These works together indicated that gut microbiota-producing SCFAs are related closely with progression of diabetes. The ability to secrete insulin depends on the functioning and mass of pancreatic b-cells. SCFA receptors GPR41 and GPR43 were expressed in pancreatic b-cells of mice and humans [112, 113]. GPR43-knockout mice shown a marginally

8 Gut Microbiota and Heart, Vascular Injury

121

significant defect in insulin secretion during hyperglycemic clamps [114], demonstrating that SCFAs increase glucose-stimulated insulin secretion via GPR43. Consistent with this finding, GPR43 knockout was related with b-cell dysfunction and increased b-cell mass in obesity mice [112]. Moreover, SCFA-activated GPR41 signaling have the capacity to mediating pancreatic b cell insulin secretion in fed and fasting states, as GPR41 knockout or overexpression in mice resulted in impairments in glucose control [113]. In a human study, inulin-propionate ester delivered in colon was associated with improved b-cell function that was independent of changes in GLP-1 levels [115]. Further in vitro study showed propionate promoted glucose-stimulated insulin secretion and maintained b cell mass by inhibiting cell apoptosis [115]. These results all revealed that SCFAs-GPR signaling has important role on regulation of insulin secretion and blood glucose homeostasis. With respect to glucose metabolism, SCFAs suppress appetite by increasing the release of satiety hormones and stimulating vagal afferent chemoreceptors, increasing energy expenditure by upregulating thermogenesisrelated proteins in the liver and adipose tissue [1]. Given that SFCAs were produced from the fermentation of these indigestible carbohydrates such as fiber, a diet strategy was to enhance the dietary fiber to colonic microorganisms for preventing and improving metabolic disorders including diabetes [23].

8.5.3

H2S and Diabetes

H2S is a major product of protein fermentation via gut microbiota. The potential role of H2S in the pathogenesis of diabetes was somewhat conflicted. Study in vitro from Yang and colleagues has demonstrated that H2S treatment led to apoptosis in rat insulinoma INS-1E cells, while pharmacological inhibition of CSE, a key enzyme of H2S production, prevents streptozotocin-induced INS-1E cell death [116]. Consistent with in vitro data, excess H2S suppressed pancreatic islet function, and therefore contributed to T2D progression in vivo [116, 117]. Moreover, H2S impaired glucose uptake and glycogen storage in hepatocytes [118]. These results indicated that H2S has pro-diabetic role. In contrast, there was some evidence showing that H2S has an antidiabetic role. For example, H2S has a protective role in mouse islets and in MIN6 cells exposed to high glucose or fatty acids by downregulating thioredoxin-interacting protein expression [119, 120]. This role of H2S also was confirmed to prevent the onset of T2D in mice study [119]. Especially, data from humans showed that patients with T2D have low plasma concentrations of H2S compared with healthy controls [121]. Although several studies have begun to characterize the potential role of H2S in the pathogenesis of T2D, subsequent studies need to be conducted to bring clarity and mechanistic insight into H2S anti- or pro-diabetic role and show whether H2S generated from gut microbiota took part in the progression of diabetes.

122

8.5.4

C. Zeng and H. Tan

TMAO and Diabetes

Some evidences show that TMAO may affect glucose metabolism and high TMAO may be closely related to diabetes. Previous studies have suggested that plasma TMAO were increased [122] and remained predictive of both major adverse cardiac events and mortality risks in T2D patients [123]. A study in HFD-fed mice showed that dietary TMAO exacerbated impaired glucose tolerance, which may be involved in glycogen synthesis, gluconeogenesis, and glucose transport in liver [124]. In contrast, Mcentyre et al. found that diabetic patients treated with metformin showed decreased glucose but increased plasma TMAO compared to untreated ones [125]. Another study showed that plasma TMAO concentration was high variable in obese diabetic patients, which suggested TMAO may be a low predictive or diagnostic value [125]. In animal models, the level of hepatic TMAO and choline was decreased in diabetic mice, while administration of TMA attenuated endoplasmic reticulum stress in STZ-induced diabetic rats [126], which implies the protective role of TMAO in diabetes [127]. Given above contradictional results, precise evidences need to be demonstrated the role of TMAO in diabetes and whether gut microbiota regulated progression of diabetes by its metabolites TMA/TMAO. All results demonstrated that gut microbiota is intimately related to progression of diabetes. Given the pathologic hallmark of diabetes involves the vasculature leading to both microvascular and macrovascular complications, and chronicity of hyperglycemia is associated with long-term damage and failure of various organ systems mainly affecting the eyes, kidneys, and the heart, there is a great possibility that gut microbiota regulated cardiovascular complications of diabetes, which need to be further studied in vivo and in vitro.

8.6

Gut Microbiota and Obesity

Obesity is becoming a global epidemic in both children and adults. It is associated with increased risk of cardiovascular disease and premature death. The Bogalusa and Pathobiological Determinants of Atherosclerosis in Youth pathology studies have shown a strong association between overweight and the increased presence of atherosclerotic lesions in both the aorta and coronary arteries [128]. Obesity is quite important in the pathogenesis of hypertension. Although the mechanism of this relationship is not completely understood, epidemiologic studies also show a strong relationship between obesity and other cardiovascular diseases for both adults and children. Thus, once gut microbiota affected obesity, there appeared a strong possibility that gut microbiota-mediated obesity contributed cardiovascular complications of obesity. Indeed, the increase in the prevalence of obesity in many parts of the world is clearly driven by changes in exposure to environmental factors [129]. Recently, many studies raise the possibility that the gut microbiota plays a

8 Gut Microbiota and Heart, Vascular Injury

123

important role in regulating host energy metabolism and may contribute toward the development of obesity. In 1983, Wostmann et al. first observed that adult male GF rats need more energy to maintain their body mass than conventional ones [130]. Recent studies showed that GF mice are leaner than their CONVR counterparts and do not trigger diet-induced obesity [131]. Colonization of GF mice with a normal microbiota increases the amount of body fat and inhibits insulin sensitivity [132], suggesting that gut microbiota may be associated with the regulation of body fat. Indeed, sequence-based studies in rodents have showed an increase in ratio of Firmicutes and Bacteroidetes and a decrease in the diversity of the microbiota in the gut [133]. Some studies report a similar increase in the ratio of Firmicutes and Bacteroides, as well as a decrease in the gut biodiversity of obese humans, but there are studies showing contradictory results in obese compared with lean humans [134]. In addition, beneficial microbiotas including Bifidobacteriumbv and anti-inflammatory Faecalibacterium, and butyrate-producing Ruminococcaceae were significantly decreased, while Bacillus and potential opportunistic pathogens were increased in patients with obesity compare to healthy controls [135]. Especially, at the phylum level, Proteobacteria have potential pathogenic features [136], and its abundance was higher in patients with obesity compared with healthy controls or patients with overweight [135]. Therefore, gut dysbiosis contributes to obesity while prebiotic agents, such as Ganoderma lucidum, have potent to prevent obesity [137].

8.6.1

SCFAs and Obesity

By analyzing 16S ribosomal RNA sequencing from 205 women, data have revealed that the abundance of the SCFA-producing bacterial families Odoribacteraceae and Clostridiaceae were inversely correlated with BMI in obese pregnant women [138]. Indeed, SCFAs activated GPR43 and thus decreased release of inflammatory cytokines in the intestines [139], which may also increase hypothalamic sensitivity to leptin, an important regulator of obesity [140, 141]. In addition, SCFAs have an important role in sustaining energy homeostasis. Butyrate has been reported to inhibit histone deacetylases (HDACs) [142] and therefore induce histone hypermethylation of genes involved in fatty acid oxidation [143]. Similarly, there are some effects of each SCFA on various gut peptides which would have a net effect of slowing digestion and nutrient intestinal transit, promoting satiety, and increasing plasma insulin. G protein mediated secretion of the satiety-stimulating hormone peptide YY (PYY) and GLP-1 as well as rates of lipolysis and lipogenesis in fat cells were upregulated in the gut after butyric acid or propionate administration [144, 145]. Butyrate and acetate have been reported to improve diet-induced obesity; propionate has capacity to decrease food intake while increase locomotor activity [146]. Based on these rodent data, colonic administration of SCFA may also have beneficial effects on human substrate and energy metabolism. However, there is a conflicting literature; Perry et al. revealed that an increased acetate

124

C. Zeng and H. Tan

turnover promotes the development of obesity and insulin resistance [147], indicating that each SCFA may have a diverse effect on pathogenesis of obesity. Human data indicating the in vivo effect of SCFA on obesity are scarce. The most abundant SCFA (acetate) via acute infusions in the distal of the colon enhanced fat oxidation and circulating levels of the PYY in overweight men, indicating an improved metabolic profile [148]. Moreover, in a randomized crossover trial that aimed to investigate the role of gut-derived SCFA by rectally administered physiologically relevant SCFA mixtures, Emanuel et al. found that SCFA mixtures upregulated not only fasting fat oxidation and resting energy expenditure but also fasting and postprandial PYY concentrations [149], providing evidence that SCFAs have important clinical implications on food intake regulation and control of body weight.

8.6.2

BAs and Obesity

BAs have been reported to affect the regulation of energy expenditure and development of obesity by action of its receptors, TGR5 and FXR. Watanabe et al. demonstrated that dietary CA administration reversed HFD-induced obesity and fat accumulation by activating TGR5-mediated intracellular thyroid hormone and thus increased thermogenesis in brown adipose tissue [150]. Subsequent studies have confirmed that TGR5 increases energy expenditure of brown adipose tissue and muscle, and release of increasing glucagon-like peptide (GLP)-1 in intestinal L cells and a cells in the pancreas [151]. Similarly, INT-777, a specific TGR5 agonist, has been reported to ameliorate hepatic steatosis and adiposity and improve insulin sensitivity in mice with high-fat-diet-induced obesity [152]. Another receptor, FXR, was also expressed in intestinal L cells where it inhibits GLP-1 synthesis [153]. Liver-specific FXR/small heterodimer partner SHP double knockout improved glucose/fatty acid homeostasis and suppressed body weight gain and adiposity in aged mice, suggesting that intrahepatic FXR activation may also regulate whole-body energy [154]. In addition, hepatic FXR inhibited steatosis [155] while intestine-selective FXR contributed to obesity [156, 157], suggesting that the role of FXR on obesity seems to depend on its location. Thus, the precise role of FXR needs to be further investigated in vivo and in vitro. A human study demonstrated a promoting effect of bile acids on energy expenditure and showed that CDCA increased whole-body energy expenditure and activated brown adipose tissue in 12 healthy women [158]. Paradoxically, most studies found increased total bile acid levels in humans with obesity [159]. Gut microbiota plays an important role in progression of obesity. Obesity is associated with vascular remodeling, inflammation, oxidative stress, insulin resistance, and endothelial dysfunction, which all lead to maladaptive structural and functional alterations of the heart such as left ventricle remodeling and diastolic dysfunction, sometimes termed obesity-associated cardiomyopathy, and through a period of asymptomatic subclinical disease may ultimately result in clinically overt

8 Gut Microbiota and Heart, Vascular Injury

125

CVD. Therefore, there are reasons to speculate that gut microbiota may be associated with cardiovascular complication of obesity even though detail evidence should be supplied.

8.7

Gut Microbiota and Heart Failure

Heart failure (HF) is growing to a modern epidemic and a main cause of mortality and morbidity worldwide despite advances in therapy; it still carries an ominous prognosis and a significant socioeconomic burden. The gut has also been implicated in the progression of heart failure because a decreased cardiac output and/or splanchnic venous impaired perfusion to the intestines and thus lead to intestinal barrier dysfunction and increased circulating endotoxins that can contribute to the underlying inflammation [160]. In addition to intestinal wall ischemia, intestinal wall edema caused by elevated systemic congestion with HF also increased bacterial translocation and circulating endotoxins. Once endotoxins such as LPS enter the mesenteric lymph nodes or the systemic circulation through the intestine, they have potentially leading to further HF exacerbations [161]. Indeed, higher concentrations of bacterial DNA in circulation have been detected in patients with CVD compared with healthy subjects [162]. The concentration of bacterial DNA has reported to be considerable impact on the onset of cardiovascular events [163]. Recently, a study in human showed that intestinal overgrowth of pathogenic bacteria and Candida species was associated with intestinal permeability and clinical disease severity in patient with chronic heart failure (CHF) [164]. In another study, HF patients had a significantly decreased diversity of gut microbiota and a depletion of key intestinal bacterial groups [165]. These data indicated that improvement of intestinal barrier function may be an effective therapy for HF. As mentioned above, there were clear evidences showing that TMAO levels have been linked to poor prognosis in patients with HF [8, 13, 166, 167]. In transverse aortic constriction(TAC)-induced HF mice, TMAO- or cholinesupplemented diets induced myocardial fibrosis and accelerated adverse ventricular remodeling compared with the control diet [168]. Recently, TMAO treatment induced cardiac hypertrophy and cardiac fibrosis, which was inhibited by pharmacological inhibition of Smad3 in SD rats [169]. In contrast, reducing TMAO synthesis by antibiotics improved TAC-induced cardiac fibrosis [169]. This role of TMAO on pro-fibrosis by transforming growth factor-b Smad3 pathway is also observed in kidney tissue of mice fed with chlorine or TMAO [170]. These animal model studies suggest that beyond association studies and adverse prognosis data in humans, the TMAO pathway may directly induce adverse ventricular remodeling and HF phenotype. Arial fibrillation, a common kind of arrhythmia, is associated with an increased risk of worsening HF [171]. A clinical study demonstrated that TMAO predicts the risk of atrial fibrillation and positively correlates with the occurred atrial fibrillation [172]. Further study in animal model revealed that TMAO facilitated the

126

C. Zeng and H. Tan

progression of atrial fibrillation by activation of cardiac autonomic nervous system [173]. In addition to microbiota-derived TMAO, increased circulation LPS was derived from gut microbiota and associated with major adverse cardiovascular event occurrence in patients with atrial fibrillation [174]. The precise relationship between gut microbiome and atrial fibrillation needs further study, both clinically and basically. Myocardial infarction is one of the most common causes leading to HF. Lam et al. demonstrated that antibiotic vancomycin altered abundance of individual groups of intestinal microbiota and resulted in reduced myocardial infarcts and improved recovery of postischemic mechanical function compared to untreated controls by changing circulating leptin levels in Dahl S rats [175]. Other low-molecular-weight metabolites, such as phenylalanine, tryptophan, and tyrosine, produced by intestinal microbiota and carried to the systemic circulation, also affect the severity of myocardial infarction [176]. These works suggested that a direct link between the gut microbiota and the severity of myocardial infarction in rats. Further study reveals that the richness of gut microbiota and the abundance of Synergistetes phylum, Spirochaetes phylum, Lachnospiraceae family, Syntrophomonadaceae family, and Tissierella Soehngenia genus was higher and associated with intestinal barrier impairment in acute myocardial infarction rat [177]. In myocardial infarction patients, the researchers found that after ST-segment elevations, the abundance and diversity of circulatory microbiota were increased and more than 12% of the blood bacteria were dominated by intestinal microbiota and contributed to cardiovascular events post-myocardial infarction [178]. Therefore, the intervention of gut microbial composition to improve myocardial infarction may be a new method of myocardial infarction treatment. Probiotic administration (Lactobacillus rhamnosus GR-1) inhibited left ventricular hypertrophy and improved left ventricular ejection fraction and fractional shortening in myocardial infarction rats [179]. Recently, Tang et al. found that dietary SCFA supplementation can improve survival of antibiotic-treated mice after myocardial infarction and suggested that microbiota-derived SCFAs play an important role in maintaining repair capacity after myocardial infarction [180]. In addition, HIV infection is associated with increased risk of coronary heart disease beyond that explained by traditional risk factors. Several studies have reported that altered composition of microbiota in HIV infection is associated with increased systemic inflammation, suggesting a potential association between altered microbiota and coronary heart disease in HIV infection. Recently, Haissman et al. reported that although there was no difference in plasma TMAO when comparing HIV-infected persons and uninfected controls, TMAO was elevated in HIV-infected persons with myocardial perfusion defects [181].

8 Gut Microbiota and Heart, Vascular Injury

8.8

127

Therapeutic Potential of Gut Microbiota in CVD

Although many types of medicines are available in the clinic to treat CVD, currently, it is still the leading cause of death worldwide. In recent years, the gut microbiota was regarded as important regulator in the development of CVD and aroused interest in the development of microbiota-targeted therapies by regulating the composition of gut microbiota community and/or gut microbial metabolites (Fig. 8.4).

8.8.1

Antibiotic, Probiotic, and Prebiotic Treatment in CVD

Antibiotic treatment was usually regarded as a experimental strategy to study how the gut microbiome influences various disease states in animal models. Recently, several randomized, placebo-controlled trials have investigated whether antibiotic

2. Antibiotic treatment, probiotic and prebiotic

Antibiotic

1. Fiber-rich diet

Lactobacillus spp. etc Reshape microbiota

4. Inhibitors targeting signalings involving gut microbial metabolites

3. Fecal microbiota transplantation

DMB etc Donor fecal sample

Fig. 8.4 Targeting gut microbiota and related metabolites are promising strategies for the treatment of CVD. Some strategies have therapeutic actions on CVD by reshaping the gut microbiota. The most frequently used approaches to manipulate the gut microbiota include antibiotic treatment, probiotic, prebiotic, diet intervention, fecal microbiota transplantation, and inhibitors targeting signaling involving in production of gut microbial metabolites

128

C. Zeng and H. Tan

treatment has beneficial effects on CVD. However, data from these clinic trials showed no benefit in reducing mortality or cardiovascular events in patients with coronary artery diseases [31]. The disappointing results are attributed to short treatment duration, bacterial resistance, or unselective targeting of protective bacteria. However, long-term treatment with antibiotics might not be sensible because antibiotic administration during early life is associated with childhood obesity in humans and mice [182, 183]. Therefore, antibiotic treatment should focus on preservation of the beneficial microbiota and duration of treatment. Probiotics are beneficial live microorganisms to improve health by re-establishing an appropriate balance in the intestinal microenvironment. Focusing on the variable hypocholesterolemic effects of probiotics, scientists have researched the mechanisms by which probiotics impact hypercholesterolemia in vitro and in vivo. Numerous studies have revealed the effect of probiotics on decreasing serum LDL cholesterol and total cholesterol levels, which are main risk of CVD [184]. Further research studies demonstrated that fermented milk containing wild-type Lactobacillus strains has hypocholesterolemic effects [185]. In 2000, a double-blind randomized study of 70 overweight and obese adults showed that consumption of probiotic yogurt containing S. thermophiles and Enterococcus faecium for 8 weeks significantly reduced 8.4% of LDL cholesterol which was associated with a reduced risk of CVD by 20–30% [186]. In addition, Kekkonen et al. further found that probiotic Lactobacillus rhamnosus GG could decrease sphingomyelins [187], a major predictor of heart disease. In a recent study, Lactobacillus reuteri was reported to induce insulin secretion in obese glucose-intolerant patients [188]. A meta-analysis of nine randomized, controlled trials found that daily consumption of  1011 CFU probiotic for 8 weeks significantly decreased BP [72]. Studies in experimental animals have also revealed beneficial effects of probiotic administration on CVD. These probiotics include VSL#3, L. rhamnosus GG, L. acidophilus ATCC 4358, and A. muciniphila, etc. [189]. Another strategy to regulate intestinal microbiota is the use of prebiotics. Prebiotics are nondigestible carbohydrates that beneficially affect host health by selectively stimulating the growth and/or activity of the health-promoting gut microbiota [190, 191]. Typical prebiotics are food indigestible fibers such as oligosaccharides or complex saccharides [192]. Oligofructose supplementation for 3 months has been reported to improve glucose tolerance and lose weight in patient with obesity [193]. Although administration of antibiotics or prebiotics improves gut permeability and glucose intolerance in genetically obese and diet-induced leptin-resistant mice [194], the prebiotic treatment did not show similar strong effects in women with obesity. Therefore, prebiotic administration for treating CVD needs to be further confirmed, particularly through clinical trial.

8 Gut Microbiota and Heart, Vascular Injury

8.8.2

129

Fecal Microbiota Transplantation in CVD Treatment

Fecal microbiota transplantation (FMT) is a therapeutic intervention in which the intestinal microbiota is transferred from a healthy donor to the patient, with the goal to displace intestinal pathogens by introducing or restoring a stable microbial community in the gut. In the past few decades, FMT was reportedly used as a therapeutic approach for antibiotic resistance in human subject infected with Clostridium difficile [195]. Recently, the application of FMT has been broadened and tested in clinical trials to improve cardiometabolic disorders [196]. Vrieze et al. demonstrated that insulin sensitivity in overweight patients with metabolic syndrome was improved about 176% after being transplanted with microbiota from lean healthy controls for 6 weeks [196]. The FMT increases overall gut microbial richness and especially abundance of butyrate-producing bacteria in treated patients. However, improvement of insulin sensitivity by FMT was unabiding as no difference was observed at 18 weeks. However, clinical application of FMT is currently limited because the treatment may induce transfer of endotoxins or infectious agents from the donors to the treated patients, thus increasing the risk of infection following FMT therapy [197]. Thus, instead of fecal contents, transplantation of only a defined group of bacteria may be advisable for FMT. Nevertheless, future studies are needed to optimize FMT in cardiometabolic disorders.

8.8.3

Reshaping the Population of Gut Bacteria by Dietary Intervention and Other Therapies

The dietary interventions could be a major relational factor in rapidly shaping gut microbiota and, hence, a dietary approach to nutritional interventions had been proved to be an effective strategy for treating CVD. Fiber-rich diets have been reported to decrease BP, cardiac hypertrophy, and fibrosis through increasing richness of acetate-producing microbiota [78]. In addition, Xiao et al. demonstrated that dietary intervention with whole grains, traditional Chinese medicinal foods, and prebiotics led to improvement in insulin sensitivity, lipid profiles, and BP which was accompanied with the reduction of opportunistic pathogen of the Enterobacteriaceae and Desulfovibrionaceae and increment in the family gut barrier-protecting bacteria Bifidobacteriaceae [198]. Modulation of gut microbiota composition through nutritional intervention represents a promising therapeutic approach. Furthermore, the metabolites of gut bacteria and their associated enzymes were also recognized as potential therapeutic targets of CVD. For example, a structural analog of choline, DMB, inhibited choline diet-enhanced endogenous macrophage foam cell formation and atherosclerotic lesion development in ApoE−/ − mice [61]. Therefore, additional research to delineate the precise mechanisms by

130

C. Zeng and H. Tan

which gut dysbiosis increases CVD risk and the novel therapeutic strategy that restores the homeostasis of gut microbiome in the host are likely to emerge.

8.9

Conclusion

The pathophysiology of CVD is complex, multifactorial, and, in many respects, poorly understood. Nevertheless, growing body of evidence suggests that gut microbiota plays a significant role in the development of CVD. Based on metabonomics and metagenomic sequencing analysis, scientists have demonstrated

Table 8.1 Gut metabolites and CVD Gut metabolites

Gut metabolite-related receptor

CVD

Detrimental or beneficial function

References

BAs

FXR FXR FXR FXR FXR TGR5 TGR5

Atherosclerosis Atherosclerosis T2D Obesity Obesity Atherosclerosis Diabetic retinopathy Obesity Atherosclerosis Atherosclerosis Atherosclerosis Diabetes Diabetes Heart failure Atrial fibrillation Hypertension T2D hypertension T2D Obesity Myocardial infarction Hypertension Diabetes Diabetes

Beneficial Detrimental Beneficial Beneficial Detrimental Beneficial Beneficial

[42–45] [46, 47] [106] [155] [156, 157] [49, 51] [89]

Beneficial Detrimental Detrimental Beneficial Detrimental Beneficial Detrimental Detrimental

[149] [54] [8, 57–62] [63] [124, 125] [126] [168] [173]

Beneficial Beneficial Beneficial Beneficial Beneficial Beneficial

[81] [113, 115] [102] [112, 114] [139] [175]

Beneficial Detrimental Beneficial

[82–84] [116–118] [119, 120]

TGR5 PXR TMA/ TMAO

SCFAs

H2S

GPR41 GPR41 Olfr78 GPR43 GPR43

8 Gut Microbiota and Heart, Vascular Injury

131

that gut microbiota impacts host health and disease by production of bacterial metabolites. The detrimental metabolites, such as TMA/TMAO, may sustain and aggravate an ongoing process of CVD, while some beneficial metabolites, such as SFCAs, may slow progression of CVD (Table 8.1). There is great promise of finding new approaches to treat CVD by using gut microbial metabolites such as supplement with beneficial gut microbiota metabolites or blocking the production of detrimental microbial metabolites. Although intervention of metabolites has been investigated in animal models, well-designed large-scale clinical studies needed to validate the feasibility of this approach.

8.10

Sources of Funding

This work was supported by the National Natural Science Foundation of China (No. 81873514 and 81570394).

References 1. Canfora EE, Meex RCR, Venema K, Blaak EE (2019) Gut microbial metabolites in obesity, NAFLD and T2DM. Nat Rev Endocrinol 15(5):261–273 2. Qin J, Li R, Raes J, Arumugam M, Burgdorf KS, Manichanh C et al (2010) A human gut microbial gene catalogue established by metagenomic sequencing. Nature 464(7285):59–65 3. David LA, Maurice CF, Carmody RN, Gootenberg DB, Button JE, Wolfe BE et al (2014) Diet rapidly and reproducibly alters the human gut microbiome. Nature 505(7484):559–563 4. Ivanov II, Frutos Rde L, Manel N, Yoshinaga K, Rifkin DB, Sartor RB et al (2008) Specific microbiota direct the differentiation of IL-17-producing T-helper cells in the mucosa of the small intestine. Cell Host Microbe 4(4):337–349 5. Min YW, Rhee PL (2015) The role of microbiota on the gut immunology. Clin Ther 37 (5):968–975 6. Sommer F, Backhed F (2013) The gut microbiota–masters of host development and physiology. Nat Rev Microbiol 11(4):227–238 7. Gopalakrishnan V, Helmink BA, Spencer CN, Reuben A, Wargo JA (2018) The influence of the gut microbiome on cancer, immunity, and cancer immunotherapy. Cancer Cell 33 (4):570–580 8. Wang Z, Klipfell E, Bennett BJ, Koeth R, Levison BS, Dugar B et al (2011) Gut flora metabolism of phosphatidylcholine promotes cardiovascular disease. Nature 472(7341): 57–63 9. Collaborators GBDCoD (2017) Global, regional, and national age-sex specific mortality for 264 causes of death, 1980–2016: a systematic analysis for the Global Burden of Disease Study 2016. Lancet 390(10100):1151-1210 10. Pedersen HK, Gudmundsdottir V, Nielsen HB, Hyotylainen T, Nielsen T, Jensen BA et al (2016) Human gut microbes impact host serum metabolome and insulin sensitivity. Nature 535(7612):376–381 11. Henao-Mejia J, Elinav E, Jin C, Hao L, Mehal WZ, Strowig T et al (2012) Inflammasome-mediated dysbiosis regulates progression of NAFLD and obesity. Nature 482(7384):179–185

132

C. Zeng and H. Tan

12. Koeth RA, Wang Z, Levison BS, Buffa JA, Org E, Sheehy BT et al (2013) Intestinal microbiota metabolism of L-carnitine, a nutrient in red meat, promotes atherosclerosis. Nat Med 19(5):576–585 13. Wang Z, Tang WH, Buffa JA, Fu X, Britt EB, Koeth RA et al (2014) Prognostic value of choline and betaine depends on intestinal microbiota-generated metabolite trimethylamineN-oxide. Eur Heart J 35(14):904–910 14. Schwarz M, Lund EG, Setchell KD, Kayden HJ, Zerwekh JE, Bjorkhem I et al (1996) Disruption of cholesterol 7 alpha-hydroxylase gene in mice. II. Bile acid deficiency is overcome by induction of oxysterol 7 alpha-hydroxylase. J Biol Chem 271(30):18024– 18031 15. Molinaro A, Wahlstrom A, Marschall HU (2018) Role of bile acids in metabolic control. Trends Endocrin Met 29(1):31–41 16. Li-Hawkins J, Gafvels M, Olin M, Lund EG, Andersson U, Schuster G et al (2002) Cholic acid mediates negative feedback regulation of bile acid synthesis in mice. J Clin Investig 110 (8):1191–1200 17. Wahlstrom A, Sayin SI, Marschall HU, Backhed F (2016) Intestinal crosstalk between bile acids and microbiota and its impact on host metabolism. Cell Metab 24(1):41–50 18. Macfarlane GT, Macfarlane S (2012) Bacteria, colonic fermentation, and gastrointestinal health. J AOAC Int 95(1):50–60 19. Cummings JH, Pomare EW, Branch WJ, Naylor CP, Macfarlane GT (1987) Short chain fatty acids in human large intestine, portal, hepatic and venous blood. Gut 28(10):1221–1227 20. Rey FE, Faith JJ, Bain J, Muehlbauer MJ, Stevens RD, Newgard CB et al (2010) Dissecting the in vivo metabolic potential of two human gut acetogens. J Biol Chem 285(29):22082– 22090 21. Louis P, Flint HJ (2017) Formation of propionate and butyrate by the human colonic microbiota. Environ Microbiol 19(1):29–41 22. Duncan SH, Barcenilla A, Stewart CS, Pryde SE, Flint HJ (2002) Acetate utilization and butyryl coenzyme A (CoA): acetate-CoA transferase in butyrate-producing bacteria from the human large intestine. Appl Environ Microb 68(10):5186–5190 23. Zhao L, Zhang F, Ding X, Wu G, Lam YY, Wang X et al (2018) Gut bacteria selectively promoted by dietary fibers alleviate type 2 diabetes. Science 359(6380):1151–1156 24. Babaei-Karamshahlou M, Hooshmand B, Hajizadeh S, Mani AR (2012) The role of endogenous hydrogen sulfide in pathogenesis of chronotropic dysfunction in rats with cirrhosis. Eur J Pharmacol 696(1–3):130–135 25. Tomasova L, Konopelski P, Ufnal M (2016) Gut bacteria and hydrogen sulfide: the new old players in circulatory system homeostasis. Molecules 21(11) 26. Fiorucci S, Antonelli E, Distrutti E, Rizzo G, Mencarelli A, Orlandi S et al (2005) Inhibition of hydrogen sulfide generation contributes to gastric injury caused by anti-inflammatory nonsteroidal drugs. Gastroenterology 129(4):1210–1224 27. Drosos I, Tavridou A, Kolios G (2015) New aspects on the metabolic role of intestinal microbiota in the development of atherosclerosis. Metab Clin Exp 64(4):476–481 28. Karlsson FH, Fak F, Nookaew I, Tremaroli V, Fagerberg B, Petranovic D et al (2012) Symptomatic atherosclerosis is associated with an altered gut metagenome. Nat Commun 3:1245 29. Li J, Lin S, Vanhoutte PM, Woo CW, Xu A (2016) Akkermansia muciniphila protects against atherosclerosis by preventing metabolic endotoxemia-induced inflammation in ApoE−/− mice. Circulation 133(24):2434–2446 30. Stepankova R, Tonar Z, Bartova J, Nedorost L, Rossman P, Poledne R et al (2010) Absence of microbiota (germ-free conditions) accelerates the atherosclerosis in ApoE-deficient mice fed standard low cholesterol diet. J Atheroscler Thromb 17(8):796–804 31. Andraws R, Berger JS, Brown DL (2005) Effects of antibiotic therapy on outcomes of patients with coronary artery disease: a meta-analysis of randomized controlled trials. Jama 293(21):2641–2647

8 Gut Microbiota and Heart, Vascular Injury

133

32. Hayashi C, Viereck J, Hua N, Phinikaridou A, Madrigal AG, Gibson FC 3rd et al (2011) Porphyromonas gingivalis accelerates inflammatory atherosclerosis in the innominate artery of ApoE deficient mice. Atherosclerosis 215(1):52–59 33. Zhang T, Kurita-Ochiai T, Hashizume T, Du Y, Oguchi S, Yamamoto M (2010) Aggregatibacter actinomycetemcomitans accelerates atherosclerosis with an increase in atherogenic factors in spontaneously hyperlipidemic mice. FEMS Immunol Med Microbiol 59(2):143–151 34. Cani PD, Amar J, Iglesias MA, Poggi M, Knauf C, Bastelica D et al (2007) Metabolic endotoxemia initiates obesity and insulin resistance. Diabetes 56(7):1761–1772 35. Akira S, Uematsu S, Takeuchi O (2006) Pathogen recognition and innate immunity. Cell 124 (4):783–801 36. Ostos MA, Recalde D, Zakin MM, Scott-Algara D (2002) Implication of natural killer T cells in atherosclerosis development during a LPS-induced chronic inflammation. FEBS Lett 519(1–3):23–29 37. Yoshida N, Emoto T, Yamashita T, Watanabe H, Hayashi T, Tabata T et al (2018) Bacteroides vulgatus and bacteroides dorei reduce gut microbial lipopolysaccharide production and inhibit atherosclerosis. Circulation 138(22):2486–2498 38. Hylemon PB, Zhou H, Pandak WM, Ren S, Gil G, Dent P (2009) Bile acids as regulatory molecules. J Lipid Res 50(8):1509–1520 39. Thomas C, Pellicciari R, Pruzanski M, Auwerx J, Schoonjans K (2008) Targeting bile-acid signalling for metabolic diseases. Nat Rev Drug Discov 7(8):678–693 40. Jiang C, Xie C, Li F, Zhang L, Nichols RG, Krausz KW et al (2015) Intestinal farnesoid X receptor signaling promotes nonalcoholic fatty liver disease. J Clin Investig 125(1):386–402 41. Lambert G, Amar MJA, Guo G, Brewer HB, Gonzalez FJ, Sinal CJ (2003) The farnesoid X-receptor is an essential regulator of cholesterol homeostasis. J Biol Chem 278(4):2563– 2570 42. Miyazaki-Anzai S, Masuda M, Levi M, Keenan AL, Miyazaki M (2014) Dual activation of the bile acid nuclear receptor FXR and G-protein-coupled receptor TGR5 protects mice against atherosclerosis. PloS one 9(9):e108270 43. Hartman HB, Gardell SJ, Petucci CJ, Wang S, Krueger JA, Evans MJ (2009) Activation of farnesoid X receptor prevents atherosclerotic lesion formation in LDLR−/− and ApoE−/− mice. J Lipid Res 50(6):1090–1100 44. Mencarelli A, Renga B, Distrutti E, Fiorucci S (2009) Antiatherosclerotic effect of farnesoid X receptor. Am J Physiol Heart Circ Physiol 296(2):H272–281 45. Hanniman EA, Lambert G, McCarthy TC, Sinal CJ (2005) Loss of functional farnesoid X receptor increases atherosclerotic lesions in apolipoprotein E-deficient mice. J Lipid Res 46 (12):2595–2604 46. Guo GL, Santamarina-Fojo S, Akiyama TE, Amar MJ, Paigen BJ, Brewer B Jr et al (2006) Effects of FXR in foam-cell formation and atherosclerosis development. Biochim Biophys Acta 1761(12):1401–1409 47. Zhang Y, Wang X, Vales C, Lee FY, Lee H, Lusis AJ et al (2006) FXR deficiency causes reduced atherosclerosis in Ldlr−/− mice. Arterioscler Thrombosis Vasc Biol 26(10):2316– 2321 48. Bishop-Bailey D, Walsh DT, Warner TD (2004) Expression and activation of the farnesoid X receptor in the vasculature. Proc Natl Acad Sci USA 101(10):3668–3673 49. Kida T, Murata T, Hori M, Ozaki H (2009) Chronic stimulation of farnesoid X receptor impairs nitric oxide sensitivity of vascular smooth muscle. Am J Physiol Heart Circ Physiol 296(1):H195–201 50. Li J, Wilson A, Kuruba R, Zhang Q, Gao X, He F et al (2008) FXR-mediated regulation of eNOS expression in vascular endothelial cells. Cardiovasc Res 77(1):169–177 51. Pols TW, Nomura M, Harach T, Lo Sasso G, Oosterveer MH, Thomas C et al (2011) TGR5 activation inhibits atherosclerosis by reducing macrophage inflammation and lipid loading. Cell Metab 14(6):747–757

134

C. Zeng and H. Tan

52. Kida T, Tsubosaka Y, Hori M, Ozaki H, Murata T (2013) Bile acid receptor TGR5 agonism induces NO production and reduces monocyte adhesion in vascular endothelial cells. Arterioscler Thrombosis Vasc Biol 33(7):1663–1669 53. Staudinger JL, Goodwin B, Jones SA, Hawkins-Brown D, MacKenzie KI, LaTour A et al (2001) The nuclear receptor PXR is a lithocholic acid sensor that protects against liver toxicity. Proc Natl Acad Sci USA 98(6):3369–3374 54. Sui Y, Xu J, Rios-Pilier J, Zhou C (2011) Deficiency of PXR decreases atherosclerosis in ApoE-deficient mice. J Lipid Res 52(9):1652–1659 55. Zhou C, King N, Chen KY, Breslow JL (2009) Activation of PXR induces hypercholesterolemia in wild-type and accelerates atherosclerosis in ApoE deficient mice. J Lipid Res 50 (10):2004–2013 56. Chen J, Guo Y, Gui Y, Xu D (2018) Physical exercise, gut, gut microbiota, and atherosclerotic cardiovascular diseases. Lipids Health Dis 17(1):17 57. Zhu W, Gregory JC, Org E, Buffa JA, Gupta N, Wang Z et al (2016) Gut microbial metabolite TMAO enhances platelet hyperreactivity and thrombosis risk. Cell 165(1): 111–124 58. Chen ML, Yi L, Zhang Y, Zhou X, Ran L, Yang J et al (2016) Resveratrol attenuates trimethylamine-N-oxide (TMAO)-induced atherosclerosis by regulating TMAO synthesis and bile acid metabolism via remodeling of the gut microbiota. mBio 7(2):e02210–02215 59. Miao J, Ling AV, Manthena PV, Gearing ME, Graham MJ, Crooke RM et al (2015) Flavin-containing monooxygenase 3 as a potential player in diabetes-associated atherosclerosis. Nat Commun 6:6498 60. Shih DM, Wang Z, Lee R, Meng Y, Che N, Charugundla S et al (2015) Flavin containing monooxygenase 3 exerts broad effects on glucose and lipid metabolism and atherosclerosis. J Lipid Res 56(1):22–37 61. Wang Z, Roberts AB, Buffa JA, Levison BS, Zhu W, Org E et al (2015) Non-lethal inhibition of gut microbial trimethylamine production for the treatment of atherosclerosis. Cell 163(7):1585–1595 62. Qiu L, Tao X, Xiong H, Yu J, Wei H (2018) Lactobacillus plantarum ZDY04 exhibits a strain-specific property of lowering TMAO via the modulation of gut microbiota in mice. Food Funct 9(8):4299–4309 63. Collins HL, Drazul-Schrader D, Sulpizio AC, Koster PD, Williamson Y, Adelman SJ et al (2016) L-Carnitine intake and high trimethylamine N-oxide plasma levels correlate with low aortic lesions in ApoE(−/−) transgenic mice expressing CETP. Atherosclerosis 244:29–37 64. Sun X, Jiao X, Ma Y, Liu Y, Zhang L, He Y et al (2016) Trimethylamine N-oxide induces inflammation and endothelial dysfunction in human umbilical vein endothelial cells via activating ROS-TXNIP-NLRP3 inflammasome. Biochem Biophys Res Commun 481(1– 2):63–70 65. Seldin MM, Meng Y, Qi H, Zhu W, Wang Z, Hazen SL et al (2016) Trimethylamine N-oxide promotes vascular inflammation through signaling of mitogen-activated protein kinase and nuclear factor-kappaB. J Am Heart Assoc 5(2):e002767 66. Ma G, Pan B, Chen Y, Guo C, Zhao M, Zheng L et al (2017) Trimethylamine N-oxide in atherogenesis: impairing endothelial self-repair capacity and enhancing monocyte adhesion. Biosci Rep 37(2) 67. Bidulescu A, Chambless LE, Siega-Riz AM, Zeisel SH, Heiss G (2007) Usual choline and betaine dietary intake and incident coronary heart disease: the atherosclerosis risk in communities (ARIC) study. BMC cardiovas Disorders7(1):20 68. Dalmeijer GW, Olthof MR, Verhoef P, Bots ML, van der Schouw YT (2008) Prospective study on dietary intakes of folate, betaine, and choline and cardiovascular disease risk in women. Eur J Clin Nutr 62(3):386–394 69. Lin KC, Tsao HM, Chen CH, Chou P (2004) Hypertension was the major risk factor leading to development of cardiovascular diseases among men with hyperuricemia. J Rheumatol 31 (6):1152–1158

8 Gut Microbiota and Heart, Vascular Injury

135

70. Kunes J, Zicha J (2009) The interaction of genetic and environmental factors in the etiology of hypertension. Physiol Res 58(Suppl 2):S33–41 71. Moghadamrad S, McCoy KD, Geuking MB, Sagesser H, Kirundi J, Macpherson AJ et al (2015) Attenuated portal hypertension in germ-free mice: function of bacterial flora on the development of mesenteric lymphatic and blood vessels. Hepatology 61(5):1685–1695 72. Khalesi S, Sun J, Buys N, Jayasinghe R (2014) Effect of probiotics on blood pressure: a systematic review and meta-analysis of randomized, controlled trials. Hypertension 64 (4):897–903 73. Li J, Zhao F, Wang Y, Chen J, Tao J, Tian G et al (2017) Gut microbiota dysbiosis contributes to the development of hypertension. Microbiome 5(1):14 74. Yan Q, Gu Y, Li X, Yang W, Jia L, Chen C et al (2017) Alterations of the gut microbiome in hypertension. Front Cell Infect Microbiol 7:381 75. Yang T, Santisteban MM, Rodriguez V, Li E, Ahmari N, Carvajal JM et al (2015) Gut dysbiosis is linked to hypertension. Hypertension 65(6):1331–1340 76. Durgan DJ, Ganesh BP, Cope JL, Ajami NJ, Phillips SC, Petrosino JF et al (2016) Role of the gut microbiome in obstructive sleep apnea-induced hypertension. Hypertension 67 (2):469–474 77. Holmes E, Loo RL, Stamler J, Bictash M, Yap IK, Chan Q et al (2008) Human metabolic phenotype diversity and its association with diet and blood pressure. Nature 453(7193): 396–400 78. Marques FZ, Nelson E, Chu PY, Horlock D, Fiedler A, Ziemann M et al (2017) High-fiber diet and acetate supplementation change the gut microbiota and prevent the development of hypertension and heart failure in hypertensive mice. Circulation 135(10):964–977 79. de la Cuesta-Zuluaga J, Mueller NT, Alvarez-Quintero R, Velasquez-Mejia EP, Sierra JA, Corrales-Agudelo V et al (2019) Higher fecal short-chain fatty acid levels are associated with gut microbiome dysbiosis, obesity, hypertension and cardiometabolic disease risk factors. Nutrients 11(1):51 80. Marques FZ, Mackay CR, Kaye DM (2018) Beyond gut feelings: how the gut microbiota regulates blood pressure. Nat Rev Cardiol 15(1):20–32 81. Pluznick JL, Protzko RJ, Gevorgyan H, Peterlin Z, Sipos A, Han J et al (2013) Olfactory receptor responding to gut microbiota-derived signals plays a role in renin secretion and blood pressure regulation. Proc Natl Acad Sci USA 110(11):4410–4415 82. Dombkowski RA, Russell MJ, Olson KR (2004) Hydrogen sulfide as an endogenous regulator of vascular smooth muscle tone in trout. Am J Physiol Regul Integr Comp Physiol 286(4):R678–685 83. Nagpure BV, Bian JS (2016) Interaction of hydrogen sulfide with nitric oxide in the cardiovascular system. Oxid Med Cell longev 2016:6904327 84. Tomasova L, Dobrowolski L, Jurkowska H, Wrobel M, Huc T, Ondrias K et al (2016) Intracolonic hydrogen sulfide lowers blood pressure in rats. Nitric Oxide Biol Chem 60: 50–58 85. Meng G, Ma Y, Xie L, Ferro A, Ji Y (2015) Emerging role of hydrogen sulfide in hypertension and related cardiovascular diseases. Br J Pharmacol 172(23):5501–5511 86. Fuller JH, Shipley MJ, Rose G, Jarrett RJ, Keen H (1983) Mortality from coronary heart disease and stroke in relation to degree of glycaemia: the Whitehall study. Br Med J 287 (6396):867–870 87. Karimsab D, Razak SK (2013) Study of aerobic bacterial conjunctival flora in patients with diabetes mellitus. Nepal J Ophthalmol Biannu Peer-Rev Acad J Nepal Ophthalmic Soc (NEPJOPH) 5(1):28–32 88. Bilen H, Ates O, Astam N, Uslu H, Akcay G, Baykal O (2007) Conjunctival flora in patients with type 1 or type 2 diabetes mellitus. Adv Ther 24(5):1028–1035 89. Beli E, Yan Y, Moldovan L, Vieira CP, Gao R, Duan Y et al (2018) Restructuring of the gut microbiome by intermittent fasting prevents retinopathy and prolongs survival in db/db mice. Diabetes 67(9):1867–1879

136

C. Zeng and H. Tan

90. Larsen N, Vogensen FK, van den Berg FW, Nielsen DS, Andreasen AS, Pedersen BK et al (2010) Gut microbiota in human adults with type 2 diabetes differs from non-diabetic adults. PloS one 5(2):e9085 91. Qin J, Li Y, Cai Z, Li S, Zhu J, Zhang F et al (2012) A metagenome-wide association study of gut microbiota in type 2 diabetes. Nature 490(7418):55–60 92. Wu H, Esteve E, Tremaroli V, Khan MT, Caesar R, Manneras-Holm L et al (2017) Metformin alters the gut microbiome of individuals with treatment-naive type 2 diabetes, contributing to the therapeutic effects of the drug. Nat Med 23(7):850–858 93. Karlsson FH, Tremaroli V, Nookaew I, Bergstrom G, Behre CJ, Fagerberg B et al (2013) Gut metagenome in European women with normal, impaired and diabetic glucose control. Nature 498(7452):99–103 94. Sato J, Kanazawa A, Ikeda F, Yoshihara T, Goto H, Abe H et al (2014) Gut dysbiosis and detection of “live gut bacteria” in blood of Japanese patients with type 2 diabetes. Diabetes Care 37(8):2343–2350 95. Katsarou A, Gudbjornsdottir S, Rawshani A, Dabelea D, Bonifacio E, Anderson BJ et al (2017) Type 1 diabetes mellitus. Nat Rev Dis Primers 3:17016 96. Giongo A, Gano KA, Crabb DB, Mukherjee N, Novelo LL, Casella G et al (2011) Toward defining the autoimmune microbiome for type 1 diabetes. ISME J 5(1):82–91 97. Brown CT, Davis-Richardson AG, Giongo A, Gano KA, Crabb DB, Mukherjee N et al (2011) Gut microbiome metagenomics analysis suggests a functional model for the development of autoimmunity for type 1 diabetes. PloS one 6(10):e25792 98. de Goffau MC, Luopajarvi K, Knip M, Ilonen J, Ruohtula T, Harkonen T et al (2013) Fecal microbiota composition differs between children with beta-cell autoimmunity and those without. Diabetes 62(4):1238–1244 99. Bennion LJ, Grundy SM (1977) Effects of diabetes mellitus on cholesterol metabolism in man. New Engl J Med 296(24):1365–1371 100. Brufau G, Bahr MJ, Staels B, Claudel T, Ockenga J, Boker KH et al (2010) Plasma bile acids are not associated with energy metabolism in humans. Nutr Metab 7(1):73 101. Li T, Owsley E, Matozel M, Hsu P, Novak CM, Chiang JY (2010) Transgenic expression of cholesterol 7alpha-hydroxylase in the liver prevents high-fat diet-induced obesity and insulin resistance in mice. Hepatology 52(2):678–690 102. Brufau G, Stellaard F, Prado K, Bloks VW, Jonkers E, Boverhof R et al (2010) Improved glycemic control with colesevelam treatment in patients with type 2 diabetes is not directly associated with changes in bile acid metabolism. Hepatology 52(4):1455–1464 103. Schwartz SL, Lai YL, Xu J, Abby SL, Misir S, Jones MR et al (2010) The effect of colesevelam hydrochloride on insulin sensitivity and secretion in patients with type 2 diabetes: a pilot study. Metab Syndr Relat Disord 8(2):179–188 104. Kobayashi M, Ikegami H, Fujisawa T, Nojima K, Kawabata Y, Noso S et al (2007) Prevention and treatment of obesity, insulin resistance, and diabetes by bile acid-binding resin. Diabetes 56(1):239–247 105. Shang Q, Saumoy M, Holst JJ, Salen G, Xu G (2010) Colesevelam improves insulin resistance in a diet-induced obesity (F-DIO) rat model by increasing the release of GLP-1. Am J Physiol Gastrointest Liver Physiol 298(3):G419–424 106. Mudaliar S, Henry RR, Sanyal AJ, Morrow L, Marschall HU, Kipnes M et al (2013) Efficacy and safety of the farnesoid X receptor agonist obeticholic acid in patients with type 2 diabetes and nonalcoholic fatty liver disease. Gastroenterology 145(3):574–582 e571 107. Neuschwander-Tetri BA, Loomba R, Sanyal AJ, Lavine JE, Van Natta ML, Abdelmalek MF et al (2015) Farnesoid X nuclear receptor ligand obeticholic acid for non-cirrhotic, non-alcoholic steatohepatitis (FLINT): a multicentre, randomised, placebo-controlled trial. Lancet 385(9972):956–965 108. Kohli R, Kirby M, Setchell KD, Jha P, Klustaitis K, Woollett LA et al (2010) Intestinal adaptation after ileal interposition surgery increases bile acid recycling and protects against obesity-related comorbidities. Am J Physiol Gastrointest Liver Physiol 299(3):G652–660

8 Gut Microbiota and Heart, Vascular Injury

137

109. Patti ME, Houten SM, Bianco AC, Bernier R, Larsen PR, Holst JJ et al (2009) Serum bile acids are higher in humans with prior gastric bypass: potential contribution to improved glucose and lipid metabolism. Obesity 17(9):1671–1677 110. Vatanen T, Franzosa EA, Schwager R, Tripathi S, Arthur TD, Vehik K et al (2018) The human gut microbiome in early-onset type 1 diabetes from the TEDDY study. Nature 562 (7728):589–594 111. Sanna S, van Zuydam NR, Mahajan A, Kurilshikov A, Vich Vila A, Vosa U et al (2019) Causal relationships among the gut microbiome, short-chain fatty acids and metabolic diseases. Nat Genet 51(4):600–605 112. McNelis JC, Lee YS, Mayoral R, van der Kant R, Johnson AM, Wollam J et al (2015) GPR43 potentiates beta-cell function in obesity. Diabetes 64(9):3203–3217 113. Veprik A, Laufer D, Weiss S, Rubins N, Walker MD (2016) GPR41 modulates insulin secretion and gene expression in pancreatic beta-cells and modifies metabolic homeostasis in fed and fasting states. FASEB J Off Publ Fed Am Soc Exp Biol 30(11):3860–3869 114. Priyadarshini M, Villa SR, Fuller M, Wicksteed B, Mackay CR, Alquier T et al (2015) An acetate-specific GPCR, FFAR2, regulates insulin secretion. Molec Endocrinol 29(7):1055– 1066 115. Pingitore A, Chambers ES, Hill T, Maldonado IR, Liu B, Bewick G et al (2017) The diet-derived short chain fatty acid propionate improves beta-cell function in humans and stimulates insulin secretion from human islets in vitro. Diabetes Obes Metab 19(2):257–265 116. Yang G, Yang W, Wu L, Wang R (2007) H2S, endoplasmic reticulum stress, and apoptosis of insulin-secreting beta cells. J Biol Chem 282(22):16567–16576 117. Wu L, Yang W, Jia X, Yang G, Duridanova D, Cao K et al (2009) Pancreatic islet overproduction of H2S and suppressed insulin release in Zucker diabetic rats. Lab Investig J Tech Methods Pathol 89(1):9–67 118. Zhang L, Yang G, Untereiner A, Ju Y, Wu L, Wang R (2013) Hydrogen sulfide impairs glucose utilization and increases gluconeogenesis in hepatocytes. Endocrinology 154 (1):114–126 119. Kaneko Y, Kimura T, Taniguchi S, Souma M, Kojima Y, Kimura Y et al (2009) Glucose-induced production of hydrogen sulfide may protect the pancreatic beta-cells from apoptotic cell death by high glucose. FEBS Lett 583(2):377–382 120. Taniguchi S, Kang L, Kimura T, Niki I (2011) Hydrogen sulphide protects mouse pancreatic beta-cells from cell death induced by oxidative stress, but not by endoplasmic reticulum stress. Br J Pharmacol 162(5):1171–1178 121. Jain SK, Bull R, Rains JL, Bass PF, Levine SN, Reddy S et al (2010) Low levels of hydrogen sulfide in the blood of diabetes patients and streptozotocin-treated rats causes vascular inflammation? Antioxid Redox Signal 12(11):1333–1337 122. Lever M, George PM, Slow S, Bellamy D, Young JM, Ho M et al (2014) Betaine and trimethylamine-N-oxide as predictors of cardiovascular outcomes show different patterns in diabetes mellitus: an observational study. PloS one 9(12):e114969 123. Tang WH, Wang Z, Li XS, Fan Y, Li DS, Wu Y et al (2017) Increased trimethylamine N-oxide portends high mortality risk independent of glycemic control in patients with type 2 diabetes mellitus. Clin Chem 63(1):297–306 124. Gao X, Liu X, Xu J, Xue C, Xue Y, Wang Y (2014) Dietary trimethylamine N-oxide exacerbates impaired glucose tolerance in mice fed a high fat diet. J Biosci Bioeng 118 (4):476–481 125. McEntyre CJ, Lever M, Chambers ST, George PM, Slow S, Elmslie JL et al (2015) Variation of betaine, N, N-dimethylglycine, choline, glycerophosphorylcholine, taurine and trimethylamine-N-oxide in the plasma and urine of overweight people with type 2 diabetes over a two-year period. Ann Clin Biochem 52(Pt 3):352–360 126. Lupachyk S, Watcho P, Stavniichuk R, Shevalye H, Obrosova IG (2013) Endoplasmic reticulum stress plays a key role in the pathogenesis of diabetic peripheral neuropathy. Diabetes 62(3):944–952

138

C. Zeng and H. Tan

127. Xu J, Zhang J, Cai S, Dong J, Yang JY, Chen Z (2009) Metabonomics studies of intact hepatic and renal cortical tissues from diabetic db/db mice using high-resolution magic-angle spinning 1H NMR spectroscopy. Anal Bioanal Chem 393(6–7):1657–1668 128. Berenson GS, Srinivasan SR, Bao W, Newman WP, Tracy RE, Wattigney WA (1998) Association between multiple cardiovascular risk factors and atherosclerosis in children and young adults. The Bogalusa heart study. New England J Med 338(23):1650–1656 129. Hill JO (2006) Understanding and addressing the epidemic of obesity: an energy balance perspective. Endocr Rev 27(7):750–761 130. Wostmann BS, Larkin C, Moriarty A, Bruckner-Kardoss E (1983) Dietary intake, energy metabolism, and excretory losses of adult male germfree Wistar rats. Lab Anim Sci 33 (1):46–50 131. Backhed F, Manchester JK, Semenkovich CF, Gordon JI (2007) Mechanisms underlying the resistance to diet-induced obesity in germ-free mice. Proc Natl Acad Sci USA 104(3):979– 984 132. Backhed F, Ding H, Wang T, Hooper LV, Koh GY, Nagy A et al (2004) The gut microbiota as an environmental factor that regulates fat storage. Proc Natl Acad Sci USA 101 (44):15718–15723 133. Ravussin Y, Koren O, Spor A, LeDuc C, Gutman R, Stombaugh J et al (2012) Responses of gut microbiota to diet composition and weight loss in lean and obese mice. Obesity 20 (4):738–747 134. Walters WA, Xu Z, Knight R (2014) Meta-analyses of human gut microbes associated with obesity and IBD. FEBS Lett 588(22):4223–4233 135. Gao R, Zhu C, Li H, Yin M, Pan C, Huang L et al (2018) Dysbiosis signatures of gut microbiota along the sequence from healthy, young patients to those with overweight and obesity. Obesity 26(2):351–361 136. Joly F, Mayeur C, Bruneau A, Noordine ML, Meylheuc T, Langella P et al (2010) Drastic changes in fecal and mucosa-associated microbiota in adult patients with short bowel syndrome. Biochimie 92(7):753–761 137. Chang CJ, Lin CS, Lu CC, Martel J, Ko YF, Ojcius DM et al (2015) Ganoderma lucidum reduces obesity in mice by modulating the composition of the gut microbiota. Nat Commun 6:7489 138. Gomez-Arango LF, Barrett HL, McIntyre HD, Callaway LK, Morrison M, Dekker Nitert M et al (2016) Increased systolic and diastolic blood pressure is associated with altered gut microbiota composition and butyrate production in early pregnancy. Hypertension 68 (4):974–981 139. Maslowski KM, Vieira AT, Ng A, Kranich J, Sierro F, Yu D et al (2009) Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature 461 (7268):1282–1286 140. de Git KC, Adan RA (2015) Leptin resistance in diet-induced obesity: the role of hypothalamic inflammation. Obes Rev Off J Int Assoc Study Obes 16(3):207–224 141. McNay DE, Speakman JR (2012) High fat diet causes rebound weight gain. Molec Metab 2 (2):103–108 142. Waldecker M, Kautenburger T, Daumann H, Busch C, Schrenk D (2008) Inhibition of histone-deacetylase activity by short-chain fatty acids and some polyphenol metabolites formed in the colon. J Nutr Biochem 19(9):587–593 143. Vanhoutvin SA, Troost FJ, Hamer HM, Lindsey PJ, Koek GH, Jonkers DM et al (2009) Butyrate-induced transcriptional changes in human colonic mucosa. PloS one 4(8):e6759 144. Kimura I, Inoue D, Hirano K, Tsujimoto G (2014) The SCFA receptor GPR43 and energy metabolism. Front Endocrinol 5:85 145. Inoue D, Tsujimoto G, Kimura I (2014) Regulation of energy homeostasis by GPR41. Front Endocrinol 5:81 146. Lin HV, Frassetto A, Kowalik EJ Jr, Nawrocki AR, Lu MM, Kosinski JR et al (2012) Butyrate and propionate protect against diet-induced obesity and regulate gut hormones via free fatty acid receptor 3-independent mechanisms. PloS one 7(4):e35240

8 Gut Microbiota and Heart, Vascular Injury

139

147. Perry RJ, Peng L, Barry NA, Cline GW, Zhang D, Cardone RL et al (2016) Acetate mediates a microbiome-brain-beta-cell axis to promote metabolic syndrome. Nature 534(7606): 213–217 148. van der Beek CM, Canfora EE, Lenaerts K, Troost FJ, Damink S, Holst JJ et al (2016) Distal, not proximal, colonic acetate infusions promote fat oxidation and improve metabolic markers in overweight/obese men. Clin Sci 130(22):2073–2082 149. Canfora EE, van der Beek CM, Jocken JWE, Goossens GH, Holst JJ, Olde Damink SWM et al (2017) Colonic infusions of short-chain fatty acid mixtures promote energy metabolism in overweight/obese men: a randomized crossover trial. Sci Rep 7(1):2360 150. Watanabe M, Houten SM, Mataki C, Christoffolete MA, Kim BW, Sato H et al (2006) Bile acids induce energy expenditure by promoting intracellular thyroid hormone activation. Nature 439(7075):484–489 151. Kumar DP, Asgharpour A, Mirshahi F, Park SH, Liu S, Imai Y et al (2016) Activation of transmembrane bile acid receptor TGR5 modulates pancreatic islet alpha cells to promote glucose homeostasis. J Biol Chem 291(13):6626–6640 152. Thomas C, Gioiello A, Noriega L, Strehle A, Oury J, Rizzo G et al (2009) TGR5-mediated bile acid sensing controls glucose homeostasis. Cell Metab 10(3):167–177 153. Trabelsi MS, Daoudi M, Prawitt J, Ducastel S, Touche V, Sayin SI et al (2015) Farnesoid X receptor inhibits glucagon-like peptide-1 production by enteroendocrine L cells. Nat Commun 6:7629 154. Kim KH, Choi S, Zhou Y, Kim EY, Lee JM, Saha PK et al (2017) Hepatic FXR/SHP axis modulates systemic glucose and fatty acid homeostasis in aged mice. Hepatology 66(2): 498–509 155. Schmitt J, Kong B, Stieger B, Tschopp O, Schultze SM, Rau M et al (2015) Protective effects of farnesoid X receptor (FXR) on hepatic lipid accumulation are mediated by hepatic FXR and independent of intestinal FGF15 signal. Liver Int Off J Int Assoc Study Liver 35 (4):1133–1144 156. Jiang C, Xie C, Lv Y, Li J, Krausz KW, Shi J et al (2015) Intestine-selective farnesoid X receptor inhibition improves obesity-related metabolic dysfunction. Nat Commun 6:10166 157. Li F, Jiang C, Krausz KW, Li Y, Albert I, Hao H et al (2013) Microbiome remodelling leads to inhibition of intestinal farnesoid X receptor signalling and decreased obesity. Nat Commun 4:2384 158. Broeders EP, Nascimento EB, Havekes B, Brans B, Roumans KH, Tailleux A et al (2015) The bile acid chenodeoxycholic acid increases human brown adipose tissue activity. Cell Metab 22(3):418–426 159. Chavez-Talavera O, Tailleux A, Lefebvre P, Staels B (2017) Bile acid control of metabolism and inflammation in obesity, type 2 diabetes, dyslipidemia, and nonalcoholic fatty liver disease. Gastroenterology 152(7):1679–1694 e1673 160. Nagatomo Y, Tang WHW (2015) Intersections between microbiome and heart failure: revisiting the gut hypothesis. J Card Fail 21(12):973–980 161. Krack A, Sharma R, Figulla HR, Anker SD (2005) The importance of the gastrointestinal system in the pathogenesis of heart failure. Eur Heart J 26(22):2368–2374 162. Dinakaran V, Rathinavel A, Pushpanathan M, Sivakumar R, Gunasekaran P, Rajendhran J (2014) Elevated levels of circulating DNA in cardiovascular disease patients: metagenomic profiling of microbiome in the circulation. PloS one 9(8):e105221 163. Amar J, Lange C, Payros G, Garret C, Chabo C, Lantieri O et al (2013) Blood microbiota dysbiosis is associated with the onset of cardiovascular events in a large general population: the D.E.S.I.R. study. PloS one 8(1):e54461 164. Pasini E, Aquilani R, Testa C, Baiardi P, Angioletti S, Boschi F et al (2016) Pathogenic gut flora in patients with chronic heart failure. JACC Heart Fail 4(3):220–227 165. Luedde M, Winkler T, Heinsen FA, Ruhlemann MC, Spehlmann ME, Bajrovic A et al (2017) Heart failure is associated with depletion of core intestinal microbiota. ESC Heart Fail 4(3):282–290

140

C. Zeng and H. Tan

166. Tang WH, Wang Z, Levison BS, Koeth RA, Britt EB, Fu X et al (2013) Intestinal microbial metabolism of phosphatidylcholine and cardiovascular risk. New Engl J Med 368(17): 1575–1584 167. Tang WH, Wang Z, Fan Y, Levison B, Hazen JE, Donahue LM et al (2014) Prognostic value of elevated levels of intestinal microbe-generated metabolite trimethylamine-N-oxide in patients with heart failure: refining the gut hypothesis. J Am Coll Cardiol 64(18):1908– 1914 168. Organ CL, Otsuka H, Bhushan S, Wang Z, Bradley J, Trivedi R et al (2016) Choline diet and its gut microbe-derived metabolite, trimethylamine N-oxide, exacerbate pressure overload-induced heart failure. Circ Heart Fail 9(1):e002314 169. Li Z, Wu Z, Yan J, Liu H, Liu Q, Deng Y et al (2019) Gut microbe-derived metabolite trimethylamine N-oxide induces cardiac hypertrophy and fibrosis. Lab Investig J Tech Methods Pathol 99(3): 46–357 170. Tang WH, Wang Z, Kennedy DJ, Wu Y, Buffa JA, Agatisa-Boyle B et al (2015) Gut microbiota-dependent trimethylamine N-oxide (TMAO) pathway contributes to both development of renal insufficiency and mortality risk in chronic kidney disease. Circ Res 116(3):448–455 171. Freeman JV, Wang Y, Akar J, Desai N, Krumholz H (2017) National trends in atrial fibrillation hospitalization, readmission, and mortality for medicare beneficiaries, 1999– 2013. Circulation 135(13):1227–1239 172. Svingen GFT, Zuo H, Ueland PM, Seifert R, Loland KH, Pedersen ER et al (2018) Increased plasma trimethylamine-N-oxide is associated with incident atrial fibrillation. Int J Cardiol 267:100–106 173. Yu L, Meng G, Huang B, Zhou X, Stavrakis S, Wang M et al (2018) A potential relationship between gut microbes and atrial fibrillation: Trimethylamine N-oxide, a gut microbe-derived metabolite, facilitates the progression of atrial fibrillation. Int J Cardiol 255:92–98 174. Pastori D, Carnevale R, Nocella C, Novo M, Santulli M, Cammisotto V et al (2017) Gut-derived serum lipopolysaccharide is associated with enhanced risk of major adverse cardiovascular events in atrial fibrillation: effect of adherence to mediterranean diet. J Am Heart Assoc 6(6) 175. Lam V, Su J, Koprowski S, Hsu A, Tweddell JS, Rafiee P et al (2012) Intestinal microbiota determine severity of myocardial infarction in rats. FASEB J Off Publ Fed Am Soc Exp Biol 26(4):1727–1735 176. Lam V, Su J, Hsu A, Gross GJ, Salzman NH, Baker JE (2016) Intestinal microbial metabolites are linked to severity of myocardial infarction in rats. PloS one 11(8):e0160840 177. Wu ZX, Li SF, Chen H, Song JX, Gao YF, Zhang F et al (2017) The changes of gut microbiota after acute myocardial infarction in rats. PloS one 12(7):e0180717 178. Zhou X, Li J, Guo J, Geng B, Ji W, Zhao Q et al (2018) Gut-dependent microbial translocation induces inflammation and cardiovascular events after ST-elevation myocardial infarction. Microbiome 6(1):66 179. Gan XT, Ettinger G, Huang CX, Burton JP, Haist JV, Rajapurohitam V et al (2014) Probiotic administration attenuates myocardial hypertrophy and heart failure after myocardial infarction in the rat. Circ Heart Fail 7(3):491–499 180. Tang TWH, Chen HC, Chen CY, Yen CYT, Lin CJ, Prajnamitra RP et al (2019) Loss of gut microbiota alters immune system composition and cripples postinfarction cardiac repair. Circulation 139(5):647–659 181. Haissman JM, Knudsen A, Hoel H, Kjaer A, Kristoffersen US, Berge RK et al (2016) Microbiota-dependent marker TMAO is elevated in silent ischemia but is not associated with first-time myocardial infarction in HIV infection. J Acquir Immune Defic Syndr 71(2): 130–136 182. Trasande L, Blustein J, Liu M, Corwin E, Cox LM, Blaser MJ (2013) Infant antibiotic exposures and early-life body mass. Int J Obes 37(1):16–23 183. Cho I, Yamanishi S, Cox L, Methe BA, Zavadil J, Li K et al (2012) Antibiotics in early life alter the murine colonic microbiome and adiposity. Nature 488(7413):621–626

8 Gut Microbiota and Heart, Vascular Injury

141

184. Guo Z, Liu XM, Zhang QX, Shen Z, Tian FW, Zhang H et al (2011) Influence of consumption of probiotics on the plasma lipid profile: a meta-analysis of randomised controlled trials. Nutr Metab Cardiovasc Dis (NMCD) 21(11):844–850 185. Sirilun S, Chaiyasut C, Kantachote D, Luxananil P (2010) Characterisation of non human origin probiotic lactobacillus plantarum with cholesterol-lowering property. Afr J Microbiol Res 4(10):994–1000 186. Agerholm-Larsen L, Raben A, Haulrik N, Hansen AS, Manders M, Astrup A (2000) Effect of 8 week intake of probiotic milk products on risk factors for cardiovascular diseases. Eur J Clin Nutr 54(4):288–297 187. Kekkonen RA, Sysi-Aho M, Seppanen-Laakso T, Julkunen I, Vapaatalo H, Oresic M et al (2008) Effect of probiotic lactobacillus rhamnosus GG intervention on global serum lipidomic profiles in healthy adults. World J Gastroenterol 14(20):3188–3194 188. Simon MC, Strassburger K, Nowotny B, Kolb H, Nowotny P, Burkart V et al (2015) Intake of lactobacillus reuteri improves incretin and insulin secretion in glucose-tolerant humans: a proof of concept. Diabet Care 38(10):1827–1834 189. He M, Shi B (2017) Gut microbiota as a potential target of metabolic syndrome: the role of probiotics and prebiotics. Cell Biosci 7:54 190. Roberfroid M, Gibson GR, Hoyles L, McCartney AL, Rastall R, Rowland I et al (2010) Prebiotic effects: metabolic and health benefits. Br J Nutr 104(Suppl 2):S1–63 191. Gibson GR, Roberfroid MB (1995) Dietary modulation of the human colonic microbiota: introducing the concept of prebiotics. J Nutr 125(6):1401–1412 192. Tang WHW, Kitai T, Hazen SL (2017) Gut microbiota in cardiovascular health and disease. Circ Res 120(7):1183–1196 193. Parnell JA, Reimer RA (2009) Weight loss during oligofructose supplementation is associated with decreased ghrelin and increased peptide YY in overweight and obese adults. Am J Clin Nutr 89(6):1751–1759 194. Everard A, Lazarevic V, Derrien M, Girard M, Muccioli GG, Neyrinck AM et al (2011) Responses of gut microbiota and glucose and lipid metabolism to prebiotics in genetic obese and diet-induced leptin-resistant mice. Diabetes 60(11):2775–2786 195. van Nood E, Vrieze A, Nieuwdorp M, Fuentes S, Zoetendal EG, de Vos WM et al (2013) Duodenal infusion of donor feces for recurrent clostridium difficile. New Engl J Med 368 (5):407–415 196. Vrieze A, Van Nood E, Holleman F, Salojarvi J, Kootte RS, Bartelsman JF et al (2012) Transfer of intestinal microbiota from lean donors increases insulin sensitivity in individuals with metabolic syndrome. Gastroenterology 143(4):913–916 e917 197. Konstantinov SR, Peppelenbosch MP (2013) Fecal microbiota transfer may increase irritable bowel syndrome and inflammatory bowel diseases-associated bacteria. Gastroenterology 144 (4):e19–20 198. Xiao S, Fei N, Pang X, Shen J, Wang L, Zhang B et al (2014) A gut microbiota-targeted dietary intervention for amelioration of chronic inflammation underlying metabolic syndrome. FEMS Microbiol Ecol 87(2):357–367

Chapter 9

Gut Microbiota and Endocrine Disorder Rui Li, Yifan Li, Cui Li, Dongying Zheng, and Peng Chen

Abstract The gut microbiome contains trillions of commensal microorganisms that maintain a symbiotic relationship with the host, and its profound effects on gastrointestinal diseases have been widely described. Recently, gut microbiota have emerged as important factors in endocrine system diseases. Disruption of the gut microbiota affects neuroendocrine homeostasis and promotes peripheral endocrine system diseases, including obesity, diabetes, and hyperuricemia. This chapter provides a comprehensive overview of the biological mechanisms of gut microbiota that participate in endocrine system pathologies and discusses potential novel therapies for these diseases. Keywords Gut microbiota

9.1

 Neuroendocrine disorder  Diabetes  Obesity

Gut Microbiota

Virtually, all multicellular organisms maintain an intimate relationship with microbes, and humans are no exception. A large number of bacteria, archaea, viruses, and unicellular eukaryotes coexist with the human body. The field of microbiota research has boomed over the last decade, resulting in the publication of a vast array of reports that describe both the composition of intestinal microbiota and their wide-ranging impact on host physiology and pathology [1–4]. Microbes live on our skin and in the genitourinary, gastrointestinal, and respiratory tracts, colonizing virtually every surface on the human body. The gastrointestinal tract (GIT) is the most heavily microbe-colonized organ [5], and recent studies suggest that over 35,000 bacterial species reside therein. Of the strict anaerobes, facultative anaerobes, and aerobes in the human GIT, the most important are Gram-positive bacteria of the phyla Firmicutes and Actinobacteria, followed by R. Li  Y. Li  C. Li  D. Zheng  P. Chen (&) Department of Pathophysiology, School of Basic Medical Sciences, Southern Medical University, No. 1838 Guangzhou Ave., Guangzhou 510515, China e-mail: [email protected] © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_9

143

144

R. Li et al.

the genera Clostridium, Bifidobacterium, Lactobacillus, Ruminococcus, and Streptococcus and Gram-negative bacteria belonging to the genera Bacteroides, Prevotella, and Akkermansia [6, 7]. These microorganisms are critical to healthy intestinal development in the host. Metabolic abnormalities, endocrine disorders, and other factors negatively impact the intestinal system, disrupting microorganisms and their secretions and causing immune system, endocrine, and urinary diseases [8–11].

9.2

Endocrine Systems

The endocrine system consists of the endocrine organs, including hypophysis, thyroid, parathyroid, adrenal gland, gonad, pancreas, and clusters of endocrine cells or tissues located in other organs, such as adipose tissue, pancreatic islets, and reproductive tissue. Endocrine glands and tissue cells secrete biologically active substances called hormones, which are released into the blood or lymphatic systems and circulate throughout the body. The endocrine system oversees the growth, development, metabolism, and reproduction of the whole organism.

9.2.1

Neuroendocrine System

A broad range of hormones are produced by the central endocrine glands, including the pineal gland, hypothalamus, and pituitary gland that comprise the neuroendocrine system. Pituitary gland The pituitary gland is a small structure at the base of the hypothalamus; its activity is regulated by projections from cell bodies located in the paraventricular nucleus (PVN). The pituitary is divided into anterior, intermediate, and posterior lobes. It secretes several different hormones, such as growth hormone, thyroid stimulating hormone, adrenocorticotropic hormone, gonadotropin, oxytocin, prolactin, and black cell-stimulating hormone, and also stores and releases antidiuretic hormone secreted by the hypothalamus. These hormones are important for metabolism, growth, development, and reproduction [7, 12–14]. Pineal gland The pineal gland is a small neuroendocrine organ that secretes melatonin. Pineal activity is strongly driven by external light information relayed through the retinohypothalamic tract and sympathetic nervous system. Melatonin has many targets throughout the body, and its receptors are located in the brain and peripheral organs like the heart, liver, adrenal glands, testes, and ovaries [15–17].

9 Gut Microbiota and Endocrine Disorder

9.2.2

145

Peripheral Endocrine System

Thyroid gland The thyroid is the largest endocrine gland in adults, containing triiodothyronine (T3) and thyroxine that regulate metabolism, growth rate, and other related biological processes and systems [18–20]. Adrenal glands The adrenal glands rest atop the kidneys and consist of an outer cortex that secretes corticosteroids and an inner medulla that produces the catecholamines epinephrine and norepinephrine. In addition to hormonal input from the pituitary gland, the adrenal gland receives neural input from the suprachiasmatic nucleus (SCN) via a projection through the PVN from the intermediolateral column of the spinal cord. Adrenal glands are critical to maintaining normal physiological function [21–23]. Adipose tissue Adipose tissue has recently come to the forefront as an important peripheral endocrine tissue. Adipocytes secrete chemical messengers such as leptin, adiponectin, tumor necrosis factor a (TNF-a), angiotensinogen, and inflammatory cytokines. These molecules affect insulin sensitivity, blood pressure, endothelial function, and inflammatory responses, all of which participate in many important pathophysiological processes [12, 24–26]. Pancreatic islets Pancreatic islets are clusters of cells scattered between pancreatic acinar cells and account for between 1 and 2% of total pancreatic volume. Pancreatic islet cells secrete glucagon, insulin, and pancreatic polypeptides (PP). Insulin is the only hormone in the body that lowers blood sugar and promotes the synthesis of glycogen, fat, and protein, and its primary physiological function is to regulate metabolic processes. Insulin can promote the uptake and utilization of glucose in tissues, cells and glycogen synthesis and can inhibit gluconeogenesis and reduce blood glucose. As for lipid metabolism, insulin is capable of promoting fatty acid synthesis and fat storage as well as reducing fat decomposition. It also facilitates amino acid to transfer into cells and modulates protein synthesis therein. Overall, insulin controls energy storage and use [27–29]. Glucagon facilitates glycogen decomposition and gluconeogenesis in the liver that significantly increases blood glucose. It also promotes fat decomposition and increases ketone body production [30, 31]. Protein intake stimulates PP secretion; fat and sugar do so as well, but to a lesser degree. PP can inhibit the secretion of pancreatic juice and bile after meals and the secretion of gastric acid caused by pentagastrin and participates in a wide range of processes in the GIT.

146

9.2.3

R. Li et al.

Consequences of Endocrine Disorders

There are dozens of endocrine diseases, and each organ malfunctions in distinctive ways. Common endocrine-related metabolic diseases include hypopituitarism, thyroid disorders, adrenal cortex diseases, diabetes, obesity, thyroid diseases, and hyperuricemia. Thyroid diseases Thyroid diseases are medical conditions resulting from impairments of thyroid function. Different thyroid diseases include Hashimoto’s thyroiditis (HT), hyperthyroidism, and hypothyroidism. Hyperthyroidism is the most common of these, such as Graves’ disease and Plummer’s disease [22, 32]. Adrenocortical diseases Adrenocortical diseases include Cushing’s syndrome and primary chronic adrenocortical dysfunction. Cushing’s syndrome is caused by the excessive production of glucocorticoids (mainly cortisol) by the adrenal glands, and reasons for which can vary. The main clinical manifestations are full-moon face, sanguinity, concentric obesity, purple lines in the skin, acne, diabetes, hypertension, and osteoporosis [33, 34]. Primary chronic adrenal cortical dysfunction, known as Addison’s disease, is caused by insufficient adrenal cortical hormone secretion due to the destruction of the bilateral adrenal glands by autoimmune pathology, tuberculosis, fungal infection, tumor, and leukemia, among others [35, 36]. Secondary chronic adrenocortical dysfunction is caused by insufficient adrenocortical hormone (ACTH) resulting from hypothalamic–pituitary axis pathologies [37]. Autoimmune diabetes Autoimmune diabetes, also called type 1 diabetes, is caused by autoimmune destruction mediated by pancreatic islet T cells. It is characterized by the development of islet inflammation and an increased presence of islet B cell autoantibodies. Type 2 diabetes mellitus and obesity Conversely, patients with type 2 diabetes mellitus (T2DM) produce normal or even excessive amounts of insulin but demonstrate significant insulin resistance [38–40]. Obesity due to an imbalance between energy intake and expenditure is an important cause of certain endocrine pathologies and T2DM. Some also have implicated socioeconomic factors in the development of obesity, such as poor diet and a sedentary lifestyle [41–43]. Hyperuricemia and urarthritis Uric acid (UA) is a terminal metabolite of human purine compounds, and purine metabolism disorders can lead to hyperuricemia. Uarthritis, or arthritis associated with gout, is a heterogeneous disease caused by monosodium urate (MSU) deposition directly related to hyperuricemia. Severe uarthritis is accompanied by joint destruction and renal function damage. Clinical studies have shown

9 Gut Microbiota and Endocrine Disorder

147

that hyperuricemia associates with chronic interstitial nephritis, UA urinary calculi, diabetes, and other endocrine disorders [44–46].

9.3

Gut Microbiota and the Endocrine System

Studies over recent years have shown that the gut microbiome can influence host endocrine functions via a range of bacteria-derived metabolites (Fig. 9.1).

9.3.1

Gut Microbiota and Neuroendocrine System Diseases

The gut–brain axis refers to the communication between the gut and central nervous system (CNS) that regulates different physiological processes via the endocrine system [47, 48]. Gut microbiota are significant players in the gut–brain axis. Changes in the microbial community can lead to the production of different hormones, metabolites, and immune factors in the GIT [49, 50]. Gut microbiota interferes with vitally important hypothalamus, pituitary gland, and pineal gland homeostasis through its interaction with the host neuroendocrine system [50–53]. The mechanisms through which gut microbiota influences neuroendocrine function and, consequently, host behavior have yet to be fully deciphered. Increasing evidence suggests, however, that gut microbiota activity is facilitated by

Fig. 9.1 Gut microbiota metabolites affect endocrine function in the host

148

R. Li et al.

direct production of neuroendocrine metabolites, such as short-chain fatty acids (SCFAs), neurotransmitters, GIT hormones, precursors to neuroactive compounds such as tryptophan and kynurenine, and, indirectly, modulation of inflammatory responses, immune responses, and hormonal secretion [54–59]. Short-Chain Fatty Acids The microbiome converts indigestible carbohydrates into SCFAs. SCFAs include formic acid, isobutyrate, propionate, acetic acid, isovalerate, and butyrate [60]. Increasing evidence suggests that SCFAs can directly influence brain function and behavior [61]. Butyric acid and propionic acid can modulate the synthesis of dopamine and norepinephrine by increasing the expression of tyrosine hydroxylase genes [62]. Propionic acid can regulate serotonin neurotransmission and reduce levels of ɣ-aminobutyric acid (GABA), serotonin, and dopamine in the body [63]. Interestingly, SCFAs have also been shown to affect the maturation and function of microglia, the macrophages of the CNS. Mice born and raised in a sterile environment show microglial defects that were reversed after long-term SCFA treatment [64]. Because gut microbial-mediated SCFAs can bind to enteroendocrine cell (EEC) receptors FFAR1 and FFAR3, stimulating the secretion of related peptides and hormones, mice deficient for the SCFA receptor FFAR2 recapitulated the microglia defects observed under germ-free conditions. These findings suggest that bacteria can regulate microglia maturation and function, whereas microglia impairment can be rectified to a certain extent by complex SCFAs [65, 66]. Dietary regulation of SCFAs also alleviated blood–brain barrier defects in sterile mice, which may be related to circuits that regulate CNS and hypothalamus function [67, 68]. Neurotransmitters and tryptophan metabolism Studies have shown that gut bacteria can produce several neurotransmitters in addition to SCFAs, such as dopamine, norepinephrine, and GABA; these molecules influence hypothalamus function and act in the main neuroendocrine axis of the host [69–72]. Microbial glutamate and GABA signaling also exert certain neuro-endocrinological effects [73]. Gut microbiota can decarboxylate levodopa to dopamine via tyrosine decarboxylases. A large number of intestinal bacteria tyrosine decarboxylases were observed at the proximal end of the small intestine, where levodopa is absorbed in Parkinson’s disease patients, which had a significant impact on levodopa levels in rat plasma. This result demonstrates that an abundance of bacterial tyrosine decarboxylase in the proximal small intestine could explain the increased dosages required during levodopa treatment in Parkinson’s disease patients [74]. Human intestinally derived strains of Lactobacillus and Bifidobacterium have been shown to produce GABA, a main inhibitory neurotransmitter in the CNS that is also associated with depression and anxiety [75, 76]. Serotonin (5-HT) is another important neurotransmitter involved in CNS adaptation responses to tryptophan metabolism [70, 77]. The gut microbiota might metabolize tryptophan to increase plasma levels of 5-HT. Streptococcus, Enterococcus, and Escherichia coli also produce 5-HT [78]. In the context of microbial-derived

9 Gut Microbiota and Endocrine Disorder

149

neurotransmitters and brain functions, further investigation is needed to delineate the direct or indirect effects of microbiota-induced changes in peripheral neurotransmitter levels on neuroendocrine functions and mechanisms and the blood– brain barrier. Enteroendocrine cells and enteroendocrine signaling EECs are widely distributed in the GIT and they can regulate two-way communication between the intestinal tract and the brain. Hormones and peptides secreted by EECs act on receptors in the vagus afferent pathway and influence different neurophysiological functions of the host [79, 80]. The main hormones secreted by EECs are cholecystokinin (CCK), peptide YY (PYY), and glucagon-like peptide-1 (GLP-1), which regulate the catabolism of fat, protein and carbohydrates [81, 82]. Studies have shown that germ-free mice have fewer EECs and lower levels of PYY, GLP-1, and CCK than conventionally colonized mice [81, 83, 84]. Enteral E. coli protein infusion can increase plasma PYY and GLP-1 levels, possibly due to casein hydrolytic protease B (ClpB) produced by E. coli [66]. Peptides such as ghrelin, GLP-1, PYY and CCK secreted by the gut significantly impact energy balance and homeostasis by inducing satiety and meal termination [85]. Gut peptides can also bind with homologous receptors on immune cells and vagus nerve endings to achieve indirect entero-brain communication, influencing the expression of endocrine peptides and regulating anxiety and depression. The concentration of intestinal peptides is regulated by signals from the intestinal microflora and by gut microbiota composition [86]. Vagus nerve The vagus nerve is a key mediator of cross-communication between gut and brain and it is a major player in homeostasis, sensing the milieu intérieur and modulating nervous and endocrine system activity to maintain gastrointestinal health. Gut microbiota can control the production and release of neurotransmitters by interacting with the CNS through the vagus nerve. Diet-driven unfavorable microbiota composition or imbalances can lead to an increase of pro-inflammatory byproducts, such as lipopolysaccharides (LPS). A chronic increase of circulating LPS has been associated with impairments in the gut–brain axis and vagus-mediated satiety signaling. Evidence suggests that correcting imbalances in microbiota composition in rats fed with high-fat diet (HFD) prevents vagal remodeling [87]. Mice that underwent vasectomies demonstrated that effects of Lactobacillus rhamnosus administration on mice behavior were dependent on the presence of the nerve. Chronic treatment with L. rhamnosus (JB-1) was shown to induce region-dependent alterations in GABA(B1b) mRNA in the brain, along with increases in the cingulate and prelimbic cortical regions and concomitant reductions in expression in the hippocampus, amygdala, and locus coeruleus, in comparison with control-fed mice. (JB-1) administration also reduced GABA (Aa2) mRNA expression in the prefrontal cortex and amygdala, but increased mRNA expression thereof in the hippocampus. Importantly, (JB-1) was

150

R. Li et al.

Fig. 9.2 The gut–brain axis is associated with neuroendocrine function

associated with reduced levels of stress-induced corticosterone and anxiety- and depression-related behavior [83, 88]. Together, these findings highlight the importance of the gut–brain axis in neuroendocrine system functioning and suggest that gut microbiota could be used to assist in the treatment of neuroendocrine diseases such as anxiety and depression (Fig. 9.2).

9.3.2

Gut Microbiota and Peripheral Endocrine System Diseases

Gut microbiota is also closely related to the peripheral endocrine system, and microbial metabolites are reported to be involved in peripheral endocrine disease processes (Fig. 9.3). Relationship between gut microbiota and thyroid diseases The functional link between the gut and thyroid has long since been confirmed. T3 is theorized to impact the development and differentiation of intestinal mucosa and execute intestinal neuromotor function [89]. The gastrointestinal system participates in the enterohepatic recycling and metabolism of thyroid hormones [90, 91]. Autoimmune thyroid disease is the most common organ-specific autoimmune disease, affecting about 5% of the world’s population, with HT as the most prevalent [92, 93]. Although the factors that underlie the genesis of HT have yet to be completely elucidated, studies have established that its pathology derives from the interaction of a predisposing genotype with endogenous and environmental triggers [32]. Peripheral thyroid homeostasis may be sensitive to microbiota changes, and there is also evidence that the genesis and progression of autoimmune

9 Gut Microbiota and Endocrine Disorder

151

Fig. 9.3 Gut microbiota interferes with peripheral endocrine system diseases through multiple pathways

thyroid disorders may be significantly affected by changing intestinal microbial composition or overt dysbiosis [94]. The abundance of Prevotella and Dialister in the guts of HT patients is decreased, while elevated genera in these patients include Escherichia-Shigella and Parasutterella. The number of E. coli in the intestinal tract of HT patients is also increased. Gut microbiota can alter the development of primary and secondary lymphoid organs and the activation and differentiation of B cells associated with autoimmune thyroid diseases [95, 96]. At present, the link between microbiota and autoimmune thyroid disease is under scrutiny but has yet to be fully elucidated [97]. Studies have shown that hypothyroidism patients have a greater chance of developing bacterial overgrowth in the small intestine that can contribute to the impairment of gastrointestinal neuromuscular function [98]. Researchers analyzed fecal samples of hyperthyroid patients and discovered a significant decrease in Bifidobacterium and Lactobacillus and an increase in Enterococcus in these patients compared with the control group [94, 99]. The impact of gut microbiota on thyroid disease is not well studied and should be further explored and verified as to molecular mechanisms and clinical presentations. The relationship of gut microbiota with obesity, insulin resistance, and diabetes Changes in lifestyle and increased intake of calorie-rich foods are significant contributors to the global obesity epidemic [100]. When energy intake exceeds energy expenditure over the long term, adipose tissue cannot store the excess triglycerides, leading to increased lipid content in non-adipose tissues such as the liver and skeletal muscle. Excessive lipids lead to lipid peroxidation, impaired metabolic lipid function, and heterotopic lipid accumulation. As the ability to store lipid decreases, adipose tissue inflammation develops, resulting in increased production and secretion of pro-inflammatory adipokines that lead to the development of peripheral and hepatic insulin resistance implicated in T2DM and nonalcoholic fatty liver disease (NAFLD) [101–104].

152

R. Li et al.

Studies have implicated gut microbiota composition in the development of obesity which consider it as an environmental factor that contributes to obesity and its comorbidities, such as insulin resistance, diabetes, and cardiovascular disease [55, 105]. Gut microbiota have emerged as important regulators of energy and substrate metabolism. Abnormalities in gut microbiota composition and function can result in energy metabolism imbalances, including effects on adipose tissue, muscle, and liver metabolism [106, 107]. The gut microbiome is also associated with the development of chronic obesity-related inflammation, and gut microbiota metabolites such as SCFAs and succinic acid are vital to the prevention and treatment of obesity and its complications [108, 109]. A. Firmicutes/Bacteroidetes ratio The phyla Firmicutes and Bacteroidetes are the predominant members of gut microbiota. The product of Firmicutes butyrate is a principal energy source used by enterocytes, and the Bacteroidetes products acetate and propionate are flushed to the liver for lipogenesis and gluconeogenesis [110]. Various studies have reported changes to the intestinal Firmicutes/Bacteroidetes ratio associated with obesity. Ley et al. showed gut microbiota alterations in genetically obese (ob/ob) mice, observing an increase in Firmicutes and a decrease in Bacteroidetes [111]. However, whether the pathology of obesity is also associated with alterations to the Firmicutes/Bacteroidetes ratio in human obese patients remains a matter of debate [112–116]. Studies suggest that Firmicutes and Clostridium species are significantly reduced in the gut microbiota of diabetics and that increases of the Bacteroides/Firmicutes and Pseudobacillus/E. coli ratios were positively correlated with higher blood glucose levels irrespective of the patient’s weight [117]. Lambeth et al. found significantly increased b-Proteobacteria in patients with T2DM along with reductions of Bifidobacterium, Firmicutes, and Fusobacterium. Ratios of Bacteroides/Firmicutes and Brevibacterium/Clostridium sphaeroides were positively correlated with blood glucose levels, demonstrating that changes in the gut microbiota were closely related to reductions in glucose tolerance [118]. B. SCFAs and T2DM Researchers reported that gut microbiota acts as a mediator, rather than an initiator, of the physiological and pathological progression of obesity and T2DM [119]. SCFAs are formed by microbial fermentation and then absorbed and transferred to various tissues. The majority of them are used as an energy resource in peripheral tissues, although they also demonstrate certain specificities. While acetate and propionate are preferentially flushed to the liver and regulate cholesterol synthesis, butyrate preferentially enters enterocytes and is used for energy producing [120, 121]. Acetates enter the lipogenesis pathway, whereas propionate participates in the gluconeogenesis. All of these molecules contribute to adipogenesis [122, 123]. SCFAs are also ligands for G-protein-coupled receptors FFAR2 and FFAR3, the activation of which affects functions related to satiety and insulin sensitivity [124].

9 Gut Microbiota and Endocrine Disorder

153

Activation of FFAR2 in adipocytes stimulates them to release leptin [125]. It has also been reported that activating FFAR2 in EECs triggers the secretion of PYY [126], and SCFA-mediated FFAR3 activation can also elevate PYY serum levels [127]. Both leptin and PYY are appetite suppressors [128]. Acetate and propionate can also suppress appetite by enhancing the secretion of GLP-1, which stimulates the production of insulin in pancreatic b-cells and promotes insulin sensitivity and satiety [129, 130]. Receptor-knockout experiments in obese and insulin-resistant mice have demonstrated that SCFAs can increase glucose-stimulated insulin secretion via GPR43 [125, 131]. In line with this finding, another study found that the depletion of GPR43 in mice with diet-induced obesity was associated with deteriorated b-cell function and increased b-cell mass [132]. SCFA–GPR41 signaling appears to be of particular importance in controlling pancreatic b-cell insulin secretion in satiety and fasting states, as GPR41 knockout or overexpression in mice resulted in impairments in glucose control but did not affect insulin sensitivity [133]. Clinical studies have shown that dietary fibers could promote SCFA-producing bacteria growth. When fiber-promoted SCFA producers were present in greater diversity and abundance, participants had more significant improvements in hemoglobin A1c levels, partly via increased GLP-1 production. Thus, targeted restoration of these SCFA producers may present a novel ecological approach for managing T2DM [39]. C. Gut microbiota influences intestinal immunomodulatory cells to regulate autoimmune diabetes As mentioned above, gut microbiota mediate the relationship between intestinal immune cells (IICs) and endocrine diseases. IICs participate in the pathogenesis of gut microbiota and endocrine diseases, especially vital T lymphocytes, including Th1, Th2, Th17, and Tregs [134]. T cells differentiate to Th1 cells after exposure to IL-12 through the STAT4 pathway and to Th2 cells after exposure to IL-4 through the STAT6 pathway [135]. The main products secreted by Th1 cells are IL-2 and IFN-c; those of Th2 are IL-4 and IL-10. Notably, Bacteroidetes and its secretions, such as polysaccharide A, induce IL-12 expression, while commensal A4 bacteria from Lachnospiraceae may inhibit intestinal Th2 cell responses [136]. IFN-c participates in macrophage activation leading to cytotoxic activity that ultimately destroys islet b-cells. IL-12 binds to IL-12 receptors on islet b cells to activate the pro-inflammatory cytokines IL-1b, TNF-a, and INF-c; these, in turn, contribute to islet b cell apoptosis via the STAT4 pathway [137]. IL-4, the main product of Th2 cells, can help improve insulin sensitivity and glucose tolerance. IL-10 binds to IL-10 receptors on cell surface, and it has been shown that when IL-10 binds to these receptors on macrophages, it inhibits phagocytotic activity, while when IL-10 binds to glucose transporter receptors on skeletal muscle cells, it reduces inflammation and protects glycometabolic pathways [138, 139]. Scientists have found that Bacteroidetes and segmented filamentous bacteria induce a striking increase in Th17 responses in the small intestine [140]. Th17 cells mainly secrete IL-17, including IL-17A and IL-17F, thereby inducing

154

R. Li et al.

pro-inflammatory cytokines and chemokines. An increase in Th17 cells correlates with a concomitant increase in Th1 cells and can contribute to the development of DM by transferring thereof into Th1-like cells. Inhibition of Th17 cells has been shown to reduce islet-specific inflammatory T cell infiltration [141]. Another group of scientists found that benign intestinal symbiotic microbiota such as Schaedler flora can induce Treg cell differentiation and inhibit Th1 and Th17 cell responses. Short-chain fatty acids (SCFAs) can also promote the development of Tregs by activating GPR43. In some animal models, the combined production of CD4+, CD25+, FoxP3+, and Tregs can mitigate the destruction of pancreatic islets and protect against autoimmune DM [142]. D. Gut microbiota participate in bile acid metabolism Bile acids, the functional components of bile, are synthesized in hepatocytes from cholesterol. After conjugating to glycine and taurine, bile acids are secreted into the duodenum from the gallbladder after a meal [143]. About 95% of intestinal bile acids are actively reabsorbed by enterocytes in the terminal ileum, while 5% of bile acids escape from enterohepatic circulation and arrive in the large bowel where they are transformed into secondary bile acids under the action of intestinal flora [144– 146]. Small intestinal bile acids act as biological detergents for the solvation, digestion, and absorption of consumed lipids and lipophilic vitamins [147]. Bile acids are also crucial signaling molecules that regulate lipid, glucose, and energy metabolism via interaction with different intracellular ligand-activated nuclear receptors [147– 149]. Gut microbiota may trigger changes in weight and lipid metabolism through mechanisms mediated by the farnesoid-X-receptor (FXR) and cell surface-located G-protein-coupled bile acid receptor TGR5. According to a recent finding, gut microbiota facilitates high-fat-diet-induced obesity by modifying the bile acid profile and altering FXR signaling. FXR may also induce an obese phenotype by affecting the composition of gut microbiota [150, 151]. Bile salt hydrolase (BSH) in the human microbiome specifically catalyzes the hydrolysis of conjugated bile salts that are subsequently transformed into free bile acids. BSH is crucial for signaling pathways mediated by bile acids that regulate glucose metabolism and lipid absorption [152]. Joyce et al. demonstrated that gut microbiota could secrete BSH to influence adiposity and cholesterol levels in the host via processes that include lipid metabolism and transport, signaling pathways, hormone production, peripheral circadian rhythms, gut homeostasis and barrier function [153]. Li et al. indicated that Tempol treatment could alter the composition of intestinal flora and result in lower BSH activity in mice. Specifically, the administration of Tempol was associated with the transformation from Firmicutes to Bacteroidetes at the phylum level and decreases in Lactobacillus. Reduction in BSH activity could potentially be directly linked to higher levels of taurine-conjugated bile acids, including tauro b-muricholic acid (T b-MCA), taurocholic acid (TCA), and taurochenodexycholic acid (TCDCA) in the intestine. An increase in taurine-conjugated bile acids contributes to the inhibition of the FXR signaling pathway, through which Tempol may alleviate obesity and insulin resistance in mice [119].

9 Gut Microbiota and Endocrine Disorder

155

Recent clinical metagenomic and metabolomic analyses showed that samples from individuals with newly diagnosed T2DM naively treated with metformin for 3 d demonstrated a decrease in Bacteroides fragilis and an increase in the bile acid glycoursodeoxycholic acid (GUDCA) in the gut. These changes were accompanied by the inhibition of FXR signaling. B. fragilis administration augmented HFD-induced glucose intolerance, and also diminished the metabolic benefits of metformin. Furthermore, GUDCA was identified as an intestinal FXR antagonist and was proven to be benefit for metabolic disorder in animals. Thus, metformin acts in part through a B. fragilis–GUDCA–intestinal FXR axis to improve metabolic dysfunction, including hyperglycemia [154]. Relationship between gut microbiota and hyperuricemia Gut microbiota are commonly believed to participate in UA metabolism. Recent evidence indicates that gut microbiota may be associated with increasing levels of UA and the development of hyperuricemia. Intestinal flora in individuals with gout have been found to be quite different from those in healthy individuals, with higher levels of Bacteroides caccaeand and Bacteroides xylanisolvens and lower levels of Faecalibacterium prausnitzii and Bifidobacterium pseudocatenulatum [155]. Hyperuricemia treatment has also been associated with changes in gut microbiota. Treatment with allopurinol and benzbromarone was shown to decrease levels of UA and was observed to alter gut microbiota in rats at the phylum and genera levels, with increasing abundances of Bifidobacterium and Collinsella and reductions of Adlercreutzia and Anaerostipes, Yu et al. suggested that the mechanisms by which allopurinol and benzbromarone decrease UA may be correlated with gut microbiota changes [156]. Garcia-Arroyo et al. showed that the administration of probiotic bacteria could accelerate UA secretion and reduce concentrations of serum UA, suggesting another possible treatment for hyperuricemia [157].

9.4

Conclusions and Potential Therapies

A great deal of research has produced strong evidence that gut microbiota and their metabolites are important for the stability of the endocrine system and the progression of diseases such as diabetes, hyperthyroidism, and hyperuricemia. Recent research also indicates that fecal bacteria transplantation (FBT) and probiotic intervention hold promise for the treatment of these endocrine diseases. Specifically, endocrine disorders, such as obesity and diabetes, can be addressed by adding normal, beneficial bacteria, or microbial metabolites to patients’ diets. Probiotics have demonstrated its anti-inflammatory, hypoglycemic, insulinotropic, antioxidative, and satietogenic properties and could be employed in the treatment of T2DM and obesity. Most available probiotics contain bifidobacteria, lactic acid bacteria (LAB), dairy propionibacteria, yeasts (Saccharomyces boulardii), Bacillus, and the Gram-negative E. coli strain Nissle [158]. Probiotic administration

156

R. Li et al.

has been associated with immune response modulation, improved lactose tolerance, prevention of diarrhea, anti-inflammatory effects, and even restorations of obesity-linked gut dysbiosis. Probiotic supplementation may also help reduce hyperphagia and improve control over weight gain, fat mass loss, and glucose tolerance [159]. Probiotics have been shown to increase the production of tight junction and adherens junction proteins, arresting gut permeability and inhibiting the passage of LPS into systemic circulation, and thereby decreasing metabolic endotoxemia. Probiotics have been investigated as a beneficial dietary supplement, positively contributing to endocrine system homeostasis. Interestingly, a growing body of research suggests that it is not the probiotics or other microbiota themselves that ameliorate endocrine pathologies, but rather the post-biologics produced by these gut microbiota. Whichever may be the case, a causal relationship between gut microbiota metabolites and endocrine disorders has yet to be fully elucidated. The function of certain metabolites, such as succinic acid and ethanol, is particularly unclear. However, manipulating microbial substrates can increase the complement of beneficial intestinal flora metabolites, and the focus on screening and identification thereof may yield new strategies for the prevention or adjuvant treatment of metabolic diseases in the future.

References 1. Gentile CL, Weir TL (2018) The gut microbiota at the intersection of diet and human health. Science 362:776–780 2. Rooks MG, Garrett WS (2016) Gut microbiota, metabolites and host immunity. Nat Rev Immunol 16:341–352 3. Kamada N, Seo SU, Chen GY, Nunez G (2013) Role of the gut microbiota in immunity and inflammatory disease. Nat Rev Immunol 13:321–335 4. Baumler AJ, Sperandio V (2016) Interactions between the microbiota and pathogenic bacteria in the gut. Nature 535:85–93 5. Rook G, Backhed F, Levin BR, McFall-Ngai MJ, McLean AR (2017) Evolution, human-microbe interactions, and life history plasticity. Lancet 390:521–530 6. Lozupone CA, Stombaugh JI, Gordon JI, Jansson JK, Knight R (2012) Diversity, stability and resilience of the human gut microbiota. Nature 489:220–230 7. Tremaroli V, Backhed F (2012) Functional interactions between the gut microbiota and host metabolism. Nature 489:242–249 8. Browne HP, Neville BA, Forster SC, Lawley TD (2017) Transmission of the gut microbiota: spreading of health. Nat Rev Microbiol 15:531–543 9. Postler TS, Ghosh S (2017) Understanding the holobiont: how microbial metabolites affect human health and shape the immune system. Cell Metab 26:110–130 10. Lagier JC, Dubourg G, Million M, Cadoret F, Bilen M, Fenollar F, Levasseur A, Rolain JM, Fournier PE, Raoult D (2018) Culturing the human microbiota and culturomics. Nat Rev Microbiol 540–550 11. Mao K, Baptista AP, Tamoutounour S, Zhuang L, Bouladoux N, Martins AJ, Huang Y, Gerner MY, Belkaid Y, Germain RN (2018) Innate and adaptive lymphocytes sequentially shape the gut microbiota and lipid metabolism. Nature 554:255–259 12. Ouchi N, Parker JL, Lugus JJ, Walsh K (2011) Adipokines in inflammation and metabolic disease. Nat Rev Immunol 11:85–97

9 Gut Microbiota and Endocrine Disorder

157

13. Newey PJ, Gorvin CM, Cleland SJ, Willberg CB, Bridge M, Azharuddin M, Drummond RS, van der Merwe PA, Klenerman P, Bountra C, Thakker RV (2013) Mutant prolactin receptor and familial hyperprolactinemia. N Engl J Med 369:2012–2020 14. Newell-Price J (2016) Pituitary gland: mortality in cushing disease. Nat Rev Endocr 12:502– 503 15. Coon SL, Munson PJ, Cherukuri PF, Sugden D, Rath MF, Moller M, Clokie SJ, Fu C, Olanich ME, Rangel Z, Werner T, Program NCS, Mullikin JC, Klein DC (2012) Circadian changes in long noncoding RNAs in the pineal gland. Proc Natl Acad Sci USA 109:13319– 13324 16. Hatori M, Hirota T, Iitsuka M, Kurabayashi N, Haraguchi S, Kokame K, Sato R, Nakai A, Miyata T, Tsutsui K, Fukada Y (2011) Light-dependent and circadian clock-regulated activation of sterol regulatory element-binding protein, X-box-binding protein 1, and heat shock factor pathways. Proc Natl Acad Sci USA 108:4864–4869 17. Hardeland R, Cardinali DP, Srinivasan V, Spence DW, Brown GM, Pandi-Perumal SR (2011) Melatonin–a pleiotropic, orchestrating regulator molecule. Prog Neurobiol 93:350– 384 18. De Leo S, Lee SY, Braverman LE (2016) Hyperthyroidism. Lancet 388:906–918 19. Lopez M, Varela L, Vazquez MJ, Rodriguez-Cuenca S, Gonzalez CR, Velagapudi VR, Morgan DA, Schoenmakers E, Agassandian K, Lage R, Martinez de Morentin PB, Tovar S, Nogueiras R, Carling D, Lelliott C, Gallego R, Oresic M, Chatterjee K, Saha AK, Rahmouni K, Dieguez C, Vidal-Puig A (2010) Hypothalamic AMPK and fatty acid metabolism mediate thyroid regulation of energy balance. Nat Med 16:1001–1008 20. Mullur R, Liu YY, Brent GA (2014) Thyroid hormone regulation of metabolism. Physiol Rev 94:355–382 21. Vidal V, Sacco S, Rocha AS, da Silva F, Panzolini C, Dumontet T, Doan TM, Shan J, Rak-Raszewska A, Bird T, Vainio S, Martinez A, Schedl A (2016) The adrenal capsule is a signaling center controlling cell renewal and zonation through Rspo3. Genes Dev 30:1389– 1394 22. Kohrle J, Jakob F, Contempre B, Dumont JE (2005) Selenium, the thyroid, and the endocrine system. Endocr Rev 26:944–984 23. Kim T, Loh YP (2005) Chromogranin A: a surprising link between granule biogenesis and hypertension. J Clin Investig 115:1711–1713 24. Ying W, Riopel M, Bandyopadhyay G, Dong Y, Birmingham A, Seo JB, Ofrecio JM, Wollam J, Hernandez-Carretero A, Fu W, Li P, Olefsky JM (2017) Adipose tissue macrophage-derived exosomal miRNAs can modulate in vivo and in vitro insulin sensitivity. Cell 171:372–384, e312 25. Trayhurn P (2013) Hypoxia and adipose tissue function and dysfunction in obesity. Physiol Rev 93:1–21 26. Wang X, Ota N, Manzanillo P, Kates L, Zavala-Solorio J, Eidenschenk C, Zhang J, Lesch J, Lee WP, Ross J, Diehl L, van Bruggen N, Kolumam G, Ouyang W (2014) Interleukin-22 alleviates metabolic disorders and restores mucosal immunity in diabetes. Nature 514:237– 241 27. Khoshi A, Bajestani MK, Shakeri H, Goodarzi G, Azizi F (2019) Association of Omentin rs2274907 and FTO rs9939609 gene polymorphisms with insulin resistance in Iranian individuals with newly diagnosed type 2 diabetes. Lipids Health Dis 18:142 28. Gray SM, Niu J, Zhang A, Svendsen B, Campbell JE, D’Alessio DA, Tong J (2019) Intra-Islet ghrelin signaling does not regulate insulin secretion from adult mice. Diabetes 68 (9):1795–1805 29. Zhou M, Jing S, Wu CY, Shu L, Dong W, Liu Y, Chen M, Wynn RM, Wang J, Wang J, Gui WJ, Qi X, Lusis AJ, Li Z, Wang W, Ning G, Yang X, Chuang DT, Wang Y, Sun H (2019) Targeting BCAA catabolism to treat obesity-associated insulin resistance. Diabetes 68(9):1730–1746

158

R. Li et al.

30. Schirra J, Nicolaus M, Roggel R, Katschinski M, Storr M, Woerle HJ, Goke B (2006) Endogenous glucagon-like peptide 1 controls endocrine pancreatic secretion and antro-pyloro-duodenal motility in humans. Gut 55:243–251 31. Chen PC, Inui A, Chen CY (2015) Paradoxical responses of plasma glucagon and bile acid levels after duodenal nutrient exclusion. Gut 64:516 32. Antonelli A, Ferrari SM, Corrado A, Di Domenicantonio A, Fallahi P (2015) Autoimmune thyroid disorders. Autoimmun Rev 14:174–180 33. Nieman LK, Ilias I (2005) Evaluation and treatment of Cushing’s syndrome. Am J Med 118:1340–1346 34. Lodish M, Stratakis CA (2016) A genetic and molecular update on adrenocortical causes of Cushing syndrome. Nat Rev Endocr 12:255–262 35. Arregger AL, Cardoso EM, Zucchini A, Aguirre EC, Elbert A, Contreras LN (2014) Adrenocortical function in hypotensive patients with end stage renal disease. Steroids 84:57– 63 36. de Loos WS (1996) Clinical problem-solving: identifying Addison’s disease. N Engl J Med 334:1403; author reply, 1404–1405 37. Matsumoto S, Hagiwara S, Kusaka J, Hasegawa R, Nonaka H, Shingu C, Noguchi T (2011) Catecholamine-resistant shock and hypoglycemic coma after cardiotomy in a patient with unexpected isolated ACTH deficiency. J Anesth 25:431–434 38. Savage DB, Petersen KF, Shulman GI (2007) Disordered lipid metabolism and the pathogenesis of insulin resistance. Physiol Rev 87:507–520 39. Zhao L, Zhang F, Ding X, Wu G, Lam YY, Wang X, Fu H, Xue X, Lu C, Ma J, Yu L, Xu C, Ren Z, Xu Y, Xu S, Shen H, Zhu X, Shi Y, Shen Q, Dong W, Liu R, Ling Y, Zeng Y, Wang X, Zhang Q, Wang J, Wang L, Wu Y, Zeng B, Wei H, Zhang M, Peng Y, Zhang C (2018) Gut bacteria selectively promoted by dietary fibers alleviate type 2 diabetes. Science 359:1151–1156 40. Yang Q, Vijayakumar A, Kahn BB (2018) Metabolites as regulators of insulin sensitivity and metabolism. Nat Rev Mol Cell Biol 19:654–672 41. Musso G, Cassader M, De Michieli F, Rosina F, Orlandi F, Gambino R (2012) Nonalcoholic steatohepatitis versus steatosis: adipose tissue insulin resistance and dysfunctional response to fat ingestion predict liver injury and altered glucose and lipoprotein metabolism. Hepatology 56:933–942 42. Johannsen DL, Tchoukalova Y, Tam CS, Covington JD, Xie W, Schwarz J-M, Bajpeyi S, Ravussin E (2014) Effect of 8 weeks of overfeeding on ectopic fat deposition and insulin sensitivity: testing the “Adipose Tissue Expandability” hypothesis. Diabetes Care 37:2789– 2797 43. Buki KG, Bauer PI, Kun E (1997) Isolation and identification of a proteinase from calf thymus that cleaves poly(ADP-ribose) polymerase and histone H1. Biochem Biophys Acta 1338:100–106 44. So A, Thorens B (2010) Uric acid transport and disease. J Clin Investig 120:1791–1799 45. Carnovale C, Venegoni M, Clementi E (2014) Allopurinol overuse in asymptomatic hyperuricemia: a teachable moment. JAMA Intern Med 174:1031–1032 46. Jeon HJ, Oh J, Shin DH (2019) Urate-lowering agents for asymptomatic hyperuricemia in stage 3–4 chronic kidney disease: controversial role of kidney function. PLoS ONE 14: e0218510 47. El Aidy S, Dinan TG, Cryan JF (2015) Gut microbiota: the conductor in the orchestra of immune-neuroendocrine communication. Clin Ther 37:954–967 48. Rieder R, Wisniewski PJ, Alderman BL, Campbell SC (2017) Microbes and mental health: a review. Brain Behav Immun 66:9–17 49. Martin-Villa JM (2014) Neuroendocrine stimulation of mucosal immune cells in inflammatory bowel disease. Curr Pharm Des 20:4766–4773 50. Caputi V, Giron MC (2018) Microbiome-gut-brain axis and toll-like receptors in Parkinson’s disease. Int J Mol Sci 19

9 Gut Microbiota and Endocrine Disorder

159

51. Farzi A, Frohlich EE, Holzer P (2018) Gut microbiota and the neuroendocrine system. Neurotherapeutics 15:5–22 52. Roman P, Rueda-Ruzafa L, Cardona D, Cortes-Rodriguez A (2018) Gut-brain axis in the executive function of austism spectrum disorder. Behav Pharmacol 29:654–663 53. Hooks KB, Konsman JP, O’Malley MA (2018) Microbiota-gut-brain research: a critical analysis. Behav Brain Sci 1–40 54. Gerhardt S, Mohajeri MH (2018) Changes of colonic bacterial composition in Parkinson’s disease and other neurodegenerative diseases. Nutrients 10 55. Moran CP, Shanahan F (2014) Gut microbiota and obesity: role in aetiology and potential therapeutic target. Best Pract Res Clin Gastroenterol 28:585–597 56. Petra AI, Panagiotidou S, Hatziagelaki E, Stewart JM, Conti P, Theoharides TC (2015) Gut-microbiota-brain axis and its effect on neuropsychiatric disorders with suspected immune dysregulation. Clin Ther 37:984–995 57. Fung TC, Olson CA, Hsiao EY (2017) Interactions between the microbiota, immune and nervous systems in health and disease. Nat Neurosci 20:145–155 58. Kennedy PJ, Cryan JF, Dinan TG, Clarke G (2017) Kynurenine pathway metabolism and the microbiota-gut-brain axis. Neuropharmacology 112:399–412 59. Bhattarai Y (2018) Microbiota-gut-brain axis: interaction of gut microbes and their metabolites with host epithelial barriers. Neurogastroenterol Motil 30:e13366 60. Koh A, De Vadder F, Kovatcheva-Datchary P, Backhed F (2016) From dietary fiber to host physiology: short-chain fatty acids as key bacterial metabolites. Cell 165:1332–1345 61. Bercik P (2011) The microbiota-gut-brain axis: learning from intestinal bacteria? Gut 60:288–289 62. Nankova BB, Agarwal R, MacFabe DF, La Gamma EF (2014) Enteric bacterial metabolites propionic and butyric acid modulate gene expression, including CREB-dependent catecholaminergic neurotransmission, in PC12 cells–possible relevance to autism spectrum disorders. PLoS ONE 9:e103740 63. Zhu H, Huang Q, Xu H, Niu L, Zhou JN (2009) Antidepressant-like effects of sodium butyrate in combination with estrogen in rat forced swimming test: involvement of 5-HT (1A) receptors. Behav Brain Res 196:200–206 64. Erny D, Hrabe de Angelis AL, Jaitin D, Wieghofer P, Staszewski O, David E, Keren-Shaul H, Mahlakoiv T, Jakobshagen K, Buch T, Schwierzeck V, Utermohlen O, Chun E, Garrett WS, McCoy KD, Diefenbach A, Staeheli P, Stecher B, Amit I, Prinz M (2015) Host microbiota constantly control maturation and function of microglia in the CNS. Nat Neurosci 18:965–977 65. Lund ML, Egerod KL, Engelstoft MS, Dmytriyeva O, Theodorsson E, Patel BA, Schwartz TW (2018) Enterochromaffin 5-HT cells—A major target for GLP-1 and gut microbial metabolites. Mol Metab 11:70–83 66. Rastelli M, Cani PD, Knauf C (2019) The gut microbiome influences host endocrine functions. Endocr Rev 40(5):1271–1284 67. Sun MF, Zhu YL, Zhou ZL, Jia XB, Xu YD, Yang Q, Cui C, Shen YQ (2018) Neuroprotective effects of fecal microbiota transplantation on MPTP-induced Parkinson’s disease mice: gut microbiota, glial reaction and TLR4/TNF-alpha signaling pathway. Brain Behav Immun 70:48–60 68. Huuskonen J, Suuronen T, Nuutinen T, Kyrylenko S, Salminen A (2004) Regulation of microglial inflammatory response by sodium butyrate and short-chain fatty acids. Br J Pharmacol 141:874–880 69. Das S, Sreevidya VS, Udvadia AJ, Gyaneshwar P (2019) Dopamine-induced sulfatase and its regulator are required for Salmonella enterica serovar Typhimurium pathogenesis. Microbiology 165:302–310 70. Strandwitz P (2018) Neurotransmitter modulation by the gut microbiota. Brain Res 1693:128–133

160

R. Li et al.

71. Cogan TA, Thomas AO, Rees LE, Taylor AH, Jepson MA, Williams PH, Ketley J, Humphrey TJ (2007) Norepinephrine increases the pathogenic potential of Campylobacter jejuni. Gut 56:1060–1065 72. Hyland NP, Cryan JF (2010) A gut feeling about GABA: focus on GABA(B) receptors. Front Pharmacol 1:124 73. Mazzoli R, Pessione E (2016) The neuro-endocrinological role of microbial glutamate and GABA signaling. Front Microbiol 7:1934 74. van Kessel SP, Frye AK, El-Gendy AO, Castejon M, Keshavarzian A, van Dijk G, El Aidy S (2019) Gut bacterial tyrosine decarboxylases restrict levels of levodopa in the treatment of Parkinson’s disease. Nat Commun 10:310 75. Yunes RA, Poluektova EU, Dyachkova MS, Klimina KM, Kovtun AS, Averina OV, Orlova VS, Danilenko VN (2016) GABA production and structure of gadB/gadC genes in Lactobacillus and Bifidobacterium strains from human microbiota. Anaerobe 42:197–204 76. Barrett E, Ross RP, O’Toole PW, Fitzgerald GF, Stanton C (2012) gamma-Aminobutyric acid production by culturable bacteria from the human intestine. J Appl Microbiol 113:411– 417 77. Polovinkin L, Hassaine G, Perot J, Neumann E, Jensen AA, Lefebvre SN, Corringer PJ, Neyton J, Chipot C, Dehez F, Schoehn G, Nury H (2018) Conformational transitions of the serotonin 5-HT3 receptor. Nature 563:275–279 78. Yano JM, Yu K, Donaldson GP, Shastri GG, Ann P, Ma L, Nagler CR, Ismagilov RF, Mazmanian SK, Hsiao EY (2015) Indigenous bacteria from the gut microbiota regulate host serotonin biosynthesis. Cell 161:264–276 79. Israelyan N, Del Colle A, Li Z, Park Y, Xing A, Jacobsen JPR, Luna RA, Jensen DD, Madra M, Saurman V, Rahim R, Latorre R, Law K, Carson W, Bunnett NW, Caron MG, Margolis KG (2019) Effects of serotonin and slow-release 5-HTP on gastrointestinal motility in a mouse model of depression. Gastroenterology 157(2) 80. Schneider S, Wright CM, Heuckeroth RO (2019) Unexpected roles for the second brain: enteric nervous system as master regulator of bowel function. Annu Rev Physiol 81:235–259 81. Bliss ES, Whiteside E (2018) The gut-brain axis, the human gut microbiota and their integration in the development of obesity. Front Physiol 9:900 82. Wu Y, He H, Cheng Z, Bai Y, Ma X (2019) The role of neuropeptide Y and peptide YY in the development of obesity via gut-brain axis. Curr Protein Pept Sci 20(7):750–758 83. Zhang X, Grosfeld A, Williams E, Vasiliauskas D, Barretto S, Smith L, Mariadassou M, Philippe C, Devime F, Melchior C, Gourcerol G, Dourmap N, Lapaque N, Larraufie P, Blottiere HM, Herberden C, Gerard P, Rehfeld JF, Ferraris RP, Fritton JC, Ellero-Simatos S, Douard V (2019) Fructose malabsorption induces cholecystokinin expression in the ileum and cecum by changing microbiota composition and metabolism. FASEB J (Official publication of the Federation of American Societies for Experimental Biology) 33:7126– 7142 84. Ripken D, van der Wielen N, Wortelboer HM, Meijerink J, Witkamp RF, Hendriks HF (2016) Nutrient-induced glucagon like peptide-1 release is modulated by serotonin. J Nutr Biochem 32:142–150 85. de Clercq NC, Frissen MN, Groen AK, Nieuwdorp M (2017) Gut microbiota and the gut-brain axis: new insights in the pathophysiology of metabolic syndrome. Psychosom Med 79:874–879 86. Lach G, Schellekens H, Dinan TG, Cryan JF (2018) Anxiety, depression, and the microbiome: a role for gut peptides. Neurotherapeutics 15:36–59 87. Klingbeil E, de La Serre CB (2018) Microbiota modulation by eating patterns and diet composition: impact on food intake. Am J Physiol Regul Integr Comp Physiol 315:R1254– R1260 88. Bravo JA, Forsythe P, Chew MV, Escaravage E, Savignac HM, Dinan TG, Bienenstock J, Cryan JF (2011) Ingestion of Lactobacillus strain regulates emotional behavior and central GABA receptor expression in a mouse via the vagus nerve. Proc Natl Acad Sci USA 108:16050–16055

9 Gut Microbiota and Endocrine Disorder

161

89. Meng S, Badrinarain J, Sibley E, Fang R, Hodin R (2001) Thyroid hormone and the d-type cyclins interact in regulating enterocyte gene transcription. J Gastrointest Surg 5:49–55 90. Virili C, Centanni M (2017) “With a little help from my friends” - The role of microbiota in thyroid hormone metabolism and enterohepatic recycling. Mol Cell Endocrinol 458:39–43 91. Patil AD (2014) Link between hypothyroidism and small intestinal bacterial overgrowth. Indian J Endocrinol Metab 18:307–309 92. Biondi B, Cappola AR, Cooper DS (2019) Subclinical hypothyroidism: a review. JAMA 322:153–160 93. Jacobson DL, Gange SJ, Rose NR, Graham NM (1997) Epidemiology and estimated population burden of selected autoimmune diseases in the United States. Clin Immunol Immunopathol 84:223–243 94. Virili C, Fallahi P, Antonelli A, Benvenga S, Centanni M (2018) Gut microbiota and Hashimoto’s thyroiditis. Rev Endocr Metab Disord 19:293–300 95. Alenghat T, Artis D (2014) Epigenomic regulation of host-microbiota interactions. Trends Immunol 35:518–525 96. Hamer HM, Jonkers D, Venema K, Vanhoutvin S, Troost FJ, Brummer RJ (2008) Review article: the role of butyrate on colonic function. Aliment Pharmacol Ther 27:104–119 97. Ishaq HM, Mohammad IS, Guo H, Shahzad M, Hou YJ, Ma C, Naseem Z, Wu X, Shi P, Xu J (2017) Molecular estimation of alteration in intestinal microbial composition in Hashimoto’s thyroiditis patients. Biomed Pharmacother 95:865–874 98. Lauritano EC, Bilotta AL, Gabrielli M, Scarpellini E, Lupascu A, Laginestra A, Novi M, Sottili S, Serricchio M, Cammarota G, Gasbarrini G, Pontecorvi A, Gasbarrini A (2007) Association between hypothyroidism and small intestinal bacterial overgrowth. J Clin Endocrinol Metab 92:4180–4184 99. Zhou L, Li X, Ahmed A, Wu D, Liu L, Qiu J, Yan Y, Jin M, Xin Y (2014) Gut microbe analysis between hyperthyroid and healthy individuals. Curr Microbiol 69:675–680 100. Sun K, Kusminski CM, Scherer PE (2011) Adipose tissue remodeling and obesity. J Clin Investig 121:2094–2101 101. Scheja L, Heeren J (2016) Metabolic interplay between white, beige, brown adipocytes and the liver. J Hepatol 64:1176–1186 102. Teperino R, Amann S, Bayer M, McGee SL, Loipetzberger A, Connor T, Jaeger C, Kammerer B, Winter L, Wiche G, Dalgaard K, Selvaraj M, Gaster M, Lee-Young RS, Febbraio MA, Knauf C, Cani PD, Aberger F, Penninger JM, Pospisilik JA, Esterbauer H (2012) Hedgehog partial agonism drives warburg-like metabolism in muscle and brown fat. Cell 151:414–426 103. Shimobayashi M, Albert V, Woelnerhanssen B, Frei IC, Weissenberger D, Meyer-Gerspach AC, Clement N, Moes S, Colombi M, Meier JA, Swierczynska MM, Jeno P, Beglinger C, Peterli R, Hall MN (2018) Insulin resistance causes inflammation in adipose tissue. J Clin Investig 128:1538–1550 104. Qatanani M, Szwergold NR, Greaves DR, Ahima RS, Lazar MA (2009) Macrophage-derived human resistin exacerbates adipose tissue inflammation and insulin resistance in mice. J Clin Investig 119:531–539 105. Sonnenburg JL, Backhed F (2016) Diet-microbiota interactions as moderators of human metabolism. Nature 535:56–64 106. Zhu Z, Zhu B, Sun Y, Ai C, Wang L, Wen C, Yang J, Song S, Liu X (2018) Sulfated polysaccharide from sea cucumber and its depolymerized derivative prevent obesity in association with modification of gut microbiota in high-fat diet-fed mice. Mol Nutr Food Res 62:e1800446 107. Zhu A, Chen J, Wu P, Luo M, Zeng Y, Liu Y, Zheng H, Zhang L, Chen Z, Sun Q, Li W, Duan Y, Su D, Xiao Z, Duan Z, Zheng S, Bai L, Zhang X, Ju Z, Li Y, Hu R, Pandol SJ, Han YP (2017) Cationic polystyrene resolves nonalcoholic steatohepatitis, obesity, and metabolic disorders by promoting eubiosis of gut microbiota and decreasing endotoxemia. Diabetes 66:2137–2143

162

R. Li et al.

108. Zhuang P, Shou Q, Lu Y, Wang G, Qiu J, Wang J, He L, Chen J, Jiao J, Zhang Y (2017) Arachidonic acid sex-dependently affects obesity through linking gut microbiota-driven inflammation to hypothalamus-adipose-liver axis. Biochimica et Biophysica Acta 1863 (11):2715–2726 109. Zou J, Chassaing B, Singh V, Pellizzon M, Ricci M, Fythe MD, Kumar MV, Gewirtz AT (2018) Fiber-mediated nourishment of gut microbiota protects against diet-induced obesity by restoring IL-22-mediated colonic health. Cell Host Microbe 23:41–53, e44 110. Flint HJ, Bayer EA, Rincon MT, Lamed R, White BA (2008) Polysaccharide utilization by gut bacteria: potential for new insights from genomic analysis. Nat Rev Microbiol 6:121– 131 111. Ley RE, Backhed F, Turnbaugh P, Lozupone CA, Knight RD, Gordon JI (2005) Obesity alters gut microbial ecology. Proc Natl Acad Sci USA 102:11070–11075 112. Gao R, Zhu C, Li H, Yin M, Pan C, Huang L, Kong C, Wang X, Zhang Y, Qu S, Qin H (2018) Dysbiosis signatures of gut microbiota along the sequence from healthy, young patients to those with overweight and obesity. Obesity (Silver Spring) 26:351–361 113. Larsen N, Vogensen FK, van den Berg FW, Nielsen DS, Andreasen AS, Pedersen BK, Al-Soud WA, Sorensen SJ, Hansen LH, Jakobsen M (2010) Gut microbiota in human adults with type 2 diabetes differs from non-diabetic adults. PLoS ONE 5:e9085 114. Wu X, Ma C, Han L, Nawaz M, Gao F, Zhang X, Yu P, Zhao C, Li L, Zhou A, Wang J, Moore JE, Millar BC, Xu J (2010) Molecular characterisation of the faecal microbiota in patients with type II diabetes. Curr Microbiol 61:69–78 115. Mai V, McCrary QM, Sinha R, Glei M (2009) Associations between dietary habits and body mass index with gut microbiota composition and fecal water genotoxicity: an observational study in African American and Caucasian American volunteers. Nutr J 8:49 116. Turnbaugh PJ, Hamady M, Yatsunenko T, Cantarel BL, Duncan A, Ley RE, Sogin ML, Jones WJ, Roe BA, Affourtit JP, Egholm M, Henrissat B, Heath AC, Knight R, Gordon JI (2009) A core gut microbiome in obese and lean twins. Nature 457:480–484 117. Brown CT, Davis-Richardson AG, Giongo A, Gano KA, Crabb DB, Mukherjee N, Casella G, Drew JC, Ilonen J, Knip M, Hyoty H, Veijola R, Simell T, Simell O, Neu J, Wasserfall CH, Schatz D, Atkinson MA, Triplett EW (2011) Gut microbiome metagenomics analysis suggests a functional model for the development of autoimmunity for type 1 diabetes. PLoS ONE 6:e25792 118. Lambeth SM, Carson T, Lowe J, Ramaraj T, Leff JW, Luo L, Bell CJ, Shah VO (2015) Composition, diversity and abundance of gut microbiome in prediabetes and type 2 diabetes. J Diabetes Obes 2:1–7 119. Knip M, Siljander H (2016) The role of the intestinal microbiota in type 1 diabetes mellitus. Nat Rev Endoc 12:154–167 120. Yamaguchi Y, Adachi K, Sugiyama T, Shimozato A, Ebi M, Ogasawara N, Funaki Y, Goto C, Sasaki M, Kasugai K (2016) Association of intestinal microbiota with metabolic markers and dietary habits in patients with type 2 diabetes. Digestion 94:66–72 121. Sabatino A, Regolisti G, Cosola C, Gesualdo L, Fiaccadori E (2017) Intestinal microbiota in type 2 diabetes and chronic kidney disease. Curr Diab Rep 17:16 122. Li G, Xie C, Lu S, Nichols RG, Tian Y, Li L, Patel D, Ma Y, Brocker CN, Yan T, Krausz KW, Xiang R, Gavrilova O, Patterson AD, Gonzalez FJ (2017) Intermittent fasting promotes white adipose browning and decreases obesity by shaping the gut microbiota. Cell Metab 26:672–685, e674 123. Trent CM, Blaser MJ (2016) Microbially produced acetate: a “Missing Link” in understanding obesity? Cell Metab 24:9–10 124. Brown AJ, Goldsworthy SM, Barnes AA, Eilert MM, Tcheang L, Daniels D, Muir AI, Wigglesworth MJ, Kinghorn I, Fraser NJ, Pike NB, Strum JC, Steplewski KM, Murdock PR, Holder JC, Marshall FH, Szekeres PG, Wilson S, Ignar DM, Foord SM, Wise A, Dowell SJ (2003) The Orphan G protein-coupled receptors GPR41 and GPR43 are activated by propionate and other short chain carboxylic acids. J Biol Chem 278:11312– 11319

9 Gut Microbiota and Endocrine Disorder

163

125. Kim MH, Kang SG, Park JH, Yanagisawa M, Kim CH (2013) Short-chain fatty acids activate GPR41 and GPR43 on intestinal epithelial cells to promote inflammatory responses in mice. Gastroenterology 145:396–406, e391-310 126. Forbes S, Stafford S, Coope G, Heffron H, Real K, Newman R, Davenport R, Barnes M, Grosse J, Cox H (2015) Selective FFA2 agonism appears to act via intestinal PYY to reduce transit and food intake but does not improve glucose tolerance in mouse models. Diabetes 64:3763–3771 127. Remely M, Aumueller E, Merold C, Dworzak S, Hippe B, Zanner J, Pointner A, Brath H, Haslberger AG (2014) Effects of short chain fatty acid producing bacteria on epigenetic regulation of FFAR3 in type 2 diabetes and obesity. Gene 537:85–92 128. Iepsen EW, Lundgren J, Dirksen C, Jensen JE, Pedersen O, Hansen T, Madsbad S, Holst JJ, Torekov SS (2015) Treatment with a GLP-1 receptor agonist diminishes the decrease in free plasma leptin during maintenance of weight loss. Int J Obes 39:834–841 129. Christiansen CB, Gabe MBN, Svendsen B, Dragsted LO, Rosenkilde MM, Holst JJ (2018) The impact of short-chain fatty acids on GLP-1 and PYY secretion from the isolated perfused rat colon. Am J Physiol Gastrointest Liver Physiol 315:G53–G65 130. Chandarana K, Gelegen C, Karra E, Choudhury AI, Drew ME, Fauveau V, Viollet B, Andreelli F, Withers DJ, Batterham RL (2011) Diet and gastrointestinal bypass-induced weight loss: the roles of ghrelin and peptide YY. Diabetes 60:810–818 131. Maslowski KM, Vieira AT, Ng A, Kranich J, Sierro F, Yu D, Schilter HC, Rolph MS, Mackay F, Artis D, Xavier RJ, Teixeira MM, Mackay CR (2009) Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature 461:1282–1286 132. Bolognini D, Tobin AB, Milligan G, Moss CE (2016) The pharmacology and function of receptors for short-chain fatty acids. Mol Pharmacol 89:388–398 133. Inoue D, Tsujimoto G, Kimura I (2014) Regulation of energy homeostasis by GPR41. Front Endocrinol (Lausanne) 5:81 134. Tanoue T, Atarashi K, Honda K (2016) Development and maintenance of intestinal regulatory T cells. Nat Rev Immunol 16:295–309 135. Neurath MF, Finotto S, Glimcher LH (2002) The role of Th1/Th2 polarization in mucosal immunity. Nat Med 8:567–573 136. Wu W, Liu HP, Chen F, Liu H, Cao AT, Yao S, Sun M, Evans-Marin HL, Zhao Y, Zhao Q, Duck LW, Elson CO, Liu Z, Cong Y (2016) Commensal A4 bacteria inhibit intestinal Th2-cell responses through induction of dendritic cell TGF-beta production. Eur J Immunol 46:1162–1167 137. Weaver JR, Nadler JL, Taylor-Fishwick DA (2015) Interleukin-12 (IL-12)/STAT4 axis is an important element for beta-cell dysfunction induced by inflammatory cytokines. PLoS ONE 10:e0142735 138. Wang Z, Shen XH, Feng WM, Ye GF, Qiu W, Li B (2016) Analysis of inflammatory mediators in prediabetes and newly diagnosed type 2 diabetes patients. J Diabetes Res 2016:7965317 139. Lee IK, Hsieh CJ, Chen RF, Yang ZS, Wang L, Chen CM, Liu CF, Huang CH, Lin CY, Chen YH, Yang KD, Liu JW (2013) Increased production of interleukin-4, interleukin-10, and granulocyte-macrophage colony-stimulating factor by type 2 diabetes’ mononuclear cells infected with dengue virus, but not increased intracellular viral multiplication. Biomed Res Int 2013:965853 140. Schnupf P, Gaboriau-Routhiau V, Gros M, Friedman R, Moya-Nilges M, Nigro G, Cerf-Bensussan N, Sansonetti PJ (2015) Growth and host interaction of mouse segmented filamentous bacteria in vitro. Nature 520:99–103 141. Arif S, Moore F, Marks K, Bouckenooghe T, Dayan CM, Planas R, Vives-Pi M, Powrie J, Tree T, Marchetti P, Huang GC, Gurzov EN, Pujol-Borrell R, Eizirik DL, Peakman M (2011) Peripheral and islet interleukin-17 pathway activation characterizes human autoimmune diabetes and promotes cytokine-mediated beta-cell death. Diabetes 60:2112–2119

164

R. Li et al.

142. Green EA, Gorelik L, McGregor CM, Tran EH, Flavell RA (2003) CD4+CD25+T regulatory cells control anti-islet CD8+T cells through TGF-beta-TGF-beta receptor interactions in type 1 diabetes. Proc Natl Acad Sci USA 100:10878–10883 143. Joyce SA, Gahan CG (2016) Bile acid modifications at the microbe-host interface: potential for nutraceutical and pharmaceutical interventions in host health. Annu Rev Food Sci Technol 7:313–333 144. Borghede MK, Schlutter JM, Agnholt JS, Christensen LA, Gormsen LC, Dahlerup JF (2011) Bile acid malabsorption investigated by selenium-75-homocholic acid taurine ((75) SeHCAT) scans: causes and treatment responses to cholestyramine in 298 patients with chronic watery diarrhoea. Eur J Intern Med 22:e137–140 145. Mekjian HS, Phillips SF, Hofmann AF (1971) Colonic secretion of water and electrolytes induced by bile acids: perfusion studies in man. J Clin Investig 50:1569–1577 146. Thaysen EH, Pedersen L (1976) Idiopathic bile acid catharsis. Gut 17:965–970 147. Ridlon JM, Kang DJ, Hylemon PB (2006) Bile salt biotransformations by human intestinal bacteria. J Lipid Res 47:241–259 148. Li T, Chiang JY (2014) Bile acid signaling in metabolic disease and drug therapy. Pharmacol Rev 66:948–983 149. Schramm C (2018) Bile acids, the microbiome, immunity, and liver tumors. N Engl J Med 379:888–890 150. Chen J, Thomsen M, Vitetta L (2019) Interaction of gut microbiota with dysregulation of bile acids in the pathogenesis of nonalcoholic fatty liver disease and potential therapeutic implications of probiotics. J Cell Biochem 120:2713–2720 151. Parseus A, Sommer N, Sommer F, Caesar R, Molinaro A, Stahlman M, Greiner TU, Perkins R, Backhed F (2017) Microbiota-induced obesity requires farnesoid X receptor. Gut 66:429–437 152. Song Z, Cai Y, Lao X, Wang X, Lin X, Cui Y, Kalavagunta PK, Liao J, Jin L, Shang J, Li J (2019) Taxonomic profiling and populational patterns of bacterial bile salt hydrolase (BSH) genes based on worldwide human gut microbiome. Microbiome 7:9 153. Joyce SA, MacSharry J, Casey PG, Kinsella M, Murphy EF, Shanahan F, Hill C, Gahan CG (2014) Regulation of host weight gain and lipid metabolism by bacterial bile acid modification in the gut. Proc Natl Acad Sci USA 111:7421–7426 154. Sun L, Xie C, Wang G, Wu Y, Wu Q, Wang X, Liu J, Deng Y, Xia J, Chen B, Zhang S, Yun C, Lian G, Zhang X, Zhang H, Bisson WH, Shi J, Gao X, Ge P, Liu C, Krausz KW, Nichols RG, Cai J, Rimal B, Patterson AD, Wang X, Gonzalez FJ, Jiang C (2018) Gut microbiota and intestinal FXR mediate the clinical benefits of metformin. Nat Med 24:1919– 1929 155. Guo Z, Zhang J, Wang Z, Ang KY, Huang S, Hou Q, Su X, Qiao J, Zheng Y, Wang L, Koh E, Danliang H, Xu J, Lee YK, Zhang H (2016) Intestinal microbiota distinguish gout patients from healthy humans. Sci Rep 6:20602 156. Yu Y, Liu Q, Li H, Wen C, He Z (2018) Alterations of the gut microbiome associated with the treatment of hyperuricaemia in male rats. Front Microbiol 9:2233 157. Garcia-Arroyo FE, Gonzaga G, Munoz-Jimenez I, Blas-Marron MG, Silverio O, Tapia E, Soto V, Ranganathan N, Ranganathan P, Vyas U, Irvin A, Ir D, Robertson CE, Frank DN, Johnson RJ, Sanchez-Lozada LG (2018) Probiotic supplements prevented oxonic acid-induced hyperuricemia and renal damage. PLoS ONE 13:e0202901 158. Yoo JY, Kim SS (2016) Probiotics and prebiotics: present status and future perspectives on metabolic disorders. Nutrients 8:173 159. Yadav H, Lee JH, Lloyd J, Walter P, Rane SG (2013) Beneficial metabolic effects of a probiotic via butyrate-induced GLP-1 hormone secretion. J Biol Chem 288:25088–25097

Chapter 10

Gut Microbiota and Immune Responses Lijun Dong, Jingwen Xie, Youyi Wang, and Daming Zuo

Abstract The gut microbiota consists of a dynamic multispecies community living within a particular niche in a mutual synergy with the host organism. Recent findings have revealed roles for the gut microbiota in the modulation of host immunity and the development and progression of immune-mediated diseases. Besides, growing evidence supports the concept that some metabolites mainly originated from gut microbiota are linked to the immune regulation implicated in systemic inflammatory and autoimmune disorders. In this chapter, we describe the recent advances in our understanding of how host–microbiota interactions shape the immune system, how they affect the pathogenesis of immune-associated diseases and the impact of these mechanisms in the efficacy of disease therapy. Keywords Gut microbiota

10.1

 Immune system  T cells  B cells  Macrophage

Introduction

Human microbiota consists of a remarkable diversity of organisms, involving bacteria, fungi, archaea, protozoans, and viruses, which seem to be more numerous than host cells [1]. Million years of symbiosis between the microorganisms and L. Dong The Fifth Affiliated Hospital, Southern Medical University, Guangzhou 510900, China L. Dong  J. Xie  Y. Wang Department of Immunology, School of Basic Medical Sciences, Southern Medical University, Guangzhou 510515, China Y. Wang  D. Zuo (&) School of Laboratory Medicine and Biotechnology, Institute of Molecular Immunology, Southern Medical University, Guangzhou 510515, China e-mail: [email protected] D. Zuo Microbiome Medicine Center, Zhujiang Hospital, Southern Medical University, Guangzhou 510282, China © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_10

165

166

L. Dong et al.

host has resulted in a mutualistic relationship in which both the microbiota and the host benefit each other. Commensal microbiota acquires niches and nutrients to survive and grow from their host and, in turn, the microbiota contributes to various host physiological and pathophysiological processes. The resident microbes digest plant structural carbohydrates into the metabolites that the host can use for energy and nutrition. Besides, the microbes produce vitamins for disease prevention. The microbiota also plays an essential role in the maintenance of the intestinal barrier, which is crucial for homeostasis and functionality of the gut. It is generally accepted that the gut microbiota colonization is in close relationship with the establishment of intestinal mucosal immunity. The commensal microbiota modulates the maturation of the mucosal immune system, whereas the pathogenic microbiota initiates immunity dysfunction, leading to disease development. Indeed, there is increasing evidence support that gut microbiome can communicate with peripheral immune cells and influence the systemic immune system works. Alternatively, the immune system has adapted to work together with microbiota, and thus control and shape the composition of the microbiota. In this chapter, we provide an overview of the current understanding of the interplay between the gut microbiota and immunity in health and disease.

10.2

Role of Gut Microbiota in Immune System Development and Differentiation

The effect of microbiota on the development and maturation of the host immune system was explored on germ-free (GF) mice which display an underdeveloped immune system associated with dramatically reduction in size and number of secondary immune organs including Peyer’s patches and cryptopatches, altered crypt structure, and limited mucus production from goblet cells [2–5]. Additionally, deficient IgA production, over-activation of anti-inflammatory Th2 cytokines, and decreased expression of pattern recognition receptors (PRRs), like the toll-like receptors (TLRs), are observed in GF animals [6].

10.2.1 Gut Bacteria Numerous studies have revealed the complex connection between gut bacteria and immunity. Initial research had suggested that the crosstalk occurred mainly in the gut lamina propria, where a toxic gut microbial signal maintains a homeostatic immune equilibrium [7]. Recent studies have reported that bacteria–immune interaction also happens at remote extraintestinal sites [8]. Normally, the gut microbial community is predominantly comprised of the two phyla Firmicutes and Bacteroidetes, followed by the phyla Actinobacteria and

10

Gut Microbiota and Immune Responses

167

Verrucomicrobia. A growing body of evidence suggests that gut bacteria are essential for the normal development and function of lymphoid organs and lymphocytes [9–11]. GF animals display extensive deficits in the development of gut-associated lymphoid tissues (GALT) [5]. Notably, GF animals exhibit impaired development and maturation of isolated lymphoid follicles (ILFs) [12]. Specific members of the commensal bacteria, such as B. fragilis and B. subtilis, promote GALT development through a subset of stress responses [13]. Regulatory T cells (Tregs) are most abundant in the colonic mucosa and play a vital role in the maintenance of immune homeostasis. The induction of colonic Tregs requires commensal bacteria with specialized properties in both inflammatory and noninflammatory conditions in the gut [14, 15]. Mazmanian et al. reported that monocolonization of mouse gut with B. fragilis leads to systemic immunological maturity of GF mice and increases the splenic CD4+ fraction to that seen in specific-pathogen-free (SPF) mice [9]. Treatment of mice with B. infantis led to a reduction in intestinal inflammation and an increase in the number of Tregs [16]. Additionally, colonization of GF mice with specific bacterial species is sufficient to instruct CD25+FoxP3+ Tregs differentiation and modulate disease phenotype [10]. Of note, each Bacteroides strain has been shown to induce significant increases in the proportions of Tregs among CD4+ T cells in the colon lamina propria from GF mice [11]. Prominent human symbiont B. fragilis prevents animals from experimental colitis through expanding Tregs, and this beneficial effect depends on a single microbial molecule (polysaccharide A, PSA) [10, 17]. Alternatively, B. fragilis-produced PSA directs the expansion and differentiation of CD4+ T cells, leading to a Th1 response shifting the Th2 immune response of GF mice [9, 18]. Interestingly, oral inoculation of mice with a defined mix of Clostridium strains also boosts colonic Tregs, resulting in resistance to colitis and systemic IgE responses [14]. Integrated microbiome and metabolome analyses reveal that bacterial metabolites mediate communication between the commensal bacteria and the immune system [19, 20]. Butyrate and propionate, two abundant short-chain fatty acids produced by the commensal bacteria, can induce extrathymic Treg induction by their histone deacetylation (HDAC) inhibitory activity and subsequently increased Foxp3 gene transcription [21, 22]. IL-17-expressing CD4+ cells (Th17 cells) constitute a remarkable proportion of lymphocytes presented in the intestinal lamina propria [23, 24]. Colonization of mice with the segmented filamentous bacterium (SFB) strongly induces intestinal Th17 cells in the lamina propria, which play a crucial role in host defense against intestinal pathogens and promote systemic autoimmunity [23, 25–27]. Microbiota low in Firmicutes promotes expression of Th17-related genes and differentiation of Th17 cells in mice, and administration of bacterial isolates from the F. phylum significantly eliminates the heightened Th17 responses in vitro [28]. Mao et al. found that sequential activation of IL22+ innate lymphoid cells (ILCs) followed by Th17 cells and Tregs occurs in the gut as a response to the gut microbiota [29]. Intriguingly, intestinal dysbiosis alters immune homeostasis in the gut, resulting in an enhancement in Tregs and a decrease in IL17-positive cd T cells [30]. Oral administration with

168

L. Dong et al.

L. casei significantly protects mice from colorectal cancer, with the modulation of Tregs to a Th17-biased immune response [31]. The immune balance of Th1/Th2 cells is crucial for immune regulation. Several studies have shown that the gut bacteria and its metabolites affect the balance of Th1/Th2 cells in the intestinal tract [32–34]. B. fragilis could promote the expression of proinflammatory cytokines, thereby enhancing the differentiation of Th1 cells by the activation of the IL-12/STAT4 pathway and major histocompatibility complex II (MHC II) expression [9, 18]. Commensal A4 bacteria, a member of the Lachnospiraceae family, which produce an immunodominant microbiota CBir1 antigen, inhibit lamina propria Th2-cell development [34]. Gut bacteria exert a dramatic, systemic effect on the activation and differentiation of B cells via direct and indirect mechanisms [13, 35, 36]. Children who colonized with E. coli and bifidobacteria early in life exhibits more circulating CD27+ memory B cells in the circulation later than infants not colonized by these bacteria [37]. Early B cell development also occurs in the mouse intestinal lamina propria (LP), where the associated V(D)J recombination/receptor editing processes are driven by interactions with antigens from commensal bacteria [13, 35]. The combination of B. fragilis and B. subtilis resulted in the development of the preimmune Ab repertoire, as indicated by a significant increase in somatic diversification of V(D)J genes [13]. Additionally, the resident bacterial metabolites short-chain fatty acids (SCFAs) modulates the metabolism status in B cells, thereby enhancing the antibody production in the host [36]. The intestinal lamina propria includes antigen-presenting cells with features of dendritic cells (DCs) and macrophages, collectively referred to as mononuclear phagocytes (MNPs), integrating microbial signals to maintain gastrointestinal homeostasis and direct adaptive immunity [38]. It has been reported that intestinal microbial colonization drives the continuous replenishment of macrophages in the intestinal mucosa by monocytes that express C–C chemokine receptor type 2 (CCR2) [39]. CX3CR1+ MNPs, derived from CCR2+Ly6Chi peripheral blood monocytes, are the most abundant intestinal macrophages in the healthy intestine [39]. The local accumulation CX3CR1+ MNPs depends on the enteric flora since a limited amount of CX3CR1+ cells was observed in the LP of GF animals [40]. CX3CR1+ MNPs ensure the maintenance of immune tolerance by facilitating the differentiation of Treg cells [41]. Stimuli from microbial commensals increase the secretion of BMP2 by macrophages, thereby controlling the gastrointestinal motility [42]. Intestinal macrophages also play a crucial role in the maintenance of mucosal T cells, with commensal bacteria-driven production of IL1b by mucosal macrophage having been shown to associate with the development of the Th17 cells [43]. Foligne et al. have determined that DCs internalize L. rhamnosus, thus maintaining an immature phenotype. Moreover, the L. rhamnosus-treated DCs can prime Tregs activity in the gut [44]. Collectively, these data make it clear that the gut bacteria are essential for the education of both the innate and adaptive immune systems (Fig. 10.1).

10

Gut Microbiota and Immune Responses

169

Fig. 10.1 The interaction between gut microbiota and immune system development.Disorders of the proportion and quantity of intestinal commensal bacteria and pathogens can lead to dysbiosis under pathological conditions. Dysbiosis enhances gut permeability and leads to bacterial and toxins translocation. Pattern recognition receptors (PRRs) expressed on activated immune cells (DC) are able to recognize pathogen-associated molecular patterns (PAMPs) expressed on bacterium, thereafter, influencing the CD4+ T cell differentiation into subsets such as Th1, Th2, Th17, and Tregs. Furthermore, the concomitant action of Tregs creates a state of immunosuppression. B cells in the lamina propria of the intestine exert immune function by secreting antibodies after antigen stimulation

10.2.2 Gut Virus Numerous studies have demonstrated the association between the host and viruses in the gut that is more similar to host–bacteria interactions and contains both beneficial and detrimental effects for the host [45–47]. Reovirus enteric infection can result in loss of tolerance to dietary antigens by influencing intestinal immune homeostasis [46, 48]. In an animal model, the differentiation of peripheral regulatory T cells via IFN-I and increased dietary antigen-specific Th1 responses through the transcription factor IFN regulatory factor 1 (IRF1) were blocked by reovirus infection [46]. In addition, patients with celiac disease exhibit elevated production of type I interferon (IFN-I), including IFN-a and IFN-b, by intestinal DCs that promote Th1 responses in the gut that is consistent with the side effect of antiviral therapy [48]. Notably, patients with celiac disease also have higher anti-reovirus antibody titers, correlated with a higher level of IRF1 in the small intestinal mucosa [46]. The gut is a major site of HIV replication and HIV-specific immune response, which leads to aggressive depletion of lamina propria CD4+ T cells during acute infection [49]. Enteric adenoviruses and anelloviruses are markedly increased in HIV-positive patients with low peripheral CD4+ T cell counts, indicating that gut commensal virus was associated with HIV infection [47]. A recent study

170

L. Dong et al.

demonstrated that murine norovirus infects tuft cells, leading to induce type 2 immune responses and promote the virus infection [50].

10.2.3 Gut Fungus The fungal community, also named as mycobiota, inhabits on various barrier surfaces of the host, which range from the skin, lung, oral cavity, and the vagina to the best-known site, the gastrointestinal tract [51]. Evidence supports that fungi can also interact with gut immune cells to maintain intestinal well-being [52, 53]. The dimorphic fungus C. albicans is the majority member of the gut mycobiota and is associated with gut immunological homeostasis [52, 54]. Of note, C. albicans stimulates Th17 cells in the mouse intestine irrelevant of the preexisting gut mycobiome composition [55, 56]. Circulating C. albicans-specific Th17 cells are observed in the peripheral blood of healthy individuals [57]. In addition to recognizing intestinal bacteria, CX3CR1+ MNPs express several antifungal C-type lectin receptors (e.g., dectin-1, dectin-2, and Mincle) and can phagocyte intestinal yeast and filamentous fungi [58]. In the gut, CX3CR1+ MNPs prime antigen-specific Th17 responses to C. albicans via the activation of Syk signaling [52]. Besides, the induction of antifungal antibodies is partially dependent on CX3CR1+ MNPs, which is associated with intestinal inflammation [52]. Interestingly, C. albicans-specific Th17 cells could cross-reactive to other fungal species, like Aspergillus fumigatus [59]. Antifungal antibodies against C. albicans detectable in the serum of patients with colitis co-occur with antibodies recognizing several other mycobiota members belonging to the Saccharomycetaceae family, such as C. parapsilosis, S. cerevisiae, and P. kudriavzevii [52]. These results show that gut mycobiota initiates trained immunity and suggests that mycobiota participates in the intestinal immune homeostasis. However, current studies are limited by the use of few model fungal strains.

10.3

Role of Gut Microbiota in the Regulation of Immune Responses

The immune system is responsible for recognizing, responding, and adapting to countless foreign and self-molecular and is therefore important during conditions of both health and disease. It is well known that the microbiome helps to regulate immune responses, and, in turn, the immune system shapes the composition of the microbiota. GF animals showed limited immune resistance to infection and increased mortality compared with conventionally colonized animals when challenged with pathogenic microorganisms. Management of commensal microbiota affects the outcome of enteric pathogen infection through its modulation of host

10

Gut Microbiota and Immune Responses

171

immune responses to infection [60, 61]. GF animals show reduced antigen-specific systemic immune responses to S. typhimurium, suggesting that enteric pathogens might develop strategies to counter both the immune system and the microbiota during the infectious process [61]. It is conceivable that GF mice have significantly different serum chemokines and cytokine profiles at baseline compared to conventionally raised mice [62]. Additionally, a marked reduction in the levels of secretory IgA in the gut was observed in GF mice [63]. Indeed, gut microbiota dysbiosis is associated with abnormal immune responses, accompanied by aberrant expression of inflammatory cytokines [64]. Gut microbiota contributes to the homeostatic production of pro-IL1b by intestinal phagocytes in a MyD88-dependent manner, thereby priming these cells to respond rapidly to pathogenic bacterial infections by conversion of pro-IL1b to mature IL1b through the NLRC4 inflammasome [65]. Intriguingly, oral administration of F. prausnitzii or supernatant from F. prausnitzii cultures enhanced the production of IL10 by peripheral blood mononuclear cells, reduced the production of TNF in the colon, and ameliorated intestinal disease in mice [66]. Interestingly, treatment of colitis mice with the probiotic cocktail (a mixture of bacteria consisting of four strains of L. casei, L. plantarum, L. acidophilus, and L. delbrueckii) increased the production of IL10 and the percentage of TGFb-expressing T cells [67]. Depletion of TGF-b-bearing CD4+ T cells from mice treated with probiotic bacteria before the transfer of lamina propria cells abolished the protective capacity of these cells [67]. The proinflammatory gut microbiota from caspase1−/− mice accelerates the atherogenesis in the mice model of atherosclerosis and increases the systemic inflammation, accompanied by enhanced proinflammatory cytokines and increased blood leukocyte numbers, implying a relationship between microbiota composition, inflammation, and atherosclerosis [68]. The gut resident microbiota composition profoundly influences the adaptive immune responses to influenza virus infection through the proper activation of inflammasomes, and antibiotic treatment predisposes the mice to high viral replication in the lung [69]. In the gastrointestinal tract, the association of commensal fungi with the innate immune receptor dectin-1 could prevent inflammation in the context of acute mucosal injury [58]. In murine experimental autoimmune encephalomyelitis (EAE) model, modifying certain gut microbiota components affects immune cell priming in the periphery, leading to dysregulation of immune responses and neuroinflammatory processes in the central nervous system (CNS) [70, 71]. Wilck et al. showed that a salt challenge in healthy individuals influences the abundance of intestinal Lactobacilli, accompanied by increased frequencies of proinflammatory Th17 cells in the blood, resulting in aggravation of experimental neuroinflammation [72]. Multiple bacterial species present in the gut produce metabolites, such as short-chain fatty acids (SCFAs), vitamin A, and butyrate, contributing to the orchestration of host immune response [19, 71]. Gut microbiota-produced SCFAs inhibit nuclear factor-jB (NF-jB) signaling and limit the production of TNF in neutrophils and peripheral blood mononuclear cells [73, 74]. Recognition of SCFA by innate immune cells for the regulation of inflammation in response also showed

172

L. Dong et al.

the models of arthritis and allergy [75]. Additionally, administration of the SCFA during EAE ameliorates disease symptoms by increasing Treg cell frequencies in the small intestine, thereby leading to a more anti-inflammatory environment in the gut [71, 76]. Gut bacteria control the production of the vitamin A metabolite retinoic acid to weaken the immune system and increase the risk of infectious diseases [77, 78]. Klebanoff et al. reported that vitamin A controls the homeostasis of pre-DC (precursor of DC)-derived splenic DCs and contributes to the regulatory specialization of mucosal DCs in the gut [78]. Butyrate was shown to markedly promote anti-inflammatory IL10 production in DCs and macrophages, implicating its regulatory role in immune functions [79]. It is noteworthy to mention that IL-10 augments the generation of functional Tregs likely mediated via the STAT3 and Foxo1 pathway [80]. The induction of Tregs in the gastrointestinal tract environment contributes to the maintenance of a mutualistic relationship with the microbiota and the systemic control of host immune responses [16].

10.4

Gut Microbiota and Immune Diseases

The interactions between the host and its gut microbiota are crucial for the maintenance of tissue homeostasis. It is unsurprising that abnormal interactions have emerged as a pivotal driver of various chronic disease states. Indeed, gut dysbiosis and the consequent development of pathogenic populations within the gut microbiota and its metabolites may contribute to a wide variety of pathologies, even in sites distant from the gut, ranging from bowel inflammation to systemic autoimmune disorders and cancer (Table 10.1).

10.4.1 Autoimmune Disorders The etiology of autoimmune disorders is multifactorial, including a mixture of genetic susceptibility and environmental factors. Among the environmental factors that gained attention during the last few decades, the imbalanced gut microbiota has been suggested to involve in the maintenance of healthy human physiology and the development of systemic autoimmune orders [81].

10.4.2 Inflammatory Bowel Diseases Inflammatory bowel disease (IBD), including Crohn’s disease (CD) and ulcerative disease (UC), is a group of chronic inflammatory disorders that affect the gastrointestinal tract and extraintestinal organs. Gut microbes were proven to be an essential factor in intestinal inflammation in IBD [82, 83]. HLA-B27 transgenic

10

Gut Microbiota and Immune Responses

173

Table 10.1 Studies investigating intestinal microbiota in patients Disease subjects Alteration in microbiota Autoimmune disorders IBD mice H. hepaticus IBD mice F. varium IBD patients F. prausnitzii UC mice B. fragilis and its polysaccharide A UC mice Clostridium UC mice B. fragilis T1D mice L. johnsonii N6.2 and some Streptococcal species T1D mice Lactobacillaceae T1D mice Segmented filamentous bacteria T1D mice A. muciniphila SLE mice Clostridia SLE mice Segmented filamentous bacteria SLE mice Lactobacilli SLE mice Synergistetes RA mice RA mice

P. gingivalis Segmented filamentous bacteria Immunodeficiency disorders CVID patient Firmicutes AIDs patients Proteobacteria, Prevotella Cancers CRC patients Fusobacterium CRC mice B. fragilis Colon cancer B. fragilis patients and mice Colorectal cancer Lactobacilli liver tumor mice Melanoma mice pancreatic cancer Epithelial tumors

Clostridium species Bifidobacterium Bifidobacterium A. muciniphila

Immune-related target IL-10 Inflammatory cells Inflammatory cytokines Tregs Tregs IL-10 Th17 cells

References 87 90 68 17 14 17 110,111

CD103+ DCs Th17 cells

112 26

Immune tolerance Th17/Th1 balance IL-17

109 127 129

IL-6 IL-6 and anti-dsDNA titer IL-17 Tfh cells

8 127

Tregs Active T cells, DCs

15 150

Antibody IL-10 CTLA4 Th17-biased immune response NKT cells PD-L1 T cells CCR9+CXCR3+CD4+ T cells

140 141

164 17,165 179 31 170 180 171 181

174

L. Dong et al.

rats, which spontaneously generate chronic colitis, do not develop the inflammatory intestinal disease when they are kept in a GF environment [84]. Uniformly, spontaneous IBD in IL-10 deficient mice associated with the infection with H. hepaticus [85]. The next-generation sequencing analysis showed that Enterobacteriaceae, Pasteurellacaea, Veillonellaceae, and Fusobacteriaceae are revealed to be increased in IBD, while the abundances of Erysipelotrichales, Bacteroidales, and Clostridiales are negatively correlated with disease status [86]. Garrett et al. found that Enterobacteriaceae act in concert with the gut microbiota to cause IBD [87]. F. varium and its metabolite, butyric acid, are able to cause UC-like lesions in mice [88]. Alternatively, there are also studies that determine the specific groups of gut microbiota that may play a protective role against IBD [28, 66, 89]. Compared with healthy controls, the gut microbiota of patients with CD exhibited a markedly reduced diversity of Firmicutes [89], and similar findings were observed in CD patients with monozygotic twins discordant [90]. Indeed, GF mice colonized with microbiota from UC patients low in Firmicutes showed increased sensitivity to colitis compared with mice transferred with synthetic ecosystems rich in Firmicutes [28]. F. prausnitzii displays anti-inflammatory properties and is underrepresented in IBD patients [66]. Indeed, CD patients with low abundances of F. prausnitzii in the mucosa are more prone to have relapse after surgery [91]. It was also confirmed that certain intestinal symbiotic bacteria and their metabolites could promote the immune balance among Th cells in mammals, thereby preventing the development of IBD [17, 92, 93]. Human symbiont B. fragilis and its polysaccharide A (PSA) can inhibit the infiltration of neutrophils in the intestinal tract of experimental animals with IBD by stimulating Treg cells to produce IL-10, thereby preventing mice from developing colitis [17]. Oral inoculation of Clostridium during the early life may alleviate colitis in mice, where the action of the bacteria is facilitated by inducing Tregs [14]. Notably, preliminary data suggest that an elevated load of Candida species in the gut may be associated with the pathogenic feature of CD, indicating a role of the gut mycobiota in the pathogenesis of IBD pathogenesis [94]. Fecal microbiota transplantation (FMT) is a safe and highly effective treatment for colitis caused by C. difficile infection (CDI) [95]. The therapeutic success of FMT in CDI-induced colitis is attributed to the restoration of the gut microbial homeostasis in patients with dysbiosis [96]. It should be noted that CDI-induced colitis is associated with reduced microbial diversity (Bacteroidetes and Firmicutes), which is similar to what is seen in IBD patients [97]. In multiple mouse models, F. prausnitzii, Clostridia strain, and B. fragilis could attenuate the severity of colitis [17, 66, 93]. In the future, FMT will be likely substituted by the use of defined microbe or their metabolites in the treatment of IBD.

10

Gut Microbiota and Immune Responses

175

10.4.3 Type 1 Diabetes Type 1 diabetes (T1D) is caused by autoimmune destruction of insulin-producing beta cells and is believed to involve both genetic and environmental factors. Altered permeability of the gut barrier with microbial dysbiosis has been implicated in the pathogenesis of T1D [98, 99]. A markedly decrease in alpha diversity of the gut microbiome in infants genetically predisposed to T1D during the time window between seroconversion and T1D diagnosis [100]. Comparison of new-onset T1D patients and control subjects suggests that functional changes in the stool metaproteome may be associated with islet autoimmunity [98]. Earlier studies determined that non-obese diabetic (NOD) mice maintained under GF conditions have an increased incidence of diabetes [101, 102]. A recent study showed that GF NOD mice colonized with human gut microbiome are able to delay the onset and reduce the incidence of diabetes [103]. Together, these data suggest that intestinal microbiota is involved in the development of T1D. Of note, NOD mice deficient for MyD88, a canonical adaptor for inflammatory signaling pathways downstream of members of the Toll-like receptor (TLR) and interleukin-1 (IL-1) receptor families, are protected against diabetes, but this effect was diminished in the absence of commensal bacteria [104]. Transfer of gut microbiota from diabetes-protected MyD88−/− NOD mice altered the mucosal immunity and further significantly delayed the development of autoimmune diabetes [105]. Moreover, NOD mice cohoused with dysbiotic inflammasome-deficient mice display lower diabetes incidence compared to noncohoused NOD mice [106], indicating that inflammation modulates the intestinal immune system and inflammation control may be important in T1D management. Many studies have shown that modulation of specific gut microbiota has a significant impact on gut mucosal and the pathogenesis of T1D [26, 107]. Colonization of NOD mice with SFB can induce a substantial population of Th17 cells in the lamina propria and prevent diabetes development [26]. Besides, oral administration of L. johnsonii N6.2 and some Streptococcal species suppresses anti-islet autoimmunity and prevents diabetes development in a spontaneous rat diabetes model by gut flora-mediated Th17 differentiation [108, 109]. Dolpady et al. reported that modification of the gut microbiota profile by oral treatment with Lactobacillaceae-enriched probiotic also affects the inflammasome activation and of tolerogenic CD103(+) DCs differentiation in the gut, therefore changing the intestinal microenvironment and preventing T1D [110]. Notably, the transfer of A. muciniphila to NOD mice promotes mucus production and significantly lowers the incidence of T1D, accompanied by a reduction in serum bacterial endotoxin levels and islet TLR expression [107]. Additionally, emerging evidence indicates that gut microbial metabolites are possible modulators of diabetes [111–114]. Gut microbiota acts as a regulator of glycaemic control in diabetic autoimmunity, and altered microbial metabolism may contribute to the pathogenesis of T1D [111]. Indeed, the beneficial effect of A. muciniphila on protection against the development of T1D is mediated by evoking metabolic and immunologic signaling to

176

L. Dong et al.

facilitate immune tolerance in the gut [107]. Dexi is considered as a candidate gene for human T1D risk [115, 116]. Disruption of Dexi in the NOD mouse depletes gut microbial metabolites and enhances serum levels of IgA and IgM, resulting in accelerated spontaneous autoimmune diabetes [116]. The gut microbiota-derived metabolites, aryl hydrocarbon receptor (AHR) ligands and butyrate, increase the IL-22 production by pancreatic ILCs, thereby preventing autoimmune diabetes in the NOD mice [113]. Strikingly, feeding NOD mice with a combined acetate- and butyrate-yielding diet protects against insulitis, mainly through decreasing the frequency of autoreactive T cells by acetate and boosting Treg cells by butyrate [114]. Collectively, alteration of the gut microbiome or metabolite profile might represent an effective approach for diabetes management.

10.4.4 Systemic Lupus Erythematosus Compared to healthy controls, a depletion of the Lactobacillus population, an enrichment in members of the family Lachnospiraceae, and an increase in overall diversity have been described in the systemic lupus erythematosus (SLE) mouse model, MRL/lpr mice [117]. In an NZB/W F1 murine model, the composition of the gut microbiota was dependent on the disease onset, showing a higher diversity and increased the representation of several bacterial species as lupus progressed from the predisease stage to the diseased stage [118]. Oral antibiotics are known to trigger lupus flares in human, suggesting a crucial role for commensal bacteria in SLE patients [117, 119]. Hevia et al. found a significantly lower Firmicutes/ Bacteroidetes ratio in SLE individuals than in healthy subjects [120]. Consistently, depletion of Firmicutes and enrichment of Bacteroidetes in SLE patients from China were also observed [121]. In contrast, Luo et al. reported that SLE patients with active disease possessed distinct relative abundances of particular Bacterial species, with an increase in Gram-negative bacteria, but the Firmicutes/ Bacteroidetes ratio was no different from that found in a healthy human cohort [118]. The inconsistency conclusion may be due to differences in the study design (active vs. inactive patients; female patients group vs. female and male patients group; patients receiving immune therapies vs. patients not receiving immune therapies). The human gut microbiota provides a major source for antigenic variation, which may trigger autoimmunity via cross-reactivity in genetically susceptible individuals. Greiling et al. reported that human Ro60 autoantigen-specific CD4 memory T cell clones from lupus patients were induced by commensal Ro60-containing bacteria. Moreover, monocolonization with a bacterial Ro60 ortholog enhances systemic lupus-like disease in a lupus mouse model [122]. A key feature of SLE is T cell activation, which facilitates the formation of self-antigen/antibody immune complexes [123]. In vitro studies revealed that microbiota isolated from SLE patient stool samples promoted highly lymphocyte activation and Th17 differentiation compared to healthy control microbiota [124]. Also, fecal microbiome from SLE mice could lead to an obvious increase of

10

Gut Microbiota and Immune Responses

177

anti-dsDNA antibodies and promote the immune response in recipient mice [125]. A mixture of two Clostridia strains significantly reduced the Th17/Th1 balance, whereas B. bifidum supplementation prevented CD4+ lymphocyte over-activation in SLE [124]. SFB strongly induces intestinal Th17 cells in the lamina propria [23, 25], and gut colonization with SFB promotes antinuclear antibodies production in mice mediated by IL-17 receptor signaling [126]. Similarly, restoration of lactobacilli levels by recolonization, antibiotic, or retinoid acid therapy ameliorated lupus symptoms and survival in lupus-prone mice by decreasing IL-17-producing cells and IL-6 secretion in the gut [8]. Intriguingly, reduced amounts of intestinal Synergistetes in SLE microbiota are linked with increased anti-dsDNA titer and IL-6 serum levels but a reduced quantity of natural protective anti-phosphorylcholine IgM antibodies [124]. It is noteworthy that the microbial components may influence levels of host hormones and play a role in disease development. Colonization of female BWF1 mice with gut microbiota from male BWF1 mice suppresses the SLE disease severity by enhancing tolerogenic CD103+ DC activity, which reinforces the role of microbiota in the sexual bias of SLE disease development [127].

10.4.5 Rheumatoid Arthritis Rheumatoid arthritis (RA) is a chronic and systemic autoimmune disease characterized by destructive alterations to the multiple joints. It results from a complex interaction between genetic and environmental factors, leading to a breakdown of immune tolerance and synovial inflammation [128]. The association of gut microbiota in RA development is supported by the observation that mice models, which spontaneously develop arthritis in a conventional environment, but do not develop arthritis under a GF condition [27, 129]. Moreover, a stronger reduction in gut microbial diversity was observed in RA patients compared to healthy controls, which is associated with the disease duration and autoantibodies levels [130]. RA patients under treatment with anti-TNF-a antibody present a partial restoration of beneficial microbiota [131]. Mice susceptible to collagen-induced arthritis (CIA) showed abundant Lactobacillus as the dominant genus prior to arthritis development. L. casei protects against the RA progression by suppressing the Th1-type cellular and humoral immune responses in arthritic inflammation [132, 133]. By contrast, L. rhamnosus and L. reuteri administration failed to attenuate the RA disease in patients suggesting that different Lactobacillus species may act differently on the development of arthritis [134, 135]. Of note, GF-free mice administrated with the commensal microbiota from CIA-susceptible mice showed a higher incidence of arthritis than those conventionalized with the microbiota from CIA-resistant mice. It was shown that the concentration of serum IL-17 and the frequency of CD8+ T cells and Th17 cells in the spleen were significantly higher in the former group, indicating that gut microbiome influences the arthritis susceptibility through regulation of systemic immune response [136]. Consistency,

178

L. Dong et al.

P. gingivalis administration significantly aggravated the arthritis induced by collagen in DBA mice with elevated sera IL-17 levels and a significant alteration in the gut microbiome [137]. Besides, administration of a single gut-residing species, SFB, into GF animals reinstates the lamina propria Th17 cell compartment and triggers autoimmune arthritis, which is associated with the differentiation and migration of Peyer’s patch T follicular helper (Tfh) cells [27, 138]. A recent study demonstrates that the high level of kaempferol in the gut regulates the intestinal flora and microbial metabolism, which are potentially responsible for the anti-arthritis effects of kaempferol [139]. RA patients receiving anti-TNF-a etanercept exhibit an enrichment of Cyanobacteria in the gut [131]. It should be noted that Cyanobacteria produce secondary metabolites with multiple bioactivities, like anti-inflammatory and immunosuppressant activities [140]. It is possible that metabolites produced by Cyanobacteria benefit RA patients. Intriguingly, ES-62, a phosphorylcholine (PC)-containing glycoprotein secreted by the filarial nematode A. viteae, has been able to modulate the gut-associated immune responses [141, 142]. Doonan et al. reported that subcutaneous administration ES-62 protects against joint disease in the CIA, which is related to the normalization of gut microbiota and prevention of loss of intestinal barrier integrity [141].

10.4.6 Immunodeficiency Disorders Immunodeficiency is a state in which the host immune system fails to fight against infection and disease. The cases of immunodeficiency disorders can either be primary by heritage or secondary due to extrinsic factors. Common variable immunodeficiency (CVID) is the most common symptomatic primary immunodeficiency characterized by recurrent infections and low antibody levels. CVID patients have reduced gut bacterial alpha diversity compared to healthy subjects, and the decreased diversity is related to elevated plasma levels of endotoxins and weakened systemic T cell activation [143]. Indeed, the functional impairment of CD4+ T cells in CVID patients is restricted to bacteria-specific CD4+ T cells. Notably, endotoxemia was associated with significantly increased expression of programmed death 1 (PD-1) on CD4+ T cells, indicating that CD4+ T cell exhaustion and functional impairment observed in CVID patients is related to bacterial translocation [144]. CVID patients have a lower proportion of Treg cells compared to healthy controls in their peripheral blood. The hallmark of CVID is hypogammaglobulinemia, which suggests that the IgA production is also impaired. Interestingly, the diversified IgA contributes to maintain the diversity and stability of healthy microbiota, in turn, facilitates the expansion of Tregs [15]. Disruption of this feedback loop in CVID patients through IgA deficiency and microbiota disorder can thereby have a profound effect on Treg cells. Human immunodeficiency virus (HIV) infection is a viral infection that progressively destroys certain immune cells and further lead to acquired immunodeficiency syndrome (AIDS). Gastrointestinal disease is a serious complication of

10

Gut Microbiota and Immune Responses

179

AIDS, indicating that dysbiosis is common among patients with AIDS [145]. Indeed, HIV infection-associated dysbiosis is characterized by reduced levels of a-diversity and Bacteroidetes as well as increased levels of Proteobacteria [146– 148]. Most importantly, the alterations of gut microbiota correlate with the route of HIV transmission [146]. Indeed, the gut microbial dysbiosis is associated with many immunological properties, including CD4+ T cell activation and inflammatory regulation, all of which have been related to the disease progression [147–149]. Interventions to reverse gut dysbiosis could be a new approach for improving immune reconstitution of individuals with HIV infection.

10.4.7 Cancer A growing body of evidence suggests a crucial role for gut microbes in the etiology of cancer, including both local and distant cancers [150]. Previously, the researchers paid much attention to identify specific microbiota that may directly promote or protect against gastrointestinal malignancies. Indeed, the capacity of gut microbes to induce systemic inflammation and immune homeostasis indicates that the commensal microbiota may also be associated with the development of cancers in tissues outside of the gastrointestinal tract. H. pylori is an important bacterium colonized in the human stomach, which provokes the chronic inflammatory state and is believed to account for the occurrence of gastric cancer [151]. Interestingly, several studies show that chronic infection with H. pylori appeared associated with moderately increased risk of colorectal cancer (CRC), particularly among African-Americans [152–154]. A recent study showed that antibodies to four H. pylori proteins were most often present among the different ethnic groups with CRC. VacA, an H. pylori protein, in particular, has a strong association with increased risk of CRC among the African-American patients. Most importantly, high levels of antibodies against VacA linked to the incidence of CRC in both African-Americans and Asian-Americans [154]. It is of interest to determine whether the types and amounts of gut microbiota people harbored based on the genetic origin or heritage, and further studies are warranted. Notably, studies have provided evidence that other species of gut bacteria may play a role in the colonic carcinogenesis [155–159]. E. coli expressing the polyketide synthase (pks+) enhance tumorigenesis in preclinical CRC models [158]. Genomic analysis of the gut microbiome of human CRC patients reveals a major enrichment of Fusobacterium species, such as F. nucleatum, in this cancer [155]. In vitro studies showed that F. nucleatum enhances the CRC cell proliferation [160]. Moreover, F. nucleatum modulates the tumor immune environment by promoting M2 polarization of macrophages and suppressing the function of tumor-infiltrating lymphocytes, thereby affecting the carcinogenesis of CRC [161]. Enterotoxigenic B. fragilis (ETBF) that can secrete B. fragilis toxin is thought to possibly promote early CRC carcinogenesis through affecting the mucosal immune responses [156]. In contrast, Lee et al. revealed that

180

L. Dong et al.

B. fragilis treatment prevents colon tumorigenesis in a mice model of colitisassociated CRC. The decreased tumorigenesis by B. fragilis administration is accompanied by an inhibited expression of C-C chemokine receptor 5 (CCR5) [157]. Indeed, B. fragilis exhibits a preventive effect against intestinal inflammatory diseases in the animal model with colitis [17, 162]. The contradictory function of B. fragilis observed in the development of CRC might due to the imbalanced composition of intestinal microbiota and that intestinal inflammation during the process of CRC. It is worthy to mention that restricting the bloom of Enterobacteriaceae suppresses intestinal inflammation and decreases the incidence of colonic tumors in mouse models of colitis-induced CRC [163]. Apart from the stomach and colorectal cancers, associations of the gut microbiome with the cancers in tissues outside of the gastrointestinal tract (e.g., liver, pancreas, prostate, breast, and others) have also been well documented [150, 164, 165]. The liver has long been known to communicate with the intestine via the portal vein and expose to the substances derived from the gut microbiota [165]. Alteration of gut microbiota modulates the secreted levels of deoxycholic acid (DCA), which can provoke the senescence-associated secretory phenotype in hepatic stellate cells (HSCs) and facilitate hepatocellular carcinoma development in mice [166]. Ma et al. found that colonization with a commensal Clostridium species, which are involved in the conversion of primary to secondary bile acids, inhibits the accumulation of hepatic NKT cells and increases the liver tumor metastasis [167]. Hence, future investigations are warranted to determine the potential effect of these factors on the immunosurveillance of liver cancers. Alteration of gut microbiota is associated with increased risk of pancreatic cancer since removing bacteria from the gut and pancreas reduced the pancreatic cancer growth [168]. Mechanistic studies showed that the abundant gut microbiome drives an immune suppression program in pancreatic cancer [168]. Of note, the capacity of gut microbiome to influence systemic hormone levels may be important in the diseases such as prostate cancer and breast cancer that is dually affected by estrogen and androgen levels [169, 170]. Together, these studies point out the crucial role of gut microbiota in modulating carcinogenesis at distant sites through systemic effects. Accumulating evidence supports that the gut microbiota can influence the effectiveness of cancer immunotherapy [171, 172]. GF mice and mice that are treated with antibiotics both show a diminished response to immunotherapy by CpG oligonucleotides and chemotherapy owing to the impaired function of myeloid-derived cells in the tumor microenvironment [171]. Indeed, some probiotic supplements have shown a potential antineoplastic activity. For example, the culture supernatant from L. casei exhibits a strong tumor-suppressive effect on colon cancer cells, which is mediated by the activation of the JNK pathway [173]. It has also been reported that Lactobacilli is able to induce a Th17-biased immune response for antitumor activity [31]. Treatment with the immune checkpoint inhibitors exhibits dramatic effects on improving survival rates for people with cancer [172]. Preclinical and clinical studies hint a link between the composition of gut microbiome composition and antitumor efficacy of immune checkpoint inhibitors

10

Gut Microbiota and Immune Responses

181

[174]. Unfolded protein response (UPR) is a cellular process that helps to keep protein populations stable by clearing away those that cell stress has caused to fold incorrectly. Li et al. found that UPR activity was lower in people with melanoma whose cancer responds to immune checkpoint inhibitors and pinpointed the UPR as an important link between the gut microbiota and antitumor immunity [175]. Experimental data from animal models demonstrate that gut microbiotas, such as Bacteroides species, Bifidobacteria species, and A. muciniphila, are able to potentiate the anticancer effect of checkpoint blockade therapies, including anti-CTLA-4 and anti-PD-1 antibodies [176–179]. A greater understanding of the communications of commensal bacteria with each other and the host immune system will allow us to improve the patient responses to cancer immune therapy through manipulating the human gut microbiome.

10.5

Concluding Remarks

The importance of the innate immune sensing of commensal microorganisms was recognized merely a decade ago. Since then, multiple levels of interaction between the microbiota and the cells of the immune system have been uncovered. This area of research has led to the integration of the microbiota as an intrinsic regulator of both local and systemic immunities. A comprehensive characterization of the microorganisms and metabolites that are sensed by the host immune system and their effects on the transcriptional and post-transcriptional landscape of the host will greatly expand our understanding about the molecular etiology of microbiotadriven tissue and organ disorders. Importantly, our deepening knowledge about the interactions between the host immune system and the gut microbiota will ultimately result in the development of therapeutic approaches for multiple diseases. However, the various candidates among these complex interactions (e.g., microbiota species, microbial metabolites, host factors, and molecular signaling pathways) as well as the diet influence remain to be determined. In fact, the microbiome is amenable to rapid change through dietary interventions that could be tailored by the host genetic background, microbiome, and metabolome, as well as nutrient intake and habitual food consumption. Identifying the association of gut microbiome alteration with immune response modulation and defining the change of metabolic pathway are necessary for the development of safer strategies for strength host immunity. The future of immunotherapy might combine direct drug-based immune modulation with microbiome and metabolome modification to collectively improve therapeutic outcomes.

182

L. Dong et al.

References 1. Neish AS (2009) Microbes in gastrointestinal health and disease. Gastroenterology 136: 65–80 2. Schaeffer EM (2018) Re: selective depletion of uropathogenic E. coli from the gut by a FimH Antagonist. J Urol 199:874–875 3. Atala A (2018) Re: selective depletion of uropathogenic E. coli from the gut by a FimH Antagonist. J Urol 199:30–31 4. Spaulding CN, Klein RD, Ruer S, Kau AL, Schreiber HL, Cusumano ZT, Dodson KW, Pinkner JS, Fremont DH, Janetka JW, Remaut H, Gordon JI, Hultgren SJ (2017) Selective depletion of uropathogenic E. coli from the gut by a FimH antagonist. Nature 546:528–532 5. Macpherson AJ, Harris NL (2004) Interactions between commensal intestinal bacteria and the immune system. Nat Rev Immunol 4:478–485 6. Round JL, Mazmanian SK (2009) The gut microbiota shapes intestinal immune responses during health and disease. Nat Rev Immunol 9:313–323 7. Gaboriau-Routhiau V, Rakotobe S, Lecuyer E, Mulder I, Lan A, Bridonneau C, Rochet V, Pisi A, De Paepe M, Brandi G, Eberl G, Snel J, Kelly D, Cerf-Bensussan N (2009) The key role of segmented filamentous bacteria in the coordinated maturation of gut helper T cell responses. Immunity 31:677–689 8. Viaud S, Saccheri F, Mignot G, Yamazaki T, Daillere R, Hannani D, Enot DP, Pfirschke C, Engblom C, Pittet MJ, Schlitzer A, Ginhoux F, Apetoh L, Chachaty E, Woerther PL, Eberl G, Berard M, Ecobichon C, Clermont D, Bizet C, Gaboriau-Routhiau V, Cerf-Bensussan N, Opolon P, Yessaad N, Vivier E, Ryffel B, Elson CO, Dore J, Kroemer G, Lepage P, Boneca IG, Ghiringhelli F, Zitvogel L (2013) The intestinal microbiota modulates the anticancer immune effects of cyclophosphamide. Science 342:971–976 9. Mazmanian SK, Liu CH, Tzianabos AO, Kasper DL (2005) An immunomodulatory molecule of symbiotic bacteria directs maturation of the host immune system. Cell 122:107– 118 10. Round JL, Mazmanian SK (2010) Inducible Foxp3+ regulatory T-cell development by a commensal bacterium of the intestinal microbiota. Proc Natl Acad Sci U S A 107:12204– 12209 11. Faith JJ, Ahern PP, Ridaura VK, Cheng J, Gordon JI (2014) Identifying gut microbe-host phenotype relationships using combinatorial communities in gnotobiotic mice. Sci Transl Med 6:220ra11 12. Bouskra D, Brezillon C, Berard M, Werts C, Varona R, Boneca IG, Eberl G (2008) Lymphoid tissue genesis induced by commensals through NOD1 regulates intestinal homeostasis. Nature 456:507–510 13. Rhee KJ, Sethupathi P, Driks A, Lanning DK, Knight KL (2004) Role of commensal bacteria in development of gut-associated lymphoid tissues and preimmune antibody repertoire. J Immunol 172:1118–1124 14. Atarashi K, Tanoue T, Shima T, Imaoka A, Kuwahara T, Momose Y, Cheng G, Yamasaki S, Saito T, Ohba Y, Taniguchi T, Takeda K, Hori S, Ivanov II, Umesaki Y, Itoh K, Honda K (2011) Induction of colonic regulatory T cells by indigenous Clostridium species. Science 331:337–341 15. Kawamoto S, Maruya M, Kato LM, Suda W, Atarashi K, Doi Y, Tsutsui Y, Qin H, Honda K, Okada T, Hattori M, Fagarasan S (2014) Foxp3(+) T cells regulate immunoglobulin a selection and facilitate diversification of bacterial species responsible for immune homeostasis. Immunity 41:152–165 16. O’Mahony C, Scully P, O’Mahony D, Murphy S, O’Brien F, Lyons A, Sherlock G, MacSharry J, Kiely B, Shanahan F, O’Mahony L (2008) Commensal-induced regulatory T cells mediate protection against pathogen-stimulated NF-kappaB activation. PLoS Pathog 4: e1000112

10

Gut Microbiota and Immune Responses

183

17. Mazmanian SK, Round JL, Kasper DL (2008) Anticancer immunotherapy by CTLA-4 blockade relies on the gut microbiota. Nature 453:620–625 18. Wang Q, McLoughlin RM, Cobb BA, Charrel-Dennis M, Zaleski KJ, Golenbock D, Tzianabos AO, Kasper DL (2006) A bacterial carbohydrate links innate and adaptive responses through Toll-like receptor 2. J Exp Med 203:2853–2863 19. Levy M, Thaiss CA, Elinav E (2016) Metabolites: messengers between the microbiota and the immune system. Genes Dev 30:1589–1597 20. Shapiro H, Thaiss CA, Levy M, Elinav E (2014) The cross talk between microbiota and the immune system: metabolites take center stage. Curr Opin Immunol 30:54–62 21. Arpaia N, Campbell C, Fan X, Dikiy S, van der Veeken J, deRoos P, Liu H, Cross JR, Pfeffer K, Coffer PJ, Rudensky AY (2013) Metabolites produced by commensal bacteria promote peripheral regulatory T-cell generation. Nature 504:451–455 22. Furusawa Y, Obata Y, Fukuda S, Endo TA, Nakato G, Takahashi D, Nakanishi Y, Uetake C, Kato K, Kato T, Takahashi M, Fukuda NN, Murakami S, Miyauchi E, Hino S, Atarashi K, Onawa S, Fujimura Y, Lockett T, Clarke JM, Topping DL, Tomita M, Hori S, Ohara O, Morita T, Koseki H, Kikuchi J, Honda K, Hase K, Ohno H (2013) Commensal microbe-derived butyrate induces the differentiation of colonic regulatory T cells. Nature 504:446–450 23. Atarashi K, Tanoue T, Honda K (2010) Induction of lamina propria Th17 cells by intestinal commensal bacteria. Vaccine 28:8036–8038 24. Liu H, Chen F, Wu W, Cao AT, Xue X, Yao S, Evans-Marin HL, Li YQ, Cong Y (2016) TLR5 mediates CD172alpha(+) intestinal lamina propria dendritic cell induction of Th17 cells. Sci Rep 6:22040 25. Ivanov II, Atarashi K, Manel N, Brodie EL, Shima T, Karaoz U, Wei D, Goldfarb KC, Santee CA, Lynch SV, Tanoue T, Imaoka A, Itoh K, Takeda K, Umesaki Y, Honda K, Littman DR (2009) Induction of intestinal Th17 cells by segmented filamentous bacteria. Cell 139:485–498 26. Kriegel MA, Sefik E, Hill JA, Wu HJ, Benoist C, Mathis D (2011) Naturally transmitted segmented filamentous bacteria segregate with diabetes protection in nonobese diabetic mice. P Natl Acad Sci USA 108:11548–11553 27. Wu HJ, Ivanov II, Darce J, Hattori K, Shima T, Umesaki Y, Littman DR, Benoist C, Mathis D (2010) Gut-residing segmented filamentous bacteria drive autoimmune arthritis via T helper 17 cells. Immunity 32:815–827 28. Natividad JM, Pinto-Sanchez MI, Galipeau HJ, Jury J, Jordana M, Reinisch W, Collins SM, Bercik P, Surette MG, Allen-Vercoe E, Verdu EF (2015) Ecobiotherapy rich in firmicutes decreases susceptibility to colitis in a humanized gnotobiotic mouse model. Inflamm Bowel Dis 21:1883–1893 29. Mao K, Baptista AP, Tamoutounour S, Zhuang L, Bouladoux N, Martins AJ, Huang Y, Gerner MY, Belkaid Y, Germain RN (2018) Innate and adaptive lymphocytes sequentially shape the gut microbiota and lipid metabolism. Nature 554:255–259 30. Benakis C, Brea D, Caballero S, Faraco G, Moore J, Murphy M, Sita G, Racchumi G, Ling L, Pamer EG, Iadecola C, Anrather J (2016) Commensal microbiota affects ischemic stroke outcome by regulating intestinal gammadelta T cells. Nat Med 22:516–523 31. Lenoir M, Del Carmen S, Cortes-Perez NG Lozano-Ojalvo D, Munoz-Provencio D, Chain F, Langella P, de Moreno de LeBlanc A, LeBlanc JG, Bermudez-Humaran LG (2016) Lactobacillus casei BL23 regulates Treg and Th17 T-cell populations and reduces DMH-associated colorectal cancer. J Gastroenterol 51:862–873 32. Brucklacher-Waldert V, Carr EJ, Linterman MA, Veldhoen M (2014) Cellular plasticity of CD4+T cells in the intestine. Front Immunol 5:488 33. Qian LJ, Kang SM, Xie JL, Huang L, Wen Q, Fan YY, Lu LJ, Jiang L (2017) Early-life gut microbial colonization shapes Th1/Th2 balance in asthma model in BALB/c mice. BMC Microbiol 17:1–8 34. Wu W, Liu HP, Chen FD, Liu H, Cao AT, Yao SX, Sun MM, Evans-Marin HL, Zhao Y, Zhao Q, Duck LW, Elson CO, Liu ZJ, Cong YZ (2016) Commensal A4 bacteria inhibit

184

35.

36. 37.

38. 39.

40.

41.

42.

43.

44.

45. 46.

47.

48.

49.

L. Dong et al. intestinal Th2-cell responses through induction of dendritic cell TGF-beta production. Eur J Immunol 46:1162–1167 Wesemann DR, Portuguese AJ, Meyers RM, Gallagher MP, Cluff-Jones K, Magee JM, Panchakshari RA, Rodig SJ, Kepler TB, Alt FW (2013) Microbial colonization influences early B-lineage development in the gut lamina propria. Nature 501:112–115 Kim CH (2016) B cell-helping functions of gut microbial metabolites. Microb Cell 3:529– 531 Lundell AC, Bjornsson V, Ljung A, Ceder M, Johansen S, Lindhagen G, Tornhage CJ, Adlerberth I, Wold AE, Rudin A (2012) Infant B cell memory differentiation and early gut bacterial colonization. J Immunol 188:4315–4322 Joeris T, Muller-Luda K, Agace WW, Mowat AM (2017) Diversity and functions of intestinal mononuclear phagocytes. Mucosal Immunol 10:845–864 Bain CC, Bravo-Blas A, Scott CL, Perdiguero EG, Geissmann F, Henri S, Malissen B, Osborne LC, Artis D, Mowat AM (2014) Constant replenishment from circulating monocytes maintains the macrophage pool in the intestine of adult mice. Nat Immunol 15:929–937 Niess JH, Adler G (2010) Enteric flora expands gut lamina propria CX3CR1+ dendritic cells supporting inflammatory immune responses under normal and inflammatory conditions. J Immunol 184:2026–2037 Kim M, Galan C, Hill AA, Wu WJ, Fehlner-Peach H, Song HW, Schady D, Bettini ML, Simpson KW, Longman RS, Littman DR, Diehl GE (2018) Critical role for the microbiota in CX(3)CR1(+) intestinal mononuclear phagocyte regulation of intestinal T cell responses. Immunity 49:151–163 Muller PA, Koscso B, Rajani GM, Stevanovic K, Berres ML, Hashimoto D, Mortha A, Leboeuf M, Li XM, Mucida D, Stanley ER, Dahan S, Margolis KG, Gershon MD, Merad M, Bogunovic M (2014) Crosstalk between muscularis macrophages and enteric neurons regulates gastrointestinal motility. Cell 158:300–313 Shaw MH, Kamada N, Kim YG, Nunez G (2012) Microbiota-induced IL-1 beta, but not IL-6, is critical for the development of steady-state T(H)17 cells in the intestine. J Exp Med 209:251–258 Foligne B, Nutten S, Grangette C, Dennin V, Goudercourt D, Poiret S, Dewulf J, Brassart D, Mercenier A, Pot B (2007) Correlation between in vitro and in vivo immunomodulatory properties of lactic acid bacteria. World J Gastroentero 13:236–243 Neil JA, Cadwell K (2018) The intestinal virome and immunity. J Immunol 201:1615–1624 Bouziat R, Hinterleitner R, Brown JJ, Stencel-Baerenwald JE, Ikizler M, Mayassi T, Meisel M, Kim SM, Discepolo V, Pruijssers AJ, Ernest JD, Iskarpatyoti JA, Costes LMM, Lawrence I, Palanski BA, Varma M, Zurenski MA, Khomandiak S, McAllister N, Aravamudhan P, Boehme KW, Hu F, Samsom JN, Reinecker HC, Kupfer SS, Guandalini S, Semrad CE, Abadie V, Khosla C, Barreiro LB, Xavier RJ, Ng A, Dermody TS, Jabri B (2017) Reovirus infection triggers inflammatory responses to dietary antigens and development of celiac disease. Science 356:44–50 Monaco CL, Gootenberg DB, Zhao G, Handley SA, Ghebremichael MS, Lim ES, Lankowski A, Baldridge MT, Wilen CB, Flagg M, Norman JM, Keller BC, Luevano JM, Wang D, Boum Y, Martin JN, Hunt PW, Bangsberg DR, Siedner MJ, Kwon DS, Virgin HW (2016) Altered virome and bacterial microbiome in human immunodeficiency virus-associated acquired immunodeficiency syndrome. Cell Host Microbe 19:311–322 Di Sabatino A, Pickard KM, Gordon JN, Salvati V, Mazzarella G, Beattie RM, Vossenkaemper A, Rovedatti L, Leakey NAB, Croft NM, Troncone R, Corazza GR, Stagg AJ, Monteleone G, MacDonald TT (2007) Evidence for the role of interferon-alfa production by dendritic cells in the Th1 response in celiac disease. Gastroenterology 133:1175–1187 Brenchley JM, Douek DC (2008) HIV infection and the gastrointestinal immune system. Mucosal Immunol 1:23–30

10

Gut Microbiota and Immune Responses

185

50. Wilen CB, Lee S, Hsieh LL, Orchard RC, Desai C, Hykes BL Jr, McAllaster MR, Balce DR, Feehley T, Brestoff JR, Hickey CA, Yokoyama CC, Wang YT, MacDuff DA, Kreamalmayer D, Howitt MR, Neil JA, Cadwell K, Allen PM, Handley SA, van Lookeren Campagne M, Baldridge MT, Virgin HW (2018) Tropism for tuft cells determines immune promotion of norovirus pathogenesis. Science 360:204–208 51. Limon JJ, Skalski JH, Underhill DM (2017) Commensal fungi in health and disease. Cell Host Microbe 22:156–165 52. Leonardi I, Li X, Semon A, Li D, Doron I, Putzel G, Bar A, Prieto D, Rescigno M, McGovern DPB, Pla J, Iliev ID (2018) CX3CR1(+) mononuclear phagocytes control immunity to intestinal fungi. Science 359:232–236 53. Li XV, Leonardi I, Iliev ID (2019) Gut mycobiota in immunity and inflammatory disease. Immunity 50:1365–1379 54. Bacher P, Hohnstein T, Beerbaum E, Rocker M, Blango MG, Kaufmann S, Rohmel J, Eschenhagen P, Grehn C, Seidel K, Rickerts V, Lozza L, Stervbo U, Nienen M, Babel N, Milleck J, Assenmacher M, Cornely OA, Ziegler M, Wisplinghoff H, Heine G, Worm M, Siegmund B, Maul J, Creutz P, Tabeling C, Ruwwe-Glosenkamp C, Sander LE, Knosalla C, Brunke S, Hube B, Kniemeyer O, Brakhage AA, Schwarz C, Scheffold A (2019) Human anti-fungal Th17 immunity and pathology rely on cross-reactivity against Candida albicans. Cell 176:1340–1355 55. Doron I, Leonardi I, Iliev ID (2019) Profound mycobiome differences between segregated mouse colonies do not influence Th17 responses to a newly introduced gut fungal commensal. Fungal Genet Biol 127:45–49 56. Atarashi K, Tanoue T, Ando M, Kamada N, Nagano Y, Narushima S, Suda W, Imaoka A, Setoyama H, Nagamori T, Ishikawa E, Shima T, Hara T, Kado S, Jinnohara T, Ohno H, Kondo T, Toyooka K, Watanabe E, Yokoyama S, Tokoro S, Mori H, Noguchi Y, Morita H, Ivanov II, Sugiyama T, Nunez G, Camp JG, Hattori M, Umesaki Y, Honda K (2015) Th17 cell induction by adhesion of microbes to intestinal epithelial cells. Cell 163:367–380 57. Becattini S, Latorre D, Mele F, Foglierini M, De Gregorio C, Cassotta A, Fernandez B, Kelderman S, Schumacher TN, Corti D, Lanzavecchia A, Sallusto F (2015) T cell immunity. Functional heterogeneity of human memory CD4(+) T cell clones primed by pathogens or vaccines. Science 347:400–406 58. Iliev ID, Funari VA, Taylor KD, Nguyen Q, Reyes CN, Strom SP, Brown J, Becker CA, Fleshner PR, Dubinsky M, Rotter JI, Wang HL, McGovern DP, Brown GD, Underhill DM (2012) Interactions between commensal fungi and the C-type lectin receptor Dectin-1 influence colitis. Science 336:1314–1317 59. Bacher P, Hohnstein T, Beerbaum E, Rocker M, Blango MG, Kaufmann S, Rohmel J, Eschenhagen P, Grehn C, Seidel K, Rickerts V, Lozza L, Stervbo U, Nienen M, Babel N, Milleck J, Assenmacher M, Cornely OA, Ziegler M, Wisplinghoff H, Heine G, Worm M, Siegmund B, Maul J, Creutz P, Tabeling C, Ruwwe-Glosenkamp C, Sander LE, Knosalla C, Brunke S, Hube B, Kniemeyer O, Brakhage AA, Schwarz C, Scheffold A (2019) Human anti-fungal Th17 immunity and pathology rely on cross-reactivity against Candida albicans. Cell 176:1340–1355 60. Berger AK, Mainou BA (2018) Interactions between enteric bacteria and eukaryotic viruses impact the outcome of infection. Viruses 10:e19 61. Stecher B, Macpherson AJ, Hapfelmeier S, Kremer M, Stallmach T, Hardt WD (2005) Comparison of Salmonella enterica serovar Typhimurium colitis in germfree mice and mice pretreated with streptomycin. Infect Immun 73:3228–3241 62. Wun K, Theriault BR, Pierre JF, Chen EB, Leone VA, Harris KG, Xiong L, Jiang Q, Spedale M, Eskandari OM, Chang EB, Ho KJ (2018) Microbiota control acute arterial inflammation and neointimal hyperplasia development after arterial injury. PLoS ONE 13: e0208426 63. Hapfelmeier S, Lawson MA, Slack E, Kirundi JK, Stoel M, Heikenwalder M, Cahenzli J, Velykoredko Y, Balmer ML, Endt K, Geuking MB, Curtiss R 3rd, McCoy KD,

186

64.

65.

66.

67.

68.

69.

70. 71. 72.

73.

74.

75.

76.

L. Dong et al. Macpherson AJ (2010) Reversible microbial colonization of germ-free mice reveals the dynamics of IgA immune responses. Science 328:1705–1709 Schirmer M, Smeekens SP, Vlamakis H, Jaeger M, Oosting M, Franzosa EA, Ter Horst R, Jansen T, Jacobs L, Bonder MJ, Kurilshikov A, Fu J, Joosten LAB, Zhernakova A, Huttenhower C, Wijmenga C, Netea MG, Xavier RJ (2016) Linking the human gut microbiome to inflammatory cytokine production capacity. Cell 167:1125–1136 Franchi L, Kamada N, Nakamura Y, Burberry A, Kuffa P, Suzuki S, Shaw MH, Kim YG, Nunez G (2012) NLRC4-driven production of IL-1beta discriminates between pathogenic and commensal bacteria and promotes host intestinal defense. Nat Immunol 13:449–456 Sokol H, Pigneur B, Watterlot L, Lakhdari O, Bermudez-Humaran LG, Gratadoux JJ, Blugeon S, Bridonneau C, Furet JP, Corthier G, Grangette C, Vasquez N, Pochart P, Trugnan G, Thomas G, Blottiere HM, Dore J, Marteau P, Seksik P, Langella P (2008) Faecalibacterium prausnitzii is an anti-inflammatory commensal bacterium identified by gut microbiota analysis of Crohn disease patients. Proc Natl Acad Sci U S A 105:16731–16736 Di Giacinto C, Marinaro M, Sanchez M, Strober W, Boirivant M (2005) Probiotics ameliorate recurrent Th1-mediated murine colitis by inducing IL-10 and IL-10-dependent TGF-beta-bearing regulatory cells. J Immunol 174:3237–3246 Brandsma E, Kloosterhuis NJ, Koster M, Dekker DC, Gijbels MJJ, van der Velden S, Rios-Morales M, van Faassen MJR, Loreti MG, de Bruin A, Fu J, Kuipers F, Bakker BM, Westerterp M, de Winther MPJ, Hofker MH, van de Sluis B, Koonen DPY (2019) A proinflammatory gut microbiota increases systemic inflammation and accelerates atherosclerosis. Circ Res 124:94–100 Ichinohe T, Pang IK, Kumamoto Y, Peaper DR, Ho JH, Murray TS, Iwasaki A (2011) Microbiota regulates immune defense against respiratory tract influenza A virus infection. Proc Natl Acad Sci U S A 108:5354–5359 Ma Q, Xing C, Long W, Wang HY, Liu Q, Wang RF (2019) Impact of microbiota on central nervous system and neurological diseases: the gut-brain axis. J Neuroinflammation 16:53 Haase S, Haghikia A, Wilck N, Muller DN, Linker RA (2018) Impacts of microbiome metabolites on immune regulation and autoimmunity. Immunology 154:230–238 Wilck N, Matus MG, Kearney SM, Olesen SW, Forslund K, Bartolomaeus H, Haase S, Mahler A, Balogh A, Marko L, Vvedenskaya O, Kleiner FH, Tsvetkov D, Klug L, Costea PI, Sunagawa S, Maier L, Rakova N, Schatz V, Neubert P, Fratzer C, Krannich A, Gollasch M, Grohme DA, Corte-Real BF, Gerlach RG, Basic M, Typas A, Wu C, Titze JM, Jantsch J, Boschmann M, Dechend R, Kleinewietfeld M, Kempa S, Bork P, Linker RA, Alm EJ, Muller DN (2017) Salt-responsive gut commensal modulates TH17 axis and disease. Nature 551:585–589 Usami M, Kishimoto K, Ohata A, Miyoshi M, Aoyama M, Fueda Y, Kotani J (2008) Butyrate and trichostatin A attenuate nuclear factor kappaB activation and tumor necrosis factor alpha secretion and increase prostaglandin E2 secretion in human peripheral blood mononuclear cells. Nutr Res 28:321–328 Vinolo MA, Rodrigues HG, Hatanaka E, Sato FT, Sampaio SC, Curi R (2011) Suppressive effect of short-chain fatty acids on production of proinflammatory mediators by neutrophils. J Nutr Biochem 22:849–855 Maslowski KM, Vieira AT, Ng A, Kranich J, Sierro F, Yu D, Schilter HC, Rolph MS, Mackay F, Artis D, Xavier RJ, Teixeira MM, Mackay CR (2009) Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature 461:1282–1286 Haghikia A, Jorg S, Duscha A, Berg J, Manzel A, Waschbisch A, Hammer A, Lee DH, May C, Wilck N, Balogh A, Ostermann AI, Schebb NH, Akkad DA, Grohme DA, Kleinewietfeld M, Kempa S, Thone J, Demir S, Muller DN, Gold R, Linker RA (2016) Dietary fatty acids directly impact central nervous system autoimmunity via the small intestine. Immunity 44:951–953

10

Gut Microbiota and Immune Responses

187

77. Grizotte-Lake M, Zhong G, Duncan K, Kirkwood J, Iyer N, Smolenski I, Isoherranen N, Vaishnava S (2018) Commensals suppress intestinal epithelial cell retinoic acid synthesis to regulate interleukin-22 activity and prevent microbial dysbiosis. Immunity 49:1103–1115 78. Klebanoff CA, Spencer SP, Torabi-Parizi P, Grainger JR, Roychoudhuri R, Ji Y, Sukumar M, Muranski P, Scott CD, Hall JA, Ferreyra GA, Leonardi AJ, Borman ZA, Wang J, Palmer DC, Wilhelm C, Cai R, Sun J, Napoli JL, Danner RL, Gattinoni L, Belkaid Y, Restifo NP (2013) Retinoic acid controls the homeostasis of pre-cDC-derived splenic and intestinal dendritic cells. J Exp Med 210:1961–1976 79. Kaisar MMM, Pelgrom LR, van der Ham AJ, Yazdanbakhsh M, Everts B (2017) Butyrate conditions human dendritic cells to prime type 1 regulatory T cells via both histone deacetylase inhibition and G protein-coupled receptor 109A signaling. Front Immunol 8:1429 80. Hsu P, Santner-Nanan B, Hu M, Skarratt K, Lee CH, Stormon M, Wong M, Fuller SJ, Nanan R (2015) IL-10 potentiates differentiation of human induced regulatory T cells via STAT3 and Foxo1. J Immunol 195:3665–3674 81. Dehner C, Fine R, Kriegel MA (2019) The microbiome in systemic autoimmune disease: mechanistic insights from recent studies. Curr Opin Rheumatol 31:201–207 82. Sartor RB (2008) Microbial influences in inflammatory bowel diseases. Gastroenterology 134:577–594 83. Tamboli CP, Desreumaux P et al (2004) Dysbiosis as a prerequisite for IBD. Gut 53:1057 84. Taurog JD, Richardson JA, Croft JT, Simmons WA, Zhou M, Fernandez-Sueiro JL, Balish E, Hammer RE (1994) The germfree state prevents development of gut and joint inflammatory disease in HLA-B27 transgenic rats. J Exp Med 180:2359–2364 85. Yang I, Eibach D, Kops F, Brenneke B, Woltemate S, Schulze J, Bleich A, Gruber AD, Muthupalani S, Fox JG, Josenhans C, Suerbaum S (2013) Intestinal microbiota composition of interleukin-10 deficient C57BL/6 J mice and susceptibility to Helicobacter hepaticus-induced colitis. PLoS ONE 8:e70783 86. Gevers D, Kugathasan S, Denson LA, Vazquez-Baeza Y, Van Treuren W, Ren B, Schwager E, Knights D, Song SJ, Yassour M, Morgan XC, Kostic AD, Luo C, Gonzalez A, McDonald D, Haberman Y, Walters T, Baker S, Rosh J, Stephens M, Heyman M, Markowitz J, Baldassano R, Griffiths A, Sylvester F, Mack D, Kim S, Crandall W, Hyams J, Huttenhower C, Knight R, Xavier RJ (2014) The treatment-naive microbiome in new-onset Crohn’s disease. Cell Host Microbe 15:382–392 87. Garrett WS, Gallini CA, Yatsunenko T, Michaud M, DuBois A, Delaney ML, Punit S, Karlsson M, Bry L, Glickman JN, Gordon JI, Onderdonk AB, Glimcher LH (2010) Enterobacteriaceae act in concert with the gut microbiota to induce spontaneous and maternally transmitted colitis. Cell Host Microbe 8:292–300 88. Ohkusa T, Okayasu I, Ogihara T, Morita K, Ogawa M, Sato N (2003) Induction of experimental ulcerative colitis by Fusobacterium varium isolated from colonic mucosa of patients with ulcerative colitis. Gut 52:79–83 89. Manichanh C, Rigottier-Gois L, Bonnaud E, Gloux K, Pelletier E, Frangeul L, Nalin R, Jarrin C, Chardon P, Marteau P, Roca J, Dore J (2006) Reduced diversity of faecal microbiota in Crohn’s disease revealed by a metagenomic approach. Gut 55:205–211 90. Dicksved J, Halfvarson J, Rosenquist M, Jarnerot G, Tysk C, Apajalahti J, Engstrand L, Jansson JK (2008) Molecular analysis of the gut microbiota of identical twins with Crohn’s disease. ISME J 2:716–727 91. Sokol H, Seksik P, Furet JP, Firmesse O, Nion-Larmurier I, Beaugerie L, Cosnes J, Corthier G, Marteau P, Dore J (2009) Low counts of Faecalibacterium prausnitzii in colitis microbiota. Inflamm Bowel Dis 15:1183–1189 92. Ishidoh K, Kominami E (2002) Processing and activation of lysosomal proteinases. Biol Chem 383:1827–1831 93. Atarashi K, Tanoue T, Oshima K, Suda W, Nagano Y, Nishikawa H, Fukuda S, Saito T, Narushima S, Hase K, Kim S, Fritz JV, Wilmes P, Ueha S, Matsushima K, Ohno H, Olle B, Sakaguchi S, Taniguchi T, Morita H, Hattori M, Honda K (2013) Treg induction by a

188

94.

95. 96. 97.

98.

99.

100.

101. 102.

103.

104.

105. 106.

107.

108.

109.

110.

L. Dong et al. rationally selected mixture of Clostridia strains from the human microbiota. Nature 500:232– 236 Mukhopadhya I, Hansen R, Meharg C, Thomson JM, Russell RK, Berry SH, El-Omar EM, Hold GL (2015) The fungal microbiota of de-novo paediatric inflammatory bowel disease. Microbes Infect 17:304–310 D’Aoust J, Battat R, Bessissow T (2017) Management of inflammatory bowel disease with Clostridium difficile infection. World J Gastroenterol 23:4986–5003 Borody TJ, Paramsothy S, Agrawal G (2013) Fecal microbiota transplantation: indications, methods, evidence, and future directions. Curr Gastroenterol Rep 15:337 Zhang L, Dong D, Jiang C, Li Z, Wang X, Peng Y (2015) Insight into alteration of gut microbiota in Clostridium difficile infection and asymptomatic C. difficile colonization. Anaerobe 34:1–7 Gavin PG, Mullaney JA, Loo D, Cao KAL, Gottlieb PA, Hill MM, Zipris D, Hamilton-Williams EE (2018) Intestinal metaproteomics reveals host-microbiota interactions in subjects at risk for type 1 diabetes. Diabetes Care 41:2178–2186 Pellegrini S, Sordi V, Bolla AM, Saita D, Ferrarese R, Canducci F, Clementi M, Invernizzi F, Mariani A, Bonfanti R, Barera G, Testoni PA, Doglioni C, Bosi E, Piemonti L (2017) Duodenal mucosa of patients with type 1 diabetes shows distinctive inflammatory profile and microbiota. J Clin Endocr Metab 102:1468–1477 Kostic AD, Gevers D, Siljander H, Vatanen T, Hyotylainen T, Hamalainen AM, Peet A, Tillmann V, Poho P, Mattila I, Lahdesmaki H, Franzosa EA, Vaarala O, de Goffau M, Harmsen H, Ilonen J, Virtanen SM, Clish CB, Oresic M, Huttenhower C, Knip M, Group DS, Xavier RJ (2015) The dynamics of the human infant gut microbiome in development and in progression toward type 1 diabetes. Cell Host Microbe 17:260–273 King C, Sarvetnick N (2011) The Incidence of type-1 diabetes in NOD mice is modulated by restricted flora not germ-free conditions. PLoS ONE 6:e17049 Alam C, Bittoun E, Bhagwat D, Valkonen S, Saari A, Jaakkola U, Eerola E, Huovinen P, Hanninen A (2011) Effects of a germ-free environment on gut immune regulation and diabetes progression in non-obese diabetic (NOD) mice. Diabetologia 54:1398–1406 Neuman V, Cinek O, Funda DP, Hudcovic T, Golias J, Kramna L, Petruzelkova L, Pruhova S, Sumnik Z (2019) Human gut microbiota transferred to germ-free NOD mice modulate the progression towards type 1 diabetes regardless of the pace of beta cell function loss in the donor. Diabetologia 62:1291–1296 Wen L, Ley RE, Volchkov PY, Stranges PB, Avanesyan L, Stonebraker AC, Hu C, Wong FS, Szot GL, Bluestone JA, Gordon JI, Chervonsky AV (2008) Innate immunity and intestinal microbiota in the development of type 1 diabetes. Nature 455:1109–1113 Peng J, Narasimhan S, Marchesi JR, Benson A, Wong FS, Wen L (2014) Long term effect of gut microbiota transfer on diabetes development. J Autoimmun 53:85–94 Yu H, Gagliani N, Ishigame H, Huber S, Zhu S, Esplugues E, Herold KC, Wen L, Flavell RA (2017) Intestinal type 1 regulatory T cells migrate to periphery to suppress diabetogenic T cells and prevent diabetes development. P Natl Acad Sci USA 114:10443– 10448 Hanninen A, Toivonen R, Poysti S, Belzer C, Plovier H, Ouwerkerk JP, Emani R, Cani PD, De Vos WM (2018) Akkermansia muciniphila induces gut microbiota remodelling and controls islet autoimmunity in NOD mice. Gut 67:1445–1453 Lau K, Benitez P, Ardissone A, Wilson TD, Collins EL, Lorca G, Li N, Sankar D, Wasserfall C, Neu J, Atkinson MA, Shatz D, Triplett EW, Larkin J (2011) Inhibition of type 1 diabetes correlated to a lactobacillus johnsonii N6.2-mediated Th17 bias. J Immunol 186:3538–3546 Satoh J, Shintani S, Oya K, Tanaka S, Nobunaga T, Toyota T, Goto Y (1988) Treatment with streptococcal preparation (OK-432) suppresses anti-islet autoimmunity and prevents diabetes in BB rats. Diabetes 37:1188–1194 Dolpady J, Sorini C, Di Pietro C, Cosorich I, Ferrarese R, Saita D, Clementi M, Canducci F, Falcone M (2016) Oral probiotic VSL#3 prevents autoimmune diabetes by modulating

10

111.

112.

113.

114.

115.

116.

117. 118.

119.

120.

121. 122.

123. 124.

125. 126.

Gut Microbiota and Immune Responses

189

microbiota and promoting indoleamine 2,3-dioxygenase-enriched tolerogenic intestinal environment. J Diabetes Res 2016:7569431 Greiner TU, Hyotylainen T, Knip M, Backhed F, Oresic M (2014) The gut microbiota modulates glycaemic control and serum metabolite profiles in non-obese diabetic mice. PLoS ONE 9:e110359 Marino E, Richards JL, McLeod KH, Yap YA, Stanley D, Mackay CR (2016) Gut microbial metabolites regulate autoimmune T cell responses and protect against type 1 diabetes. Eur J Immunol 46:538–538 Miani M, Le Naour J, Waeckel-Enee E, Verma SC, Straube M, Emond P, Ryffel B, Van Endert P, Sokol H, Diana J (2018) Gut microbiota-stimulated innate lymphoid cells support beta-defensin 14 expression in pancreatic endocrine cells. Prev Autoimmune Diabetes Cell Metab 28:557–572 Marino E, Richards JL, McLeod KH, Stanley D, Yap YA, Knight J, McKenzie C, Kranich J, Oliveira AC, Rossello FJ, Krishnamurthy B, Nefzger CM, Macia L, Thorburn A, Baxter AG, Morahan G, Wong LH, Polo JM, Moore RJ, Lockett TJ, Clarke JM, Topping DL, Harrison LC, Mackay CR (2017) Gut microbial metabolites limit the frequency of autoimmune T cells and protect against type 1 diabetes. Nat Immunol 18:552–562 Dos Santos RS, Marroqui L, Velayos T, Olazagoitia-Garmendia A, Jauregi-Miguel A, Castellanos-Rubio A, Eizirik DL, Castano L, Santin I (2019) DEXI, a candidate gene for type 1 diabetes, modulates rat and human pancreatic beta cell inflammation via regulation of the type I IFN/STAT signalling pathway. Diabetologia 62:459–472 Davison LJ, Wallace MD, Preece C, Hughes K, Todd JA, Davies B, Callaghan CO (2018) Dexi disruption depletes gut microbial metabolites and accelerates autoimmune diabetes. BioRxiv Zhang H, Liao X, Sparks JB, Luo XM, Schloss PD (2014) Dynamics of gut microbiota in autoimmune lupus. Appl Environ Microbiol 80:7551–7560 Luo XM, Edwards MR, Mu Q, Yu Y, Vieson MD, Reilly CM, Ahmed SA, Bankole AA (2018) Gut microbiota in human systemic lupus erythematosus and a mouse model of lupus. Appl Environ Microbiol 84:e02288 Margolis DJ, Hoffstad O, Bilker W (2007) Association or lack of association between tetracycline class antibiotics used for acne vulgaris and lupus erythematosus. Br J Dermatol 157:540–546 Hevia A, Milani C, Lopez P, Cuervo A, Arboleya S, Duranti S, Turroni F, Gonzalez S, Suarez A, Gueimonde M, Ventura M, Sanchez B, Margolles A (2014) Intestinal dysbiosis associated with systemic lupus erythematosus. MBio 5:e01548–e01514 He Z, Shao T, Li H, Xie Z, Wen C (2016) Alterations of the gut microbiome in Chinese patients with systemic lupus erythematosus. Gut Pathog 8:64 Greiling TM, Dehner C, Chen XG, Hughes K, Iniguez AJ, Boccitto M, Ruiz DZ, Renfroe SC, Vieira SM, Ruff WE, Sim S, Kriegel C, Glanternik J, Chen XD, Girardi M, Degnan P, Costenbader KH, Goodman AL, Wolin SL, Kriegel MA (2018) Commensal orthologs of the human autoantigen Ro60 as triggers of autoimmunity in lupus. Sci Transl Med 10:eaan2306 Comte D, Karampetsou MP, Tsokos GC (2015) T cells as a therapeutic target in SLE. Lupus 24:351–363 Lopez P, de Paz B, Rodriguez-Carrio J, Hevia A, Sanchez B, Margolles A, Suarez A (2016) Th17 responses and natural IgM antibodies are related to gut microbiota composition in systemic lupus erythematosus patients. Sci Rep 6:24072 Ma Y, Xu X, Li M, Cai J, Wei Q, Niu H (2019) Gut microbiota promote the inflammatory response in the pathogenesis of systemic lupus erythematosus. Mol Med 25 Van Praet JT, Donovan E, Vanassche I, Drennan MB, Windels F, Dendooven A, Allais L, Cuvelier CA, van de Loo F, Norris PS, Kruglov AA, Nedospasov SA, Rabot S, Tito R, Raes J, Gaboriau-Routhiau V, Cerf-Bensussan N, Van de Wiele T, Eberl G, Ware CF, Elewaut D (2015) Commensal microbiota influence systemic autoimmune responses. EMBO J 34:466–474

190

L. Dong et al.

127. Chhabra A, Alard P, Jala V, Haribabu B, Kosiewicz M (2014) A role for hormones, gut microbiota and tolerogenic CD103DC in protection of male (NZBxNZW)F1 (BWF1) mice from lupus. J Immunol 192:198 128. Croia C, Bursi R, Sutera D, Petrelli F, Alunno A, Puxeddu I (2019) One year in review 2019: pathogenesis of rheumatoid arthritis. Clin Exp Rheumatol 37:347–357 129. Maeda Y, Kurakawa T, Umemoto E, Motooka D, Ito Y, Gotoh K, Hirota K, Matsushita M, Furuta Y, Narazaki M, Sakaguchi N, Kayama H, Nakamura S, Iida T, Saeki Y, Kumanogoh A, Sakaguchi S, Takeda K (2016) Dysbiosis contributes to arthritis development via activation of autoreactive T cells in the intestine. Arthritis Rheumatol 68:2646–2661 130. Chen J, Wright K, Davis JM, Jeraldo P, Marietta EV, Murray J, Nelson H, Matteson EL, Taneja V (2016) An expansion of rare lineage intestinal microbes characterizes rheumatoid arthritis. Genome Med 8:43 131. Picchianti-Diamanti A, Panebianco C, Salemi S, Sorgi ML, Di Rosa R, Tropea A, Sgrulletti M, Salerno G, Terracciano F, D’Amelio R, Lagana B, Pazienza V (2018) Analysis of gut microbiota in rheumatoid arthritis patients: disease-related Dysbiosis and modifications induced by Etanercept. Int J Mol Sci 19:e2938 132. So JS, Kwon HK, Lee CG, Yi HJ, Park JA, Lim SY, Hwang KC, Jeon YH, Im SH (2008) Lactobacillus casei suppresses experimental arthritis by down-regulating T helper 1 effector functions. Mol Immunol 45:2690–2699 133. So JS, Lee CG, Kwon HK, Yi HJ, Chae CS, Park JA, Hwang KC, Im SH (2008) Lactobacillus casei potentiates induction of oral tolerance in experimental arthritis. Mol Immunol 46:172–180 134. Pineda Mde L, Thompson SF, Summers K, de Leon F, Pope J, Reid G (2011) A randomized, double-blinded, placebo-controlled pilot study of probiotics in active rheumatoid arthritis. Med Sci Monit 17:CR347–CR354 135. Hatakka K, Martio J, Korpela M, Herranen M, Poussa T, Laasanen T, Saxelin M, Vapaatalo H, Moilanen E, Korpela R (2003) Effects of probiotic therapy on the activity and activation of mild rheumatoid arthritis–a pilot study. Scand J Rheumatol 32:211–215 136. Liu X, Zeng B, Zhang J, Li W, Mou F, Wang H, Zou Q, Zhong B, Wu L, Wei H, Fang Y (2016) Role of the gut microbiome in modulating arthritis progression in mice. Sci Rep 6:30594 137. Sato K, Takahashi N, Kato T, Matsuda Y, Yokoji M, Yamada M, Nakajima T, Kondo N, Endo N, Yamamoto R, Noiri Y, Ohno H, Yamazaki K (2017) Aggravation of collageninduced arthritis by orally administered porphyromonas gingivalis through modulation of the gut microbiota and gut immune system. Sci Rep 7:6955 138. Teng F, Klinger CN, Felix KM, Bradley CP, Wu E, Tran NL, Umesaki Y, Wu HJJ (2016) Gut microbiota drive autoimmune arthritis by promoting differentiation and Migration of Peyer’s patch T follicular helper cells. Immunity 44:875–888 139. Aa L-X, Fei F, Qi Q, Sun R-B, Gu S-H, Di Z-Z, Aa J-Y, Wang G-J, Liu C-X (2020) Rebalancing of the gut flora and microbial metabolism is responsible for the anti-arthritis effect of kaempferol. Acta Pharmacol Sin 41:73–81 140. Raja R, Hemaiswarya S, Ganesan V, Carvalho IS (2016) Recent developments in therapeutic applications of Cyanobacteria. Crit Rev Microbiol 42:394–405 141. Doonan J, Tarafdar A, Pineda MA, Lumb FE, Crowe J, Khan AM, Hoskisson PA, Harnett MM, Harnett W (2019) The parasitic worm product ES-62 normalises the gut microbiota bone marrow axis in inflammatory arthritis. Nat Commun 10:1554 142. Eason RJ, Bell KS, Marshall FA, Rodgers DT, Pineda MA, Steiger CN, Al-Riyami L, Harnett W, Harnett MM (2016) The helminth product, ES-62 modulates dendritic cell responses by inducing the selective autophagolysosomal degradation of TLR-transducers, as exemplified by PKCdelta. Sci Rep 6:37276 143. Jorgensen SF, Troseid M, Kummen M, Anmarkrud JA, Michelsen AE, Osnes LT, Holm K, Hoivik ML, Rashidi A, Dahl CP, Vesterhus M, Halvorsen B, Mollnes TE, Berge RK, Moum B, Lundin KE, Fevang B, Ueland T, Karlsen TH, Aukrust P, Hov JR (2016) Altered

10

144.

145.

146.

147.

148.

149.

150. 151. 152. 153.

154.

155.

156.

157.

158.

Gut Microbiota and Immune Responses

191

gut microbiota profile in common variable immunodeficiency associates with levels of lipopolysaccharide and markers of systemic immune activation. Mucosal Immunol 9:1455– 1465 Perreau M, Vigano S, Bellanger F, Pellaton C, Buss G, Comte D, Roger T, Lacabaratz C, Bart PA, Levy Y, Pantaleo G (2014) Exhaustion of bacteria-specific CD4 T cells and microbial translocation in common variable immunodeficiency disorders. J Exp Med 211:2033–2045 Bhaijee F, Subramony C, Tang SJ, Pepper DJ (2011) Human immunodeficiency virus-associated gastrointestinal disease: common endoscopic biopsy diagnoses. Patholog Res Int 2011:247923 Zhou Y, Ou Z, Tang X, Zhou Y, Xu H, Wang X, Li K, He J, Du Y, Wang H, Chen Y, Nie Y (2018) Alterations in the gut microbiota of patients with acquired immune deficiency syndrome. J Cell Mol Med 22:2263–2271 Dillon SM, Lee EJ, Kotter CV, Austin GL, Dong Z, Hecht DK, Gianella S, Siewe B, Smith DM, Landay AL, Robertson CE, Frank DN, Wilson CC (2014) An altered intestinal mucosal microbiome in HIV-1 infection is associated with mucosal and systemic immune activation and endotoxemia. Mucosal Immunol 7:983–994 Lu W, Feng Y, Jing F, Han Y, Lyu N, Liu F, Li J, Song X, Xie J, Qiu Z, Zhu T, Routy B, Routy JP, Li T, Zhu B (2018) Association between gut microbiota and CD4 recovery in HIV-1 infected patients. Front Microbiol 9:1451 Nowak P, Troseid M, Avershina E, Barqasho B, Neogi U, Holm K, Hov JR, Noyan K, Vesterbacka J, Svard J, Rudi K, Sonnerborg A (2015) Gut microbiota diversity predicts immune status in HIV-1 infection. AIDS 29:2409–2418 Hullar MA, Burnett-Hartman AN, Lampe JW (2014) Gut microbes, diet, and cancer. Cancer Treat Res 159:377–399 Wroblewski LE, Peek RM Jr, Wilson KT (2010) Helicobacter pylori and gastric cancer: factors that modulate disease risk. Clin Microbiol Rev 23:713–739 Papastergiou V, Karatapanis S, Georgopoulos SD (2016) Helicobacter pylori and colorectal neoplasia: is there a causal link? World J Gastroenterol 22:649–658 Park H, Park JJ, Park YM, Baik SJ, Lee HJ, Jung DH, Kim JH, Youn YH, Park H (2018) The association between Helicobacter pylori infection and the risk of advanced colorectal neoplasia may differ according to age and cigarette smoking. Helicobacter 23:e12477 Butt J, Varga MG, Blot WJ, Teras L, Visvanathan K, Le Marchand L, Haiman C, Chen Y, Bao Y, Sesso HD, Wassertheil-Smoller S, Ho GYF, Tinker LE, Peek RM, Potter JD, Cover TL, Hendrix LH, Huang LC, Hyslop T, Um C, Grodstein F, Song M, Zeleniuch-Jacquotte A, Berndt S, Hildesheim A, Waterboer T, Pawlita M, Epplein M (2019) Serologic response to helicobacter pylori proteins associated with risk of colorectal cancer among diverse populations in the United States. Gastroenterology 156:175–186 Kostic AD, Gevers D, Pedamallu CS, Michaud M, Duke F, Earl AM, Ojesina AI, Jung J, Bass AJ, Tabernero J, Baselga J, Liu C, Shivdasani RA, Ogino S, Birren BW, Huttenhower C, Garrett WS, Meyerson M (2012) Genomic analysis identifies association of Fusobacterium with colorectal carcinoma. Genome Res 22:292–298 Purcell RV, Pearson J, Aitchison A, Dixon L, Frizelle FA, Keenan JI (2017) Colonization with enterotoxigenic Bacteroides fragilis is associated with early-stage colorectal neoplasia. PLoS ONE 12:e0171602 Lee YK, Mehrabian P, Boyajian S, Wu WL, Selicha J, Vonderfecht S, Mazmanian SK (2018) The protective role of bacteroides fragilis in a murine model of colitis-associated colorectal cancer. mSphere 3:e00587 Arthur JC, Perez-Chanona E, Muhlbauer M, Tomkovich S, Uronis JM, Fan TJ, Campbell BJ, Abujamel T, Dogan B, Rogers AB, Rhodes JM, Stintzi A, Simpson KW, Hansen JJ, Keku TO, Fodor AA, Jobin C (2012) Intestinal inflammation targets cancerinducing activity of the microbiota. Science 338:120–123

192

L. Dong et al.

159. Sobhani I, Amiot A, Le Baleur Y, Levy M, Auriault ML, Van Nhieu JT, Delchier JC (2013) Microbial dysbiosis and colon carcinogenesis: could colon cancer be considered a bacteria-related disease? Therap Adv Gastroenterol 6:215–229 160. Bullman S, Pedamallu CS, Sicinska E, Clancy TE, Zhang X, Cai D, Neuberg D, Huang K, Guevara F, Nelson T, Chipashvili O, Hagan T, Walker M, Ramachandran A, Diosdado B, Serna G, Mulet N, Landolfi S, Ramon YCS, Fasani R, Aguirre AJ, Ng K, Elez E, Ogino S, Tabernero J, Fuchs CS, Hahn WC, Nuciforo P, Meyerson M (2017) Analysis of Fusobacterium persistence and antibiotic response in colorectal cancer. Science 358:1443– 1448 161. Wu J, Li Q, Fu X (2019) Fusobacterium nucleatum contributes to the carcinogenesis of colorectal cancer by inducing inflammation and suppressing host immunity. Transl Oncol 12:846–851 162. Chang YC, Ching YH, Chiu CC, Liu JY, Hung SW, Huang WC, Huang YT, Chuang HL (2017) TLR2 and interleukin-10 are involved in Bacteroides fragilis-mediated prevention of DSS-induced colitis in gnotobiotic mice. PLoS ONE 12:e0180025 163. Zhu W, Miyata N, Winter MG, Arenales A, Hughes ER, Spiga L, Kim J, Sifuentes-Dominguez L, Starokadomskyy P, Gopal P, Byndloss MX, Santos RL, Burstein E, Winter SE (2019) Editing of the gut microbiota reduces carcinogenesis in mouse models of colitis-associated colorectal cancer. J Exp Med 216:2378–2393 164. Wong SH, Kwong TNY, Wu CY, Yu J (2019) Clinical applications of gut microbiota in cancer biology. Semin Cancer Biol 55:28–36 165. Ohtani N, Kawada N (2019) Role of the gut-liver axis in liver inflammation, fibrosis, and cancer: a special focus on the gut microbiota relationship. Hepatol Commun 3:456–470 166. Yoshimoto S, Loo TM, Atarashi K, Kanda H, Sato S, Oyadomari S, Iwakura Y, Oshima K, Morita H, Hattori M, Honda K, Ishikawa Y, Hara E, Ohtani N (2013) Obesity-induced gut microbial metabolite promotes liver cancer through senescence secretome. Nature 499: 97–101 167. Ma C, Han M, Heinrich B, Fu Q, Zhang Q, Sandhu M, Agdashian D, Terabe M, Berzofsky JA, Fako V, Ritz T, Longerich T, Theriot CM, McCulloch JA, Roy S, Yuan W, Thovarai V, Sen SK, Ruchirawat M, Korangy F, Wang XW, Trinchieri G, Greten TF (2018) Gut microbiome-mediated bile acid metabolism regulates liver cancer via NKT cells. Science 360:eaan5931 168. Pushalkar S, Hundeyin M, Daley D, Zambirinis CP, Kurz E, Mishra A, Mohan N, Aykut B, Usyk M, Torres LE, Werba G, Zhang K, Guo K, Li Q, Akkad Q, Lall S, Wadowski B, Gutierrez J, Kochen Rossi JA, Herzog JW, Diskin B, Torres-Hernandez A, Leinwand J, Wang W, Taunk PS, Savadkar S, Janal M, Saxena A, Li X, Cohen D, Sartor RB, Saxena D, Miller G (2018)The pancreatic cancer microbiome promotes oncogenesis by induction of innate and adaptive immune suppression. Cancer Discov 8:403–416 169. Porter CM, Shrestha E, Peiffer LB, Sfanos KS (2018) The microbiome in prostate inflammation and prostate cancer. Prostate Cancer Prostatic Dis 21:345–354 170. Fernandez MF, Reina-Perez I, Astorga JM, Rodriguez-Carrillo A, Plaza-Diaz J, Fontana L (2018) Breast cancer and its relationship with the microbiota. Int J Environ Res Public Health 15:e1747 171. Iida N, Dzutsev A, Stewart CA, Smith L, Bouladoux N, Weingarten RA, Molina DA, Salcedo R, Back T, Cramer S, Dai RM, Kiu H, Cardone M, Naik S, Patri AK, Wang E, Marincola FM, Frank KM, Belkaid Y, Trinchieri G, Goldszmid RS (2013) Commensal bacteria control cancer response to therapy by modulating the tumor microenvironment. Science 342:967–970 172. Frankel AE, Deshmukh S, Reddy A, Lightcap J, Hayes M, McClellan S, Singh S, Rabideau B, Glover TG, Roberts B, Koh AY (2019) Cancer immune checkpoint inhibitor therapy and the gut microbiota. Integr Cancer Ther 18:1534735419846379 173. Konishi H, Fujiya M, Tanaka H, Ueno N, Moriichi K, Sasajima J, Ikuta K, Akutsu H, Tanabe H, Kohgo Y (2016) Probiotic-derived ferrichrome inhibits colon cancer progression via JNK-mediated apoptosis. Nat Commun 7:12365

10

Gut Microbiota and Immune Responses

193

174. Gong J, Chehrazi-Raffle A, Placencio-Hickok V, Guan M, Hendifar A, Salgia R (2019) The gut microbiome and response to immune checkpoint inhibitors: preclinical and clinical strategies. Clin Transl Med 8:9 175. Li Y, Tinoco R, Elmen L, Segota I, Xian Y, Fujita Y, Sahu A, Zarecki R, Marie K, Feng Y, Khateb A, Frederick DT, Ashkenazi SK, Kim H, Perez EG, Day CP, Segura Munoz RS, Schmaltz R, Yooseph S, Tam MA, Zhang T, Avitan-Hersh E, Tzur L, Roizman S, Boyango I, Bar-Sela G, Orian A, Kaufman RJ, Bosenberg M, Goding CR, Baaten B, Levesque MP, Dummer R, Brown K, Merlino G, Ruppin E, Flaherty K, Ramer-Tait A, Long T, Peterson SN, Bradley LM, Ronai ZA (2019) Gut microbiota dependent anti-tumor immunity restricts melanoma growth in Rnf5(-/-) mice. Nat Commun 10:1492 176. Vetizou M, Pitt JM, Daillere R, Lepage P, Waldschmitt N, Flament C, Rusakiewicz S, Routy B, Roberti MP, Duong CP, Poirier-Colame V, Roux A, Becharef S, Formenti S, Golden E, Cording S, Eberl G, Schlitzer A, Ginhoux F, Mani S, Yamazaki T, Jacquelot N, Enot DP, Berard M, Nigou J, Opolon P, Eggermont A, Woerther PL, Chachaty E, Chaput N, Robert C, Mateus C, Kroemer G, Raoult D, Boneca IG, Carbonnel F, Chamaillard M, Zitvogel L (2015) Anticancer immunotherapy by CTLA-4 blockade relies on the gut microbiota. Science 350:1079–1084 177. Sivan A, Corrales L, Hubert N, Williams JB, Aquino-Michaels K, Earley ZM, Benyamin FW, Lei YM, Jabri B, Alegre ML, Chang EB, Gajewski TF (2015) Commensal Bifidobacterium promotes antitumor immunity and facilitates anti-PD-L1 efficacy. Science 350:1084–1089 178. Routy B, Le Chatelier E, Derosa L, Duong CPM, Alou MT, Daillere R, Fluckiger A, Messaoudene M, Rauber C, Roberti MP, Fidelle M, Flament C, Poirier-Colame V, Opolon P, Klein C, Iribarren K, Mondragon L, Jacquelot N, Qu B, Ferrere G, Clemenson C, Mezquita L, Masip JR, Naltet C, Brosseau S, Kaderbhai C, Richard C, Rizvi H, Levenez F, Galleron N, Quinquis B, Pons N, Ryffel B, Minard-Colin V, Gonin P, Soria JC, Deutsch E, Loriot Y, Ghiringhelli F, Zalcman G, Goldwasser F, Escudier B, Hellmann MD, Eggermont A, Raoult D, Albiges L, Kroemer G, Zitvogel L (2018) Gut microbiome influences efficacy of PD-1-based immunotherapy against epithelial tumors. Science 359: 91–97 179. Matson V, Fessler J, Bao R, Chongsuwat T, Zha Y, Alegre ML, Luke JJ, Gajewski TF (2018) The commensal microbiome is associated with anti-PD-1 efficacy in metastatic melanoma patients. Science 359:104–108

Chapter 11

Gut Microbiota and Multiple Organ Dysfunction Syndrome (MODS) Peng Chen and Timothy Billiar

Abstract Multiple organ dysfunction syndrome (MODS), also referred to as external challenge-induced multiple organ injury, is characterized by dysfunction of two or more organs during infection or following shock or trauma. The pathogenesis of MODS is multifactorial and involves systemic inflammation and cell stress responses including cell death; sepsis is defined as an infection with MODS. Gut microbiota contributes significantly to organ dysfunction and to the pathobiology of sepsis. However, the relationship between the development of sepsis and the composition of gut microbiota is equivocal and is only now starting to be elucidated. Recent studies by our group and others reveal that enteric microbial composition and function are disrupted during sepsis, and that microbial products can either promote or alleviate the progression of sepsis. Here, we summarize the current research on the functional link between gut microbiota and sepsis, and argue the point that gut microbiota is a potential therapeutic target in the management of sepsis. Keywords Gut microbiota

11.1

 MODS  Sepsis

Sepsis

Sepsis is one of the leading public health problems worldwide and is caused by a dysregulated host immune response during infection, leading to multiple organ dysfunction syndrome (MODS) [1, 2]. Overall, the patient mortality from sepsis is P. Chen Department of Pathophysiology, Guangdong Provincial Key Laboratory of Proteomics, School of Basic Medical Sciences, Southern Medical University, N.No.1838 Guangzhou Ave., Guangzhou 510515, China e-mail: [email protected] T. Billiar (&) Department of Surgery, University of Pittsburgh, Pittsburgh, PA, USA e-mail: [email protected] © Springer Nature Singapore Pte Ltd. 2020 P. Chen (ed.), Gut Microbiota and Pathogenesis of Organ Injury, Advances in Experimental Medicine and Biology 1238, https://doi.org/10.1007/978-981-15-2385-4_11

195

196

P. Chen and T. Billiar

approximately 20–30% and it is the principle cause of death in intensive care units (ICU) [3–5]. In 2016, the definition of sepsis was updated and now incorporates infection, dysregulated immune response, and organ dysfunction as contributing factors. Specifically, the combination of an infection and a sequential organ failure assessment (SOFA) score of  2 is recognized as sepsis [6]. A SOFA score is calculated from the ratio of arterial oxygen partial pressure to fractional inspired oxygen (PaO2/FiO2), and using plasma bilirubin and creatinine concentrations and blood pressure readings, among other parameters. The detail information about SOFA score could refer to reference 6. When sepsis is complicated by a low mean arterial pressure (2 mmol/L), the condition is defined as septic shock. The cardiovascular, neurologic, and hematologic systems, with liver, kidney, and lung, encompass the respective systems and organs disrupted in sepsis. During the initial stages of sepsis, the lungs are usually first affected, resulting in decreased respiratory function, as characterized by reduced PaO2/FiO2. Furthermore, sepsis involves the liver, with its dysfunction evidenced by increase in plasma bilirubin and aminotransferases. In addition, coagulation disturbances are often observed and are associated with diminished platelet numbers. This effect is partially due to microthrombi formation leading to malperfusion at the level of microcirculation. Vascular dysfunction and capillary leaks lead to drops in blood pressure that are resistant to vasopressor treatments. Renal dysfunction, as measured by elevated creatinine, is one of the major causes of adverse outcomes and often requires organ replacement therapy. Finally, impairment of central nervous system function is estimated by the Glasgow Coma Score. Sepsis induced MODS is identified by dysfunction of two or more organs, complications and high mortality [1]. Thus, the pathogenesis of sepsis is complex and involves dysregulated host immune responses, and damage to normal organs, tissues, and cells. For example, uncontrolled infection precedes systemic inflammation from cytokine over-production by macrophages and other immune cells, that is, “cytokine storm” [7, 8]. Depressed immune defenses render the host susceptible to secondary microbial infections and further propagate organ dysfunction [9]. Overall, the dysregulated immune responses in sepsis are complicated and are the topics of ongoing investigation.

11.2

Gut Microbiota and Sepsis

Knowledge about the link between gut microbiota and sepsis is limited. Traditionally, it was considered that sepsis augmented intestinal permeability and promoted translocation of gut microbiota to the circulatory system and peripheral organs, amplifying inflammation and organ injury [10–12]. However, the specific mechanisms by which sepsis enhances gut permeability remain unclear, and few data are available to link intestinal microbial composition and/or functional

11

Gut Microbiota and Multiple Organ Dysfunction Syndrome (MODS)

197

alterations to sepsis. Investigating this intriguing hypothesis, we and others are the first to establish a direct link between sepsis and enteric microbial eubiosis [13]. In the clinic, we found that sepsis patients show enteric dysbiosis compared with healthy subjects. Thus, the microbial composition between the two groups was different; alpha-diversity metrics (Chao1 index, Observed species, Shannon index, and Simpson index) were decreased in patients with sepsis as compared with healthy subjects. Moreover, specific strains such as Bacteroidetes, Proteobacteria, and Actinobacteria were enriched, whereas Firmicutes was reduced in patients with sepsis. However, most patients with sepsis were treated with antibiotics, so it is reasonable to question whether the microbial disruptions resulted from sepsis or were due to antibiotics. As we could not withdraw antibiotics from patients, we confirmed our results in animal models. We established murine sepsis models by cecum ligation punch (CLP) and by Streptococcus pneumoniae administration; both models caused obvious alterations in gut microbial composition. These data indicate that sepsis may directly lead to enteric dysbiosis. Microbial functions were also disrupted during sepsis development. By employing metabolomic and proteomic analysis techniques, we reveal that metabolites and proteins generated from intestinal microbiota were different between healthy individuals and patients with sepsis. For example, dimethyl trisulfide and ceramide (d18:0/25:0) were elevated, and cytosine and 6-tridecene were decreased in septic feces. Using bioinformatics to classify our proteomic data, gene ontology enrichment analysis indicated that microbial proteins belong to the biological processes pathway, and that cellular components and molecular functions were likely changed in patients with sepsis. Hence, these results directly signify that microbial functions were influenced during sepsis development. Disturbed eubiosis could directly promote organ damage during sepsis. We performed fecal microbiota transplantation (FMT) experiments and found that mice given feces from patients with sepsis versus that from healthy subjects had more severe liver injury and inflammation. Thus, the altered gut microbiota observed in sepsis was not only the consequence, as there was increased sepsis-related organ injury. Overall, our clinical and experimental pilot data have systematically uncovered the association between sepsis and enteric eubiosis at both the compositional and functional levels for microbiota. However, further mechanistic studies are required to extend this novel investigative direction in intensive care medicine. Besides our work, other groups also found similar microbial phenotypes. In ICU patients, it is reported that gut microbial composition at the phyla level showed dynamic changes after their admission to ICU. Specifically, Bacteroidetes to Firmicutes (B/F ratio) in deceased patients tended to increase, whereas the B/F ratio was stable in survivors. Although this was a small-scale study in only 12 patients, the data suggest a new direction by which gut microbiota may serve as a predictor of ICU mortality [14]. Large-scale clinical trials are urgently required to confirm this hypothesis. Zaborin et al. showed that alpha-diversity (Chao1 index) was decreased in ICU patients compared with healthy subjects [15]. More importantly, they found that the intestinal microbial community in prolonged critical illness was dramatically

198

P. Chen and T. Billiar

changed, as characterized by ultra-low-diversity bacterial communities. They further demonstrated that the “commensal lifestyle” of bacteria could be shifted to a “pathogenic style”, strengthening the hypothesis that gut microbiota plays a key role during sepsis progression. Besides clinical observations, basic research from animal models also unmasks essential roles for gut microbiota in intensive care diseases. It is reported that germ-free mice were more sensitive to Pseudomonas aeruginosa-induced pneumonia, and this phenotype may have resulted from altered gut epithelial cell apoptosis [16]. Similarly, Schuijt et al. disclosed that antibiotic-treated mice were more sensitive to pneumococcal pneumonia, and furthermore, the protective effects of the gut microbiota were possibly mediated by modulation of alveolar macrophage function [17]. Although these studies clearly suggest that the gut microbiota is important for sepsis and/or pneumonia, the actual mechanisms are largely unknown. Using a CLP-induced sepsis model to investigate the novel mechanism of gut microbial action in sepsis progression [18, 19], we found that the metabolic profile was different between sepsis-sensitive mice (mice who died either before or around 24 h after CLP) and sepsis-resistant mice (mice who survived until 7 days after CLP). Our findings were noteworthy, as fecal and plasma granisetron levels were significantly greater in sepsis-resistant mice than in sepsis-sensitive mice. Moreover, we also found that granisetron-mitigated CLP-induced lung and liver damage via reduction of cytokine/chemokine over-expression in macrophages. This molecular mechanism may involve the Toll-like receptor 4 (TLR4)-signaling pathway, because the protective effect of granisetron was blocked in TLR4 knockout mice. Additionally, granisetron may affect inflammasome activation in macrophages. Thus, the anti-inflammatory effects of granisetron may be complex. To translate this basic study to the clinic, we are now performing a clinical trial to determine whether granisetron can be used as a novel treatment strategy for patients with sepsis. Although the fundamental mechanisms by which the gut microbiota influences sepsis remain unclear, we can conclude that several key inflammatory pathways are involved. First, gut microbial strains can exert pathogenic changes during sepsis development [20]; for example, Escherichia coli, the classic gram-negative bacterium, augments host infection and inflammation through lipopolysaccharide toxicity. Similarly, other commensal bacteria may also become pathogenic, likely due to disrupted host immune systems. Secondly, gut commensal microbiota could modulate gut barrier integrity. Interactions between microbiota and intestinal epithelial cells (IECs) are complex. Bacteria-derived products such as short-chain fatty acids may act directly on IECs and strengthen the gut barrier. Notably, microbiota may also modulate goblet cells function, which influences mucus production [1]. However, inflammatory responses due to changes in gut microbiota are recognized as a main regulatory pathway during interactions between the gut microbiota and gut barrier. Inflammatory

11

Gut Microbiota and Multiple Organ Dysfunction Syndrome (MODS)

199

factors, such as immune cell activation, and cytokine/chemokine generation, disrupt gut barrier integrity. Augmented gut leakiness could amplify systemic inflammation via the release of microbiota and/or microbial products from the intestine into the circulatory system and peripheral organs; this mechanism is a key pathophysiologic process contributing to organ deterioration during sepsis. Thirdly, intestinal immunologic responses modulated by microbiota may impact the gut barrier, and influence immune reactions and inflammation in the periphery. Gut microbiota may modulate adaptive immune cells like Treg, Th17, B cells, and innate immune cells, such as macrophages and neutrophils. In fact, microbial-generated lipids may activate intestinal natural killer T cells, which could migrate to the liver and cause hepatocyte apoptosis and subsequent liver damage [21]. There are few direct clues about whether intestinal-derived immune cells can migrate into the circulatory system and promote sepsis progression; therefore, additional investigations should focus on this topic. Targeting gut microbiota is increasingly recognized as a novel and effective approach for sepsis management, and this approach translates from basic research to clinical practice. Rodent studies revealed that a high-fiber diet could increase survival during CLP-induced sepsis, and this sepsis-resistant phenotype was accompanied by reduced systemic pro-inflammatory cytokine production and decreased organ inflammation [22]. The authors further demonstrated that a high-fiber diet could modulate gut microbial composition, as characterized by enrichment of Lachnospiraceae and Akkermansia, which is known as a new “probiotic” and could provide beneficial effects to the host. Promising clinical data by Shimizu et al. found that synbiotics (Bifidobacterium breve strain Yakult, Lactobacillus casei strain Shirota, and galactooligosaccharides) significantly reduced enteritis and ventilator-associated pneumonia, without affecting mortality in patients with sepsis [23]. One study confirmed that probiotics could provide protection to ICU patients against intestinal inflammation [24], and a meta-analysis revealed that probiotics and synbiotics may be beneficial for postoperative sepsis [25]. Besides specific strain therapies, whole microbiome interventions, such as FMT, were also considered as a potential treatment approach. For example, FMT was an efficient method for curing ICU patients with Clostridium difficile infection [26]. Moreover, some case reports suggest that FMT could treat severe diarrhea in ICU patients [27], but definitive data are urgently required to provide guidance to clinicians and intensive care units for successful FMT administration. Importantly, the FMT approach has both advantages and disadvantages. As a full spectrum treatment, it may improve the whole microbial environment in the intestine and may be more effective than a single strain for maintaining a healthy gut. However, current reports about FMT use in the ICU mainly focus on sepsis in patients with intestinal abnormalities. Little information is available about whether FMT may improve extra-intestinal organ injuries, although basic research has demonstrated that the gut microbiome could affect multiple organ damage. Future clinical studies are needed to expand FMT

200

P. Chen and T. Billiar

application to the pathophysiology of other organs. Moreover, donor selection is complicated, and there are no common criteria for donor selection in this situation. Thus, the outcome of FMT depends largely on the interaction between donor feces and the recipient gut. Because of the complicated pathogenesis of sepsis, FMT therapy may not be appropriate with all patients with sepsis, and more detailed clinical practice knowledge is required [28].

11.3

Conclusion

Although increasing attention has been paid to the link between gut microbiota and sepsis pathogenesis (Fig. 11.1), mechanistic insights are still limited. The types of microbiota and microbial products which impact the host to influence sepsis development and progression, as well as the molecular pathways which mediate it, are still largely unknown. Nonetheless, focusing on the gut microbiota as a therapeutic target is a promising strategy for the future treatment of sepsis. Basic research scientists and clinicians should combine their efforts to fill this knowledge gap in intensive care medicine.

Organs injury

Gut microbiota Modulate organs injury

Generate protecve products, e.g. SCFAs, granisetron Disrupt gut barrier Exert “pathogenic”effects ……

Disrupt enteric eubiosis Alpha-diversity↓Metabolites, Bacteroidetes↑ proteins Acnobacteria↑ disrupon ……

Sepsis development Fig. 11.1 The bidirectional crosstalk between gut microbiota and organ injury during sepsis

11

Gut Microbiota and Multiple Organ Dysfunction Syndrome (MODS)

201

References 1. Haak BW, Wiersinga WJ (2017) The role of the gut microbiota in sepsis. Lancet Gastroenterol Hepatol 2(2):135–143 2. Livingston DH, Mosenthal AC, Deitch EA (1995) Sepsis and multiple organ dysfunction syndrome: a clinical-mechanistic overview. New Horiz 3(2):257–266 3. Vincent JL, Marshall JC, Namendys-Silva SA, François B, Martin-Loeches I, Lipman J, Reinhart K, Antonelli M, Pickkers P, Njimi H, Jimenez E, Sakr Y (2014) ICON investigators. Assessment of the worldwide burden of critical illness: the intensive care over nations (ICON) audit. Lancet Respir Med 2(5):380–386 4. Alberti C, Brun-Buisson C, Burchardi H, Martin C, Goodman S, Artigas A, Sicignano A, Palazzo M, Moreno R, Boulmé R, Lepage E, Le Gall R (2002) Epidemiology of sepsis and infection in ICU patients from an international multicentre cohort study. Intensive Care Med 28(2):108–121 5. Pittet D, Rangel-Frausto S, Li N, Tarara D, Costigan M, Rempe L, Jebson P, Wenzel RP (1995) Systemic inflammatory response syndrome, sepsis, severe sepsis and septic shock: incidence, morbidities and outcomes in surgical ICU patients. Intensive Care Med 21(4):302– 309 6. Singer M, Deutschman CS, Seymour CW, Shankar-Hari M, Annane D, Bauer M, Bellomo R, Bernard GR, Chiche JD, Coopersmith CM, Hotchkiss RS, Levy MM, Marshall JC, Martin GS, Opal SM, Rubenfeld GD, van der Poll T, Vincent JL, Angus DC (2016) The third international consensus definitions for sepsis and septic shock (Sepsis-3). JAMA 315(8):801– 810 7. Chousterman BG, Swirski FK, Weber GF (2017) Cytokine storm and sepsis disease pathogenesis. Semin Immunopathol 39(5):517–528 8. London NR, Zhu W, Bozza FA, Smith MC, Greif DM, Sorensen LK, Chen L, Kaminoh Y, Chan AC, Passi SF, Day CW, Barnard DL, Zimmerman GA, Krasnow MA, Li DY (2010) Targeting Robo4-dependent slit signaling to survive the cytokine storm in sepsis and influenza. Sci Transl Med 2(23):23ra19 9. Angele MK, Wichmann MW, Ayala A, Cioffi WG, Chaudry IH (1997) Testosterone receptor blockade after hemorrhage in males. Restoration of the depressed immune functions and improved survival following subsequent sepsis. Arch Surg 132(11):1207–1214 10. Kamada N, Seo SU, Chen GY, Núñez G (2013) Role of the gut microbiota in immunity and inflammatory disease. Nat Rev Immunol 13(5):321–335 11. Tripathi A, Debelius J, Brenner DA, Karin M, Loomba R, Schnabl B, Knight R (2018) The gut-liver axis and the intersection with the microbiome. Nat Rev Gastroenterol Hepatol 15 (7):397–411 12. Chen P, Stärkel P, Turner JR, Ho SB, Schnabl B (2015) Dysbiosis-induced intestinal inflammation activates tumor necrosis factor receptor I and mediates alcoholic liver disease in mice. Hepatology 61(3):883–894 13. Liu Z, Li N, Fang H, Chen X, Guo Y, Gong S, Niu M, Zhou H, Jiang Y, Chang P, Chen P (2019) Enteric dysbiosis is associated with sepsis in patients. FASEB J 33(11):12299-12310 14. Ojima M, Motooka D, Shimizu K, Gotoh K, Shintani A, Yoshiya K, Nakamura S, Ogura H, Iida T, Shimazu T (2016) Metagenomic analysis reveals dynamic changes of whole gut microbiota in the acute phase of intensive care unit patients. Dig Dis Sci 61(6):1628–1634 15. Zaborin A, Smith D, Garfield K, Quensen J, Shakhsheer B, Kade M, Tirrell M, Tiedje J, Gilbert JA, Zaborina O, Alverdy JC (2014) Membership and behavior of ultra-low-diversity pathogen communities present in the gut of humans during prolonged critical illness. mBio 5 (5):e01361-14 16. Fox AC, McConnell KW, Yoseph BP, Breed E, Liang Z, Clark AT, O’Donnell D, Zee-Cheng B, Jung E, Dominguez JA, Dunne WM, Burd EM, Coopersmith CM (2012) The endogenous bacteria alter gut epithelial apoptosis and decrease mortality following Pseudomonas aeruginosa pneumonia. Shock 38(5):508–514

202

P. Chen and T. Billiar

17. Schuijt TJ, Lankelma JM, Scicluna BP, de Sousa e Melo F, Roelofs JJ, de Boer JD, Hoogendijk AJ, de Beer R, de Vos A, Belzer C, de Vos WM, van der Poll T, Wiersinga WJ (2016) The gut microbiota plays a protective role in the host defence against pneumococcal pneumonia. Gut 65(4):575–583 18. Wang J, Gong S, Wang F, Niu M, Wei G, He Z, Gu T, Jiang Y, Liu A, Chen P (2019) Granisetron protects polymicrobial sepsis-induced acute lung injury in mice. Biochem Biophys Res Commun 508(4):1004–1010 19. Gong S, Yan Z, Liu Z, Niu M, Fang H, Li N, Huang C, Li L, Chen G, Luo H, Chen X, Zhou H, Hu J, Yang W, Huang Q, Schnabl B, Chang P, Billiar TR, Jiang Y, Chen P (2019) Intestinal microbiota mediates the susceptibility to polymicrobial sepsis-induced liver injury by granisetron generation in mice. Hepatology 69(4):1751–1767 20. Taylor FB Jr, Stearns-Kurosawa DJ, Kurosawa S, Ferrell G, Chang AC, Laszik Z, Kosanke S, Peer G, Esmon CT (2000) The endothelial cell protein C receptor aids in host defense against Escherichia coli sepsis. Blood 95(5):1680–1686 21. Lee KC, Chen P, Maricic I, Inamine T, Hu J, Gong S, Sun JC, Dasgupta S, Lin HC, Lin YT, Loomba R, Stärkel P, Kumar V, Schnabl B (2019) Intestinal iNKT cells migrate to liver and contribute to hepatocyte apoptosis during alcoholic liver disease. Am J Physiol Gastrointest Liver Physiol 316(5):G585–G597 22. Morowitz MJ, Di Caro V, Pang D, Cummings J, Firek B, Rogers MB, Ranganathan S, Clark RSB, Aneja RK (2017) Dietary supplementation with nonfermentable fiber alters the gut microbiota and confers protection in murine models of sepsis. Crit Care Med 45(5):e516– e523 23. Shimizu K, Yamada T, Ogura H, Mohri T, Kiguchi T, Fujimi S, Asahara T, Yamada T, Ojima M, Ikeda M, Shimazu T (2018) Synbiotics modulate gut microbiota and reduce enteritis and ventilator-associated pneumonia in patients with sepsis: a randomized controlled trial. Crit Care 22(1):239 24. AlFaleh K, Anabrees J (2014) Probiotics for prevention of necrotizing enterocolitis in preterm infants. Cochrane Database Syst Rev 4:CD005496 25. Arumugam S, Lau CS, Chamberlain RS (2016) Probiotics and synbiotics decrease postoperative sepsis in elective gastrointestinal surgical patients: a meta-analysis. J Gastrointest Surg 20(6):1123–1131 26. Han S, Shannahan S, Pellish R (2016) Fecal microbiota transplant: treatment options for clostridium difficile infection in the intensive care unit. J Intensive Care Med. 31(9):577–586 27. Wei Y, Yang J, Wang J, Yang Y, Huang J, Gong H, Cui H, Chen D (2016) Successful treatment with fecal microbiota transplantation in patients with multiple organ dysfunction syndrome and diarrhea following severe sepsis. Crit Care 20(1):332 28. Brandt LJ, Aroniadis OC (2013) An overview of fecal microbiota transplantation: techniques, indications, and outcomes. Gastrointest Endosc 78(2):240–249