Reviews of Physiology, Biochemistry and Pharmacology [1st ed.] 9783030615062, 9783030615079

Leading researchers are specially invited to provide a complete understanding of a key topic within the multidisciplinar

1,475 189 4MB

English Pages VII, 110 [116] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Reviews of Physiology, Biochemistry and Pharmacology [1st ed.]
 9783030615062, 9783030615079

Table of contents :
Front Matter ....Pages i-vii
CXC Chemokine Receptors in the Tumor Microenvironment and an Update of Antagonist Development (Yang Xun, Hua Yang, Jiekai Li, Fuling Wu, Fang Liu)....Pages 1-40
Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal Dystrophies (Darpan Malhotra, Joseph R. Casey)....Pages 41-81
Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In Vivo, and Clinical Studies (Alireza Fatahian, Saeed Mohammadian Haftcheshmeh, Sara Azhdari, Helaleh Kaboli Farshchi, Banafsheh Nikfar, Amir Abbas Momtazi-Borojeni)....Pages 83-110

Citation preview

Stine Helene Falsig Pedersen  Editor

Reviews of Physiology, Biochemistry and Pharmacology 178

Reviews of Physiology, Biochemistry and Pharmacology Volume 178

Editor-in-Chief Stine Helene Falsig Pedersen, Department of Biology, University of Copenhagen, Copenhagen, Denmark Series Editors Emmanuelle Cordat, Department of Physiology, University of Alberta, Edmonton, Alberta, Canada Diane L. Barber, Department of Cell and Tissue Biology, University of California San Francisco, CA, USA Jens Leipziger, Department of Biomedicine, Aarhus University, Aarhus, Denmark Luis A. Pardo, Max Planck Institute for Experimental Medicine, G€ ottingen, Germany Christian Stock, Department of Gastroenterology, Hannover Medical School, Hannover, Germany Nicole Schmitt, Department of Biomedical Sciences, University of Copenhagen, Copenhagen, Denmark Martha E. O’Donnell, Department of Physiology and Membrane Biology, University of California, Davis School of Medicine, Davis, CA, USA

The highly successful Reviews of Physiology, Biochemistry and Pharmacology continue to offer high-quality, in-depth reviews covering the full range of modern physiology, biochemistry and pharmacology. Leading researchers are specially invited to provide a complete understanding of the key topics in these archetypal multidisciplinary fields. In a form immediately useful to scientists, this periodical aims to filter, highlight and review the latest developments in these rapidly advancing fields. 2019 Impact Factor: 4.700, 5-Year Impact Factor: 6.000 2019 Eigenfaktor Score: 0.00067, Article Influence Score: 1.570 More information about this series at http://www.springer.com/series/112

Stine Helene Falsig Pedersen Editor

Reviews of Physiology, Biochemistry and Pharmacology

Editor Stine Helene Falsig Pedersen Department of Biology University of Copenhagen Copenhagen, Denmark

ISSN 0303-4240 ISSN 1617-5786 (electronic) Reviews of Physiology, Biochemistry and Pharmacology ISBN 978-3-030-61506-2 ISBN 978-3-030-61507-9 (eBook) https://doi.org/10.1007/978-3-030-61507-9 © The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are solely and exclusively licensed 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 Switzerland AG. The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

Acknowledgements

Contributions to this volume have partly been personally invited, with kind support of the series editors D.L. Barber, E. Cordat, J. Leipziger, M.E. O’Donnell, L. Pardo, N. Schmitt, C. Stock.

v

Contents

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of Antagonist Development . . . . . . . . . . . . . . . . . . . . . . . Yang Xun, Hua Yang, Jiekai Li, Fuling Wu, and Fang Liu

1

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal Dystrophies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 Darpan Malhotra and Joseph R. Casey Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In Vivo, and Clinical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 Alireza Fatahian, Saeed Mohammadian Haftcheshmeh, Sara Azhdari, Helaleh Kaboli Farshchi, Banafsheh Nikfar, and Amir Abbas Momtazi-Borojeni

vii

Rev Physiol Biochem Pharmacol (2020) 178: 1–40 https://doi.org/10.1007/112_2020_35 © The Author(s), under exclusive licence to Springer Nature Switzerland AG 2020 Published online: 21 August 2020

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of Antagonist Development Yang Xun, Hua Yang, Jiekai Li, Fuling Wu, and Fang Liu

Contents 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 CXCR1 and CXCR2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Cell Signaling Induced by CXCR1 and CXCR2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 CXCR1 and CXCR2 in Tumor Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 CXCR1 and CXCR2 in Immune Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Antagonist Development Targeting CXCR1 and CXCR2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 CXCR3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 CXCR3 Structure and Signaling Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Autocrine and Paracrine Mechanisms of CXCR3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Antagonists Targeting CXCR3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 CXCR4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1 CXCR4 Signaling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 CXCR4 Expression and Signaling in Tumor Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Association Between CXCR4 and Leukocyte Trafficking . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Antagonists Targeting CXCR4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 CXCR5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 CXCR5 Expression Patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 CXCR5 in Tumor Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 CXCR5 in Immune Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4 Absence of Antagonists Targeting CXCR5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 CXCR6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1 Expression of CXCR6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2 CXCR6–TM-CXCL16 and CXCR6–sCXCL16 Axes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3 CXCR6 Expression by Tumor Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4 Association Between CXCR6 and Immune Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Yang Xun and Hua Yang contributed equally to this work. Y. Xun, H. Yang, J. Li, and F. Liu (*) Department of Basic Medicine and Biomedical Engineering, School of Stomatology and Medicine, Foshan University, Foshan, Guangdong Province, China e-mail: [email protected] F. Wu Department of Pharmacy, Nanfang Hospital, Southern Medical University, Guangzhou, Guangdong Province, China

3 4 4 6 6 10 13 13 13 15 16 16 16 17 17 18 18 19 19 20 20 20 20 21 22

2

Y. Xun et al.

6.5 Absence of Antagonist Targeting CXCR6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 CXCR7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1 CXCR7 Signaling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2 CXCR7 in Tumor Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3 CXCR7 and Drug Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4 Crosstalk of CXCR7 with CXCR3 and CXCR4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.5 Antagonist Development Targeting CXCR7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

22 23 23 23 24 24 25 25 29

Abstract Chemokine receptors, a diverse group within the seven-transmembrane G protein-coupled receptor superfamily, are frequently overexpressed in malignant tumors. Ligand binding activates multiple downstream signal transduction cascades that drive tumor growth and metastasis, resulting in poor clinical outcome. These receptors are thus considered promising targets for anti-tumor therapy. This article reviews recent studies on the expression and function of CXC chemokine receptors in various tumor microenvironments and recent developments in cancer therapy using CXC chemokine receptor antagonists. Keywords CXC antagonists · CXC chemokine receptors · Tumor microenvironment

Abbreviations ADAM10 Akt AML CCRCC CLL CRPC CTL CXCR DLBCL ECM EGFR EMT ERK HCC HER2 HIV IgG IL JNK MAPK MDSC

Disintegrin-like metalloproteinase 10 Protein kinase B Acute myeloid leukemia Clear cell renal cell carcinoma Chronic lymphocytic leukemia Castrated-resistant prostate cancer Cytotoxicity T-cell CXC chemokine receptor Diffuse large B-cell lymphoma Extracellular matrix Epidermal growth factor receptor Epithelial-to-mesenchymal transition Extracellular signal-regulated kinase Hepatocellular carcinoma Epidermal growth factor receptor type 2 Human immunodeficiency virus Immunoglobulin G Interleukin c-Jun N-terminal kinases Mitogen-activated protein kinase Myeloid-derived suppressor cells

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

MMP NHERF1 NHL NK NKT NSCLC PD-1 PDAC PI3K PMN PTEN sCXCL16 STAT TAK1 TAM Tfh TIL TM-CXCL16 TNBC Treg Trm VEGF

3

Matrix metalloproteinases Na+/H+ exchanger regulatory factor 1 Non-Hodgkin’s lymphoma Natural killer Natural killer T-cell Non-small cell lung cancer Programmed death-ligand 1 Pancreatic ductal adenocarcinoma Phosphoinositide 3-kinase Polymorphonuclear Phosphatase and tensin homolog Soluble CXCL16 Signal transducers and activators of transcription Transforming growth factor-β-activated kinase 1 Tumor-associated microglia/macrophage T-follicular helper cells Tumor-infiltrating lymphocyte Transmembrane CXCL16 Triple-negative breast cancer Regulatory T-cells Resident memory T-cells Vascular endothelial growth factor

1 Introduction Tumor development is strongly dependent on factors in the surrounding matrix or tumor microenvironment, including anchored and diffusible signaling factors such as hormones, growth factors, and cytokines. These factors collectively constitute a complex molecular network, regulating tumor cell proliferation, motility, and survival. The relationship between tumor cells and the microenvironment has been equated to that between seed and soil, as the seed can only grow in congenial soil (Kenny et al. 2007). Nowadays, tumor microenvironment has been intensively studied as a hot spot in aiding the development of anti-tumor drugs, among which the roles of chemokines and their receptors have attracted many attentions. The chemokine receptors constitute a large and diverse subgroup within the seven-transmembrane G protein-coupled receptor superfamily. Based on the sequence of conserved cysteine residues at the N-terminal of the ligand molecule, these receptors are divided into four subtypes, CR, CCR, CXCR, and CX3CR, where C is cysteine and X is an arbitrary amino acid. Chemokine receptors bind specific ligands and transmit extracellular information into cells through structural changes of the G protein to regulate cell growth, differentiation, and apoptosis among other functions under physiological and pathological conditions (Fig. 1).

4

Y. Xun et al.

Fig. 1 Chemokine ligand–receptor regulatory mechanisms. Chemokine ligands (red) bind to receptors (blue) to induce signal transduction pathways that activate a myriad of physiological and pathological processes

Chemokine receptors are expressed in many tumors and are strongly linked to malignancy. For example, CCR7 promotes lung cancer cell proliferation and lymph node metastasis (Zhang et al. 2013), CX3CR1 expression is associated with poor prognosis of colorectal cancer (Erreni et al. 2010), and CCR2 contributes to the changes in CDb/Gr myelogenous cell number and influences tumor load in colorectal cancer liver metastasis (Zhao et al. 2013). CXC chemokine receptors (CXCRs) (Table 1) in the tumor microenvironment act as seminal regulators of tumor progression by binding to unique ligands or subgroups of ligands (Fig. 2). This review focuses on the expression and function of CXCRs in various tumor microenvironments as well as on the development of CXCR antagonists as anti-tumor drugs.

2 CXCR1 and CXCR2 2.1

Cell Signaling Induced by CXCR1 and CXCR2

CXCR1 and CXCR2 are closely related receptors, also known as interleukin-8receptor A and B, respectively, as both bind interleukin (IL)-8 (CXCL8). These receptors share 77% sequence homology and an ELR motif immediately adjacent to the CXC motif, and both receptors bind CXCL6 and CXCL8, while only CXCR2 binds CXCL1–3, CXCL5, and CXCL7. Two conserved domains at the CXCR2 carboxyl end (ILXLL and PDZ-ligand domains) are critical for signal transduction (Baugher and Richmond 2008). The three dimensional structures of monomeric

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

5

Table 1 Basic information of CXC chemokine receptors in human (Homo sapiens)

Name CXCR1 CXCR2

CXCR3A CXCR3B CXCR4

CXCR5 CXCR6 CXCR7

Alias IL8RA, IL8R1, CD181, CD128, CKR-1, CMKAR1 IL8RB, IL8R2, CD182, CDw128b, CMKAR2

Location 2q35

Number of amino acids 350

2q35

360

40,759

CD183, CKR-L2, CMKAR3, GPR9, IP10-R, Mig-R

Xq13

368

40,659

CD184, D2S201E, FB22, HM89, HSY3RR, LAP3, LCR1, LESTR, NPY3R, NPYR, NPYRL, NPYY3R, WHIM CD185, BLR1, MDR15 CD186, BONZO, STRL33, TYMSTR ACKR3, RDC1, CMKOR1, GPR159

2q21

415 360

45,523 40,607

11q23 3p21

372 342

41,925 39,280

– –

2q37

362

41,478



Protein size 39,771

Structure solved Full monomeric structure C-terminal CXCR2NHERF1 PDZ2 complex structure Fragmental structure – CXCR4– CXCL12 complex structure

“–” indicates unavailable structural information

Fig. 2 CXC chemokine receptors (blue) and their corresponding CXC chemokine ligands (gray)

CXCR1 in phospholipid bilayers and the C-terminal of CXCR2 in complex with the Na+/H+ exchanger regulatory factor 1 (NHERF1) PDZ2 domain have been described (Park et al. 2012; Lu et al. 2013), providing crucial information on CXCR1/2-ligand interaction sites for antagonist design. Due to the high level of homology between CXCR1 and CXCR2, the two receptors are usually investigated together.

6

2.2

Y. Xun et al.

CXCR1 and CXCR2 in Tumor Cells

High expression levels of CXCR1 and/or CXCR2 are usually associated with accelerated tumor development (Fig. 3). Both CXCR1–CXCL8 and CXCR2– CXCL8 signaling axes promote lymphatic and distant metastasis of liver cancer (Bi et al. 2019). Overexpression of CXCR1 and CXCR2 also promotes metastasis and chemoresistance of breast cancer cells via epithelial-to-mesenchymal transition (EMT) and neutrophil infiltration, processes mediated by epidermal growth factor receptor type 2 (HER2), protein kinase B (Akt), and cyclooxygenase-2 among other signaling factors (Kaunisto et al. 2015; Xu et al. 2018a; Shah and Osipo 2016). The CXCR2–CXCL8 axis stimulates the growth and proliferation of prostate cancer cells by activating cyclin D via phosphoinositide 3-kinase (PI3K)/Akt/mTOR and mitogen-activated protein kinase (MAPK) pathways (MacManus et al. 2007). In addition, the CXCR2–CXCL7 axis promotes angiogenesis in metastatic colorectal cancer, and high expression of CXCR2 is predictive of poor survival (Desurmont et al. 2015). The CXCR2–CXCL5 axis is a key inducer of EMT and metastatic colonization of bone from breast cancer via Raf/MAPK/extracellular signalregulated kinase (ERK)/Snail signaling (Romero-Moreno et al. 2019) and liver cancer metastasis via PI3K/Akt/GSK-3β/Snail signaling (Zhou et al. 2015). CXCR2 interacts with different CXC ligands to mediate IL-17A-induced vascular endothelial growth factor (VEGF)-independent tumor angiogenesis and endothelial chemotaxis in liver cancer (Liu et al. 2019a). Hypoxia-induced upregulation of CXCR1 and CXCR2 by malignant melanoma cells also enhances CXCL8dependent tumor cell proliferation, invasion, and angiogenesis (Gabellini et al. 2009).

2.3

CXCR1 and CXCR2 in Immune Cells

Both CXCR1 and CXCR2 are expressed on immune cells and facilitate recruitment to the tumor microenvironment, thereby influencing tumor progression (Table 2). In multiple tumor types, CXCR1+ Foxp3+ CD4+ regulatory T-cells (Tregs) promote tumor cell migration toward a CXCL8 gradient upon IL-6 stimulation (Eikawa et al. 2010). In liver cancer and breast cancer, CXCR2 expressed on neutrophils regulates infiltration into tumoral and peri-tumoral stroma, which in turn promotes tumor progression and metastasis mediated by the CXCR2–CXCL1 axis (Li et al. 2015). Further, this infiltration can be stimulated by TNFα-activated mesenchymal stromal cells (Yu et al. 2017). A secretion loop is formed in colorectal cancer when tumor suppressor gene SMAD4 stimulates CXCL1 and CXCL8 secretion to recruit CXCR2+ neutrophils, which produce more CXCL1 and CXCL8, resulting in further CXCR2+ neutrophil infiltration to the tumor microenvironment (Ogawa et al. 2019). Such infiltration at the tumor nest subsequently induces the secretion of inflammatory and angiogenic factors, such as matrix metalloproteinases (MMPs) and VEGF.

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

7

Fig. 3 Expression of CXC chemokine receptors in various normal and tumor tissues. The X-axis represents the expression levels of CXCRs on a log2 scale. The Y-axis represents different normal (blue), tumor (red), and metastatic tumor (purple) tissues. ACC adrenocortical carcinoma, BLCA bladder cancer, BRCA breast cancer, CESC cervical squamous cell carcinoma, CHOL cholesterol tumor, COAD colon adenocarcinoma, DLBC Diffuse large B-cell lymphoma, ESCA esophageal carcinoma, GBM glioblastoma multiforme, HNSC head and neck squamous cell carcinoma, KICH kidney chromophobe, KIRC kidney renal clear cell carcinoma, KIRP kidney renal papillary cell carcinoma, LAML acute myeloid leukemia, LGG brain lower grade glioma, LIHC liver hepatocellular carcinoma, LUAD lung adenocarcinoma, LUSC lung squamous cell carcinoma, MESO

8

Y. Xun et al.

Fig. 3 (continued) mesothelioma, OV ovarian serous cystadenocarcinoma, PAAD pancreatic adenocarcinoma, PCPG pheochromocytoma and paraganglioma, PRAD prostate adenocarcinoma, READ rectum adenocarcinoma, SARC Sarcoma, SKCM skin cutaneous melanoma, STAD stomach adenocarcinoma, TGCT testicular germ cell tumors, THCA thyroid carcinoma, THYM thymoma, UCEC uterine corpus endometrial carcinoma, UCS uterine carcinosarcoma, UVM uveal melanoma. The analysis was performed by our group

In addition, elevated CXCR1 and CXCR2 expression levels have been detected in tumor polymorphonuclear (PMN) myeloid-derived suppressor cells (MDSCs), which are known to reduce T-cell infiltration and thus worsen clinical outcome (Zhao et al. 2015; Najjar et al. 2017; Sun et al. 2019). Knockdown of CXCR1 and/or CXCR2 can effectively inhibit tumor development in vitro and in vivo. For instance, CXCR2 knockout in mice significantly suppresses neutrophil infiltration, tumor growth, angiogenesis, and metastasis of

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

9

Table 2 CXCRs expressed on immune cells and their impacts on tumor progression

CXCR2

CXCR+ immune cells Foxp3+ CD4 Treg cells Neutrophils

CXCR1/ CXCR2

MDSC cells

CXCR3

NK cells

CXCR CXCR1

CXCR4

CXCR5

Tumor Lung cancer, malignant mesothelioma, melanoma Colorectal cancer, HCC, pancreatic cancer, breast cancer HCC, renal cell carcinoma, oral carcinoma, lung carcinoma Lymphoma

Dendritic cells Treg cells

Gastric cancer

B-cells

HCC

MDSC cells NK cells

Osteosarcoma

CD4+ T-cells CD8 T-cell

NSCLC

Tfh cells

CRC

Glioma

Thyroid cancer, HCC, follicular lymphoma, colorectal cancer, pancreatic cancer HCC

MDSC cells

Gastric cancer

CD4 T-cells

DLBCL

Outcomes Elevate tumor cell migration

Ref. (Eikawa et al. 2010)

Induce secretion of inflammatory and angiogenic factors, reduce T-cell infiltration, poor prognosis Reduce T-cell infiltration, poor clinical outcomes

(Li et al. 2015; Yu et al. 2017; Ogawa et al. 2019; Nywening et al. 2018) (Zhao et al. 2015; Najjar et al. 2017; Sun et al. 2019)

Increase NK cell infiltration, anti-tumor effects Increase TIL infiltration, anti-tumor effects Induce cancer initiating cell production Induce cytokine production and reduce antitumor T-cell response Reduce CTL activity and anti-tumor effects Increase anti-tumor effects Increase tumor cell migration Increase cytotoxicity and anti-tumor effects

(Wendel et al. 2008)

Increase IL-21 secretion and anti-tumor effects Inhibit T-cell expansion and promote tumor growth Upregulate IL-10 and downregulate IL-21 to promote tumor progression

(Chen et al. 2018) (Yang et al. 2011) (Wei et al. 2019)

(Jiang et al. 2019b) (Muller et al. 2015) (Wald et al. 2006) (Zhou et al. 2018b; Jin et al. 2017a; Chu et al. 2019; Bai et al. 2017; Jifu et al. 2018) (Duan et al. 2015) (Ding et al. 2015)

(Cha et al. 2017)

(continued)

10

Y. Xun et al.

Table 2 (continued)

CXCR CXCR6

CXCR+ immune cells NK cells

Tumor Breast cancer

CD4+ T-cells CD8+ T-cells

Colorectal cancer, NSCLC Colorectal cancer, NSCLC

Outcomes Increase anti-tumor effects – –

Ref. (Yoon et al. 2016) (Lofroos et al. 2017; Oja et al. 2018) (Lofroos et al. 2017; Oja et al. 2018)

The functions of tumor-associated CXCR7+ immune cells have not been investigated in detail and thus are not listed in the table. “–” is an indication for unavailable information

several cancers (Keane et al. 2004; Singh et al. 2009a; Lee et al. 2012). In addition, CXCR1 knockdown effectively reduced CXCL8 expression and Akt signaling in osteosarcoma (Han et al. 2015). Thus, both receptors may be promising targets for anti-tumor therapy.

2.4

Antagonist Development Targeting CXCR1 and CXCR2

Over the last few years, tumor therapeutic strategies targeting CXCR1 and CXCR2 have shown promising results (Table 3). The dual CXCR1/CXCR2 small-molecule antagonist SCH-527123, also known as navarixin or MK-527123, has been shown to inhibit melanoma cell proliferation, migration, and invasion via blockade of PI3K/ Akt and MAPK/ERK pathways (Shang and Li 2018; Singh et al. 2009b) and also to inhibit CXCR2 signal transduction through deactivation of NF-κB/MAPK/Akt signaling, leading to colorectal cancer cell apoptosis (Ning et al. 2012). Application of SCH-527123 together with another dual receptor antagonist, SCH-479833, has been investigated for the treatment of spontaneous liver metastasis of colon cancer (Varney et al. 2011). In preclinical colon cancer models, combination of SCH-527123 with oxaliplatin, a standard therapy for colorectal cancer, exhibits stronger inhibition of tumor cell proliferation and angiogenesis compared with using either therapy alone (Ning et al. 2012). At present, a clinical phase 2 trial (NCT03473925) is investigating the efficacy and safety of SCH-527123 combined with pembrolizumab, a humanized antibody targeting the inhibitory T-cell receptor programmed death-ligand 1 (PD-1), for treatment of non-small cell lung cancer (NSCLC), castrated-resistant prostate cancer (CRPC), and microsatellite stable of colorectal cancer. Reparixin, another dual antagonist that reduces tumor volume and metastasis, augments the therapeutic effects of paclitaxel (Schott et al. 2017) or 5-fluorouracil (Wang et al. 2016) for treatment of metastatic breast cancer and gastric cancer, respectively. Apart from these dual functional antagonists, antagonists specifically targeting CXCR2 have been designed, including SB225002 (Manjavachi et al. 2010), which

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

11

Table 3 Summary of antagonists targeting CXC receptors and corresponding outcomes

Targets CXCR1/ CXCR2

Antagonist SCH-527123

SCH-479833

CXCR2

Type of cancers Melanoma, colorectal cancer, pancreatic cancer, NSCLC, CRPC Melanoma, colorectal cancer, pancreatic cancer

Outcomes Reduced tumor growth, metastasis, angiogenesis

Reduced tumor growth, metastasis, and angiogenesis

SB225002

Prostate cancer, ovarian cancer, esophageal cancer, pancreatic cancer, nasopharyngeal cancer, CCRCC, intrahepatic bile duct cell cancer, breast cancer

Reduced tumor growth and metastasis

AZD-5069 SB332235

Prostate cancer Esophageal cancer, AML

AZ10397767

Type 2 alveolar epithelial adenocarcinoma

CXCR3

AMG 487

Breast cancer, osteosarcoma, colon cancer

– Reduced tumor invasion, viability and clonogenic capacity Reduced chemotaxis of neutrophils to cancer cells, reduced tumor proliferation and microvascular formation Reduced tumor metastasis

CXCR4

AMD3100

Osteosarcoma, colon cancer, lung cancer, prostate cancer, ovarian cancer, NHL, multiple myeloma, B acute lymphoblastic leukemia

Inhibited tumor migration and metastasis

Ref. (Singh et al. 2009b; Varney et al. 2011; Purohit et al. 2016) (Singh et al. 2009b; Varney et al. 2011; Purohit et al. 2016) (Xu et al. 2018b; Yung et al. 2018; Wang et al. 2006; Sueoka et al. 2014; Saintigny et al. 2013; Matsuo et al. 2009; Lo et al. 2013; Ijichi et al. 2011; Grepin et al. 2014; Erin et al. 2015) – (Schinke et al. 2015; Shrivastava et al. 2014)

In vitro or in vivo Both

Both

Both

– In vitro

(Tazzyman et al. 2011)

Both

(Cambien et al. 2009; Zhu et al. 2015a; Pradelli et al. 2009) (Randhawa et al. 2016; Liao et al. 2015; Reeves et al. 2017; Zhu et al. 2019; Bachelerie et al. 2014; Li et al. 2017a)

In vivo

Both

(continued)

12

Y. Xun et al.

Table 3 (continued)

Targets

CXCR7

Antagonist AMD070

Type of cancers Oral cancer, acute lymphoblastic leukemia, pancreatic cancer

Outcomes Reduced tumor invasiveness and metastasis

Ref. (Uchida et al. 2018; Morimoto et al. 2016; Parameswaran et al. 2011)

MDX-1338

AML, CLL, multiple myeloma

IT1t

Triple-negative breast cancer

(Kashyap et al. 2016; Kuhne et al. 2013) (Tulotta et al. 2016)

POL6326 BKT140

Breast cancer AML, multiple myeloma, NSCLC, NHL, prostate cancer, pancreatic cancer

Induced apoptosis and inhibited proliferation Reduced formation of early metastasis – Inhibited tumor proliferation and metastasis

X4-136

Melanoma

CCX771

Prostate cancer, liver cancer, breast cancer

CCX662

Glioblastoma multiforme

Inhibited tumor growth Reduced tumor growth, migration, and angiogenesis –

– (Fahham et al. 2012; Beider et al. 2013, 2014; Beider et al. 2011; Peng and Kopecek 2014) (Saxena et al. 2019) (Lin et al. 2014; Luo et al. 2018; Qian et al. 2018) –

In vitro or in vivo Both

Both

In vivo

In vivo Both

In vivo Both

In vivo

inhibit the expression of invasion-related proteins, such as bovine seminal plasma, osteopontin, MMPs, and αvβ3 by prostate cancer cells. This leads to blockade of Akt and mTOR activation and reduction of tumor cell proliferation and invasion (Xu et al. 2018b). In addition, SB225002 significantly attenuated the carcinogenic and metastatic potential of ovarian cancer both in vitro and in vivo, likely by inhibiting transforming growth factor-β-activated kinase 1 (TAK1)/NF-κB signal transduction (Yung et al. 2018). In deltaNp63+ triple-negative breast cancer (TNBC) patients, application of SB225002 reduced MDSC recruitment and tumor metastasis (Kumar et al. 2018). A recent study also reported that the CXCR2 antagonist AZD5069 aids in activating CXCR2+ tumor-associated macrophages, which promote an anti-tumorigenic state upon infiltration (Di Mitri et al. 2019). Combination of AZD5069 with anti-PD-1 antibody prolonged survival of mice with pancreatic ductal adenocarcinoma (PDAC) by improving T-cell infiltration (Steele et al. 2016). Another such CXCR2 antagonist, AZ10397767, developed to specifically inhibit CXCR2-mediated neutrophil chemotaxis, reduced neutrophil infiltration into

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

13

xenografts of type 2 alveolar epithelial adenocarcinoma, thereby indirectly suppressing tumor cell proliferation and angiogenesis (Tazzyman et al. 2011).

3 CXCR3 3.1

CXCR3 Structure and Signaling Mechanisms

CXCR3, a predominantly IFN-γ-driven CXC chemokine receptor, interacts with CXCL9–11 to regulate immune cell proliferation, cytokine secretion, and migration (Nakajima et al. 2002). In addition to the structures conserved across all CXCRs, CXCR3 has two different activation domains, one targeting the C-terminal domains of CXCL9 and CXCL10 and the other targeting the third intracellular ring of CXCL11 (Colvin et al. 2004). Structural analysis has revealed three CXCR3 variants, CXCR3A, CXCR3B, and CXCR3-ALT (Lasagni et al. 2003), with distinct intracellular signaling functions. Variant CXCR3A functions as the classical CXCR3, while CXCR3B binds to CXCL9–11 as well as CXCL4 to inhibit endothelial cell growth, which may contribute to suppression of tumor angiogenesis (Lasagni et al. 2003; Puchert et al. 2019), and CXCR3-ALT is a 101-amino acidtruncated CXCR3 that mainly mediates CXCL11 signaling (Lasagni et al. 2003). The CXCR3A variant (also referred to as simply CXCR3) has been the major focus of research because of its known contributions to tumor development. CXCR3 is expressed not only by immune cells, including macrophages, T-cells, natural killer (NK) cells, and dendritic cells but is also detected in various tumor cells (Fig. 3) (Ma et al. 2009; Zhou et al. 2016; Winkler et al. 2011; Cambien et al. 2009; Maekawa et al. 2008; Rezakhaniha et al. 2016).

3.2

Autocrine and Paracrine Mechanisms of CXCR3

Signaling by CXCR3 can be either autocrine or paracrine (Fig. 4). CXCR3 overexpression in cancer cells tends to elevate tumor proliferation, migration, metastasis, and angiogenesis through autocrine signaling, resulting in poor clinical outcomes, while immune cells expressing CXCR3 are recruited to the tumor nest through paracrine signaling, leading to either anti-tumor or pro-tumor effects depending on the recruited population (Table 2). Autocrine signaling: Overexpression of CXCR3 by breast cancer cells is associated with tumor progression and poor survival (Bronger et al. 2017). Binding to CXCL9 or CXCL10 elevates the expression of cathepsin B, a protease that degrades chemokine ligands. This creates a negative feedback loop that disrupts the chemokine gradient at the tumor nest and suppresses CXCR3+ lymphocyte infiltration, resulting in tumor progression (Bronger et al. 2017). High CXCR3 expression also induces MMP and IL-6 activation, resulting in degradation of extracellular matrix

14

Y. Xun et al.

Fig. 4 Functions of CXCR3 in the tumor microenvironment. CXCR3 signals through autocrine and paracrine mechanisms to either promote or suppress tumor development. CXCR3 expressed on tumor cells interacts with tumor-derived chemokines via autocrine signaling to trigger tumor cell proliferation, migration, and metastasis. CXCR3+ leukocytes are differentiated and recruited toward tumor sites via paracrine signaling to induce or suppress tumor growth

(ECM) proteins, which ultimately leads to vascular invasion, lymph node metastasis, and poor survival (Zhou et al. 2016; Yang et al. 2016; Shin et al. 2010). Paracrine signaling: In a mouse model of lymphoma, secretion of CXCL10 by tumor cells increased CXCR3+ CD27+ NK cell infiltration to tumor sites through CXCR3–CXCL10 signaling, prolonging survival (Wendel et al. 2008). Similarly, paracrine signaling recruits CXCR3+ dendritic cells and tumor-infiltrating lymphocytes (TILs) to gastric cancer tissues, resulting in anti-tumor activity (Chen et al. 2018). High expression of CXCR3 is also correlated with low pro-tumor M2 infiltration into gastric cancer (Chen et al. 2019a); however, detailed mechanisms

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

15

were not clarified. In contrast, elevated CXCR3 expression promotes CXCR3+ Foxp3+ IL-17 Treg cell infiltration toward high CXCL11 gradient colorectal tumor tissues, which then triggers IL-17-dependent production of cancer initiation cells (Yang et al. 2011). In hepatocellular carcinoma (HCC), CXCR3 expressed on B-cells interacts with CXCL10 secreted by macrophages upon CD4+ T-cell activation, causing CXCR3+ B-cells to produce immunoglobulin G (IgG), leading to elevated cytokine expression by macrophage and a reduced anti-tumor T-cell response (Wei et al. 2019). Due to the bidirectional function of CXCR3 in tumor promotion and suppression, CXCR3 knockdown may either suppress or stimulate tumor progression and metastasis. In mouse metastatic breast cancer and melanoma models, CXCR3 depletion effectively reduces tumor progression and metastasis (Zhu et al. 2015a; Kawada et al. 2004). Alternatively, CXCR3 deficient mice exhibited macrophage polarization toward the M2 phenotype and IL-12/IFN-γ/NO axis disruption, resulting in a tumor-promoting microenvironment (Oghumu et al. 2014; Chheda et al. 2016). In addition, although CXCR3B has been reported to inhibit tumor angiogenesis via CXCR3B–CXCL10 signaling (Proost et al. 2001), tumor occurrence is more likely when the expression of CXCR3A is dominant over CXCR3B (Kawada et al. 2007; Goldberg-Bittman et al. 2004). Recently, Saahene et al. (Saahene et al. 2019) suggested that CXCR3B interaction with CXCL4 may be an indicator of poor prognosis. Thus, CXCR3 has diverse functions in the tumor microenvironment that can enhance or suppress tumor progression.

3.3

Antagonists Targeting CXCR3

The fragmental structures of CXCR3–CXCL complexes (Palladino et al. 2012; Trotta et al. 2009) have greatly assisted in the development of antagonists targeting CXCR3 for anti-tumor therapy (Table 3). The small-molecule antagonist of all CXCR3 variants AMG487 exhibits significant inhibitory effects on tumor progression and metastasis and improves host anti-tumor immunity. Experiments on a rat breast cancer model revealed that this anti-tumor mechanism involves reduction of tumor cell migration and improvement of both myeloid cell function and T-cell response (Zhu et al. 2015a). AMG487 also prevented CXCR3-induced migration of colorectal cancer cells in mouse model of colon carcinoma and lung metastasis (Cambien et al. 2009) and blocked MMP-2 and MMP-9 activity induced by the CXCR3–CXCL10 axis, thereby suppressing lung metastasis of osteosarcoma (Pradelli et al. 2009). Other CXCR3 antagonists include SCH-546738 (Du et al. 2017), Takeda 779 (Takama et al. 2011), and NBI-74330 (Gao et al. 2018). However, the anti-tumor effects of these agents have not been assessed.

16

Y. Xun et al.

4 CXCR4 4.1

CXCR4 Signaling

CXCR4 was first reported as a highly efficient lymphocyte chemoattractant in 1996 (Bleul et al. 1996). In 2001, Muller et al. (Muller et al. 2001) demonstrated elevated expression of CXCR4 in human breast cancer, which influenced tumor cell migration and metastasis through exclusive interaction with CXCL12. Since these seminal reports, many studies of CXCR4 influences on tumor progression have been conducted, making it the most thoroughly investigated CXC chemokine receptor. To date, CXCR4 expression has been reported in more than 20 tumor types (Fig. 3). Elevated expression of CXCR4 and CXCR4–CXCL12 signaling has diverse effects (Cambien et al. 2009; Muller et al. 2001; Scala et al. 2005; Kaifi et al. 2005; Lu et al. 2014; Oliveira Frick et al. 2011) but is generally a biomarker for poor prognosis (Furusato et al. 2010; Du et al. 2019). CXCR4 is also expressed in immune cells, including macrophages, neutrophils, and T and B lymphocytes, where it regulates infiltration and thus the tumor microenvironment. As there are a number of quality reviews available on CXCR4 functions in the tumor microenvironment (Chatterjee et al. 2014; Walenkamp et al. 2017), this review presents a selection of the most recent studies.

4.2

CXCR4 Expression and Signaling in Tumor Cells

Overexpression of CXCR4 was recently identified as a prognostic biomarker for gastrointestinal cancer, acute myeloid leukemia, and tongue squamous carcinoma in addition to many other tumor types (Du et al. 2019; Jiang et al. 2019a; Ciuca et al. 2019). In ovarian cancer, CXCR4 promotes tumor cell invasion by stimulating MMP production and inhibiting expression of the Rho-activating protein ARHGAP10 via upregulation of VEGF/VEGFR2 signaling (Luo et al. 2019). High expression of CXCR4 in glioblastoma multiforme tissue and glioma-initiating cells modulated by Notch1 through Akt/mTOR signaling promotes stem maintenance and migration of glioma-initiating cells (Yi et al. 2019). Combined inhibition of CXCR4 signal transducers and activators of transcription 3 (STAT3), a transcription factor that converts cytokines from anti-tumor to oncogenic, reduced breast cancer lung metastasis by increasing CD8 T-cell infiltration and downregulating MMPs and VEGF (Li et al. 2019a). In addition, blocking B-cell receptor/PI3K signaling upregulated CXCR4–CXCL12-induced chemotaxis in B-cell receptor-dependent diffuse large B-cell lymphoma (DLBCL), suggesting that CXCR4 expression level may explain tumor sensitivity to B-cell receptor inhibitors (Chen et al. 2019b).

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

4.3

17

Association Between CXCR4 and Leukocyte Trafficking

Signaling through CXCR4 is a crucial controller of tumor leukocyte trafficking and thus a major influence on tumor progression. In osteosarcoma, CXCR4+ MDSCs migrate toward a CXCL12 gradient at the tumor nest, where these cells activate the Akt pathway. In turn, Akt activation reduces MDSC cell apoptosis, forming a positive feedback loop that leads to inhibition of cytotoxic T-cell anti-tumor effects (Jiang et al. 2019b). An in vitro study examining infiltration of NK cells expressing genetically modified CXCR4 found an augmented cytotoxic response against glioma cells via CXCR4–CXCL12 (Muller et al. 2015). In addition, high plasmacytoid dendritic cell infiltration was detected, which stimulated CXCR4 expression via TNF-α/NF-κB signaling. Overexpression of CXCL12 at tumor sites attracts more CXCR4+ cells, forming a positive feedback cycle leading to lymph node metastasis of breast cancer (Gadalla et al. 2019). Alternatively, blocking CXCR4 effectively suppressed tumor metastasis and improved the therapeutic efficacy of an immune checkpoint blocker via upregulation of Treg infiltration, downregulation of tumorassociated fibroblasts, and immunosuppression (Chen et al. 2019c).

4.4

Antagonists Targeting CXCR4

Since elucidation of the CXCR4–CXCL12 complex structure (Wu et al. 2010), a number of antagonists have been developed to target the CXCR4–CXCL12 signaling axis (Table 3). Among all the antagonists available, AMD3100, also known as plerixafor, is the best studied and so far the only one approved by the US Food and Drug Administration for clinical use in treating non-Hodgkin’s lymphoma (NHL) and auto-transplantation of multiple myeloma. AMD3100 was originally designed to antagonize human immunodeficiency virus (HIV) (Donzella et al. 1998) and its potential anticancer efficacy subsequently explored based on studies showing the importance of the CXCR4–CXCL12 axis in tumor progression. Indeed, AMD3100 has been shown to inhibit the migration and adhesion of acute B lymphoblastic leukemia cells (Randhawa et al. 2016), the survival and metastasis of osteosarcoma by blocking c-Jun N-terminal kinase (JNK) and Akt signaling pathways (Liao et al. 2015), and ovarian cancer cell proliferation, while improving the therapeutic efficacy of low-dose paclitaxel and reducing recurrence of ovarian cancer (Reeves et al. 2017), and enhance cellular radiosensitivity and primary tumor response in mouse models of TNBC (Zhou et al. 2018a) and cervical cancer in radiation therapy (Lecavalier-Barsoum et al. 2018). In addition, AMD3100 suppressed EMT and prostate cancer cell migration by inhibiting the CXCR4–CXCL12α pathway (Zhu et al. 2019). It also reduced lung cancer cell proliferation and invasion and suppressed brain metastasis by diminishing the expression levels of CXCR4, VEGF, and MMPs (Li et al. 2017a).

18

Y. Xun et al.

Combination of CXCR4 antagonist and immune checkpoint inhibitor has been reported to improve anti-tumor immune response by reshaping the tumor microenvironment. Dual blockade of CXCR4–CXCL12 and PD-1–PDL-1 pathways using AMD3100 and anti-PD-1 antibody significantly promoted anti-tumor immunity in PDAC (Seo et al. 2019), osteosarcoma (Jiang et al. 2019b), and ovarian cancer (Zeng et al. 2019). An in vitro study showed that immunotherapy alone failed to effectively induce anti-tumor activity in PDAC due to geographic sequestration of T-cells from the tumor site. This can be reversed by CXCR4 antagonist which greatly enhances anti-tumor CD8+ T-cell infiltration in the tumor microenvironment (Seo et al. 2019). In a mouse model of osteosarcoma, application of AMD3100 showed synergistic effect with immunotherapy by diminishing the immunosuppressive CXCR4+ MDSCs within tumor tissues and thus promoting effector T-cell infiltration (Jiang et al. 2019b). Similarly, AMD3100 improved the efficacy of anti-PD-1 and prolonged survival of ovarian tumor-bearing mice by increasing memory and CD8 + T-cell infiltration, increasing conversion from Treg cells into helper T-cells, and decreasing MDSCs (Zeng et al. 2019). Other CXCR4 antagonists include AMD070, MDX-1338, IT1t, BKT140, POL6326, USL-311, T22, and TG-0054. AMD070, also named AMD11070/X4P001, was reported to suppress CXCR4–CXCL12 signaling-dependent migration and invasion of oral cancer cells and to significantly inhibit lung metastasis (Uchida et al. 2018), while MDX-1338, also known as Medarex, was found to inhibit F-actin aggregation and chronic lymphocytic leukemia (CLL) cell migration and induce tumor cell death by producing reactive oxygen species in CLL cells (Kashyap et al. 2016). Further, IT1t was reported to suppress CXCR4 signaling and reduce early metastasis of TNBC (Tulotta et al. 2016), while BKT140 (BL-8040) was shown to improve sensitivity of tumor cells to chemotherapeutic drugs in NSCLC (Fahham et al. 2012), NHL (Beider et al. 2013; Burger et al. 2011), and pancreatic cancer (Bockorny et al. 2020).

5 CXCR5 5.1

CXCR5 Expression Patterns

CXCR5, also known as Burkitt’s lymphoma receptor 1, exclusively interacts with B-cell chemoattractant (CXCL13). CXCR5 is expressed in all mature peripheral B-cells, a small number of T-cell subtypes, mature dendritic cells, and tumor cells (Qi et al. 2014).

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

5.2

19

CXCR5 in Tumor Cells

In CXCR5-expressing tumor cells, the CXCR5–CXCL13 axis activates divergent cell signaling cascades to affect tumor progression and lymph node metastasis, thereby influencing prognosis. The CXCR5–CXCL13 axis was reported to promote the growth of clear cell renal cell carcinoma (CCRCC) (Zheng et al. 2018) and colorectal cancer (Zhu et al. 2015b) via activation of PI3K/AKT/mTOR signaling. This was accompanied by activation of MMP cascades for ECM degradation, thereby enhancing tumor metastasis (Zhu et al. 2015b). Overexpression of CXCR5 upregulates ERK signaling in breast and prostate cancer, thus reducing tumor cell apoptosis (El Haibi et al. 2010; Xu et al. 2018c), and promotes tumor cell proliferation via JNK signaling in prostate cancer (El Haibi et al. 2010; Singh et al. 2009c). In addition, loss of the oncogenic protein PKCε and the tumor suppressor phosphatase and tensin homolog (PTEN) can induce prostate cancer progression through NF-κB-mediated upregulation of CXCR5–CXCL13 signaling (Garg et al. 2017). In contrast, suppressing CXCR5 in colorectal cancer reduces tumor growth and liver metastasis (Meijer et al. 2006). Taken together, these findings suggest that targeting the CXCR5–CXCL13 axis and related signaling pathways may be an effective antitumor strategy.

5.3

CXCR5 in Immune Cells

The main functions of CXCR5–CXCL13 signaling in the immune response are the regulation of B-cells, T-cells, and dendritic cells in secondary lymphoid organs (Schiffer et al. 2015). In turn, the balance of these influences determines tumor suppression or progression. High CXCR5+ CD8 T-cell infiltration was detected at tumor nests and tumor-involved lymph nodes in thyroid cancer (Zhou et al. 2018b), HBV-related HCC (Jin et al. 2017a), follicular lymphoma (Chu et al. 2019), colorectal cancer (Jifu et al. 2018), and pancreatic cancer (Bai et al. 2017). Compared to CXCR5 CD8 T-cells, CXCR5+ CD8 T-cells generally exhibit higher proliferation potency, granzyme B production, PD-1 and TIM-3 expression levels, and cytotoxicity against tumors. In addition, CXCL13 was reported to recruit CXCR5+ tumorinfiltrating T-follicular helper (Tfh) cells into HCC sites, which in turn promoted section of the anti-tumor cytokine IL-21(Duan et al. 2015). In contrast, the CXCR5– CXCL13 axis can also promote escape of tumor cells from immune surveillance. For instance, CXCR5+ CD40+ MDSCs migrate and accumulate at gastric tumor sites via CXCR5–CXCL13 signaling to suppress T-cell expansion and promote tumor growth (Ding et al. 2015). In DLBCL as well, CXCR5+ CD4+ T-cells promote survival and proliferation of tumor cells through upregulation of IL-10 and downregulation of IL-21 (Cha et al. 2017).

20

5.4

Y. Xun et al.

Absence of Antagonists Targeting CXCR5

Little progress has been made in the development of antagonists that directly target CXCR5 due to the lack of CXCR5 structural information, so a more common inhibitory strategy is blockade of the ligand CXCL13. To date, only a few methods are available to target CXCR5 directly for anti-tumor activity. For example, CXCR5::CD3, a trifunctional antibody that targets both CXCR5 and CD3, has been developed to suppress NHL progression via stimulation of CD4+ and CD8+ T-cells, cytokine secretion, and CXCR5+ B-cell lysis (Panjideh et al. 2014). Nevertheless, CXCR5 antagonist development still holds greater promise for anti-tumor therapy.

6 CXCR6 6.1

Expression of CXCR6

CXCR6, originally known as orphan receptor BONZO (or STRL33) as it is the only receptor for CXCL16, is expressed in both normal and tumor tissues (Fig. 3). Investigations of CXCR6 in the tumor microenvironment have only started recently. In 2004, the CXCR6–CXCL16 signaling axis was reported to be significantly downregulated in colon cancer (Wagsater and Dimberg 2004; Wagsater et al. 2004). Since this initial discovery, many studies have investigated CXCR6 functions in regulating the tumor microenvironment.

6.2

CXCR6–TM-CXCL16 and CXCR6–sCXCL16 Axes

Immune signaling by CXCR6 is unique in that this receptor binds to both a full length transmembrane (TM)-ligand (TM-CXCL16) and a secreted soluble (s) ligand (sCXCL16), a shorter version generated by disintegrin-like metalloproteinase 10 (ADAM10). The CXCR6–TM-CXCL16 axis has been reported to inhibit renal and breast cancer progression (Chung et al. 2017; Gutwein et al. 2009), possibly by enhancing immune cell infiltration into tumor sites (Hojo et al. 2007), while CXCR6–sCXCL16 signaling may promote or suppress tumor progression, depending on expression site (Chung et al. 2017; Lang et al. 2017) (Fig. 5).

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

21

Fig. 5 CXCR6 interactions with transmembrane (TM) or soluble (s) CXCL16. TM-CXCL16 undergoes cleavage by ADAM10 to form sCXCL16. (a) CXCR6 interacts with TM-CXCL16 to suppress tumor progression via leukocyte recruitment into the tumor nest. (b) CXCR6 interacts with sCXCL16 to promote or suppress tumor progression via downstream pro-tumor signaling pathways or by leukocyte trafficking, respectively

6.3

CXCR6 Expression by Tumor Cells

CXCR6 is overexpressed in multiple tumor types (Fig. 3), where enhanced signaling promotes tumor cell proliferation, invasion, metastasis, malignant transformation, reoccurrence, advanced clinical stage, and lower patient survival (Mir et al. 2019; Ma et al. 2017; Ke et al. 2017; Chang et al. 2017; Jin et al. 2017b; Hong et al. 2018). Activation of the CXCR6–CXCL16 axis is mainly associated with PI3K/Akt and NF-κB signaling pathways. For example, overexpression of CXCR6 promotes

22

Y. Xun et al.

tumor cell proliferation, invasion, and metastasis through upregulation of Akt signaling in osteosarcoma and gastric cancer (Ma et al. 2017; Jin et al. 2017b), through the PI3K/Akt pathway in ovarian cancer (Hong et al. 2018), and via the NF-κB pathway in prostate cancer (Kapur et al. 2019). This pro-tumor activity is usually associated with elevated EMT and expression levels of MMPs, ADAM10, and CXCL16 (Mir et al. 2019; Ma et al. 2017; Jin et al. 2017b; Hong et al. 2018; Kapur et al. 2019). High expression of CXCR6 is also associated with Ezrin activation and αvβ3 integrin clustering, which promotes gastric cancer by reduced cellular adhesion and concomitant mobilization (Singh et al. 2016). Alternatively, blockade of CXCR6 significantly inhibits tumor growth and potentiates drug efficacy against prostate cancer (Kapur et al. 2019) and HCC (Xu et al. 2014) by suppressing the CXCR6–CXCL16 and VEGF signaling.

6.4

Association Between CXCR6 and Immune Cells

Similar to other CXCR signaling pathways, CXCR6–CXCL16 signaling can recruit immune cells to tumor sites and modulate immune cell phenotype transformation to influence tumor progression. Elevated CXCL16 upon irradiation enhances CXCR6+ NK cell infiltration into tumor sites and promote anti-tumor activity in breast cancer (Yoon et al. 2016). Both CD4+ and CD8+ T-cells also express CXCR6, and the proportions of CXCR6+ CD4+ and CD8+ T-cells differ between normal tissue and colorectal cancer, suggesting that T-cell recruitment to tumors may involve distinct mechanisms (Lofroos et al. 2017). Elevated CXCR6 expression was also detected in CD103+ CD4+ and CD13+ CD8+ T-cells from NSCLC tissues and served as a novel biomarker for resident memory T-cells (Trm), which prevent secondary infection by releasing IFN-γ, TNF-α, and IL-2 (Oja et al. 2018). Immune responses mediated by natural killer T-cells (NKT) and CD4+ T-cells, which are critical for hepatocyte senescence recognition and hepatocarcinogenesis suppression, appear to be CXCR6-dependent. CXCR6-deficient mice exhibit reduced NKT and CD4+ T-cell numbers and concomitantly greater tumor burdens (Mossanen et al. 2019). In glioma, CXCL16 released by tumor cells drives tumorassociated microglia/macrophage (TAM) polarization toward the anti-inflammatory (pro-tumor) phenotype. These cells then infiltrate and stimulate glioma development via the CXCR6–CXCL16 axis. Indeed, CXCR6 knockout greatly suppressed this effect and prolonged survival in mice (Lepore et al. 2018). Unfortunately, the cellspecific functions of CXCR6 are still largely unknown.

6.5

Absence of Antagonist Targeting CXCR6

Overall, the CXCR6–CXCL16 axis presents a promising target for anti-tumor therapy. Although there are several different methods to block CXCR6–CXCL16

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

23

activity, such as by gene silencing or inhibition of related signaling pathways, no information is yet available on CXCR6 antagonist development. Elucidation of CXCR6–CXCL16 structure should greatly aid antagonist design.

7 CXCR7 7.1

CXCR7 Signaling

CXCR7, better known as ACKR3, is an atypical chemokine receptor as it contains an Asp-Arg-Tyr-Leu-Ser-Ile-THR (DRYL-SIT) sequence instead of the classical chemokine receptor motif Asp-Arg-Tyr-Leu-Ala-Ile-Val (DRYLAIV). Due to this alteration, CXCR7 signaling is not mediated through conventional G proteins but rather through β-arrestin2 upon ligand interaction (Naumann et al. 2010). CXCR7 is expressed in normal tissues such as activated endothelial cells (Thelen and Thelen 2008), as well as in the tumor microenvironment upon ligand binding. CXCR7 is a second receptor for CXCL11 and CXCL12 and binds to CXCL12 at a tenfold higher binding affinity compared to CXCR4 (Wang et al. 2018a).

7.2

CXCR7 in Tumor Cells

Both CXCR7–CXCL11 and CXCR7–CXCL12 signaling axes have been reported to participate in tumor progression. Overactivation of the CXCR7–CXCL11 axis upon estrogen stimulation promotes EMT and ECM remodeling, thereby contributing to ovarian cancer progression (Benhadjeba et al. 2018). Significant overexpression of CXCR7 and CXCL12 was also found in lung metastasis of colorectal cancer (Wang et al. 2018b) and breast cancer (Al-Toub et al. 2019) and was associated with elevated cell adhesion to fibronectin and laminin (Qian et al. 2018), triggering tumor metastasis. There is accumulating evidence demonstrating that CXCR7 elevation can induce tumor growth and angiogenesis mainly by facilitating VEGF secretion. This enhanced VEGF secretion is mediated via Akt/MAPK signaling in HCC (Chen et al. 2016) and colon cancer (Li et al. 2019b), the MAPK pathway in gastric cancer (Shi et al. 2017), and the CXCR7–Src kinase axis in melanoma (Xu et al. 2019). In addition, lipopolysaccharide-induced CXCR7 expression promotes gastric cancer cell proliferation and migration via transmembrane Toll-like receptor 4 (TLR4)/myeloid differential protein-2 (MD-2) signaling (Li et al. 2019c). Alternatively, downregulation of CXCR7 suppresses migration and invasion of HCC-derived tumor endothelial cells by inhibiting MMP2 and VEGF via STAT3 signaling (Wu et al. 2018). In addition to interacting with cognate CXC ligands, CXCR7 also binds directly to the epidermal growth factor receptor (EGFR) to trigger activation of MAPK and Akt pathways, leading to tumor cell proliferation in prostate cancer (Singh and Lokeshwar 2011), breast cancer (Salazar et al. 2014), and NSCLC

24

Y. Xun et al.

(Liu et al. 2019b). Downregulation of CXCR7 or dual EGFR/CXCR7 inhibition disrupted this interaction, diminished EGFR and ERK1/2 activities, and suppressed tumor cell proliferation (Salazar et al. 2014; Liu et al. 2019b).

7.3

CXCR7 and Drug Resistance

In recent years, the contributions of CXCR7 signaling to drug resistance have attracted widespread attention. In CRPC, high CXCR7 expression confers resistance to enzalutamide through activation of MAPK/ERK signaling (Li et al. 2019d) and induction of M-phase cell cycle genes through PI3K/Akt signaling upon association with macrophage migration inhibitory factor (Rafiei et al. 2019). Both mechanisms lead to prostate cancer progression. Overexpression of CXCR7 in NSCLC promotes resistance to the EGFR TKI (a class of EGFR inhibitor) and EMT upregulation via MAPK signaling. The combination of CXCR7 depletion and EGFR inhibition diminished ERK activation and improved drug activity (Becker et al. 2019). The expression level of CXCR7 may also correlate with imatinib resistance of chronic myelogenous leukemia through upregulation of ERK signaling (Li et al. 2017b), cisplatin resistance of esophageal squamous cell carcinoma via IL-6-mediated NF-κB signaling (Qiao et al. 2018), and Taxotere resistance of breast cancer (Rong et al. 2013). Collectively, these findings indicate that CXCR7 may serve as a valuable biomarker for drug-resistant cancers.

7.4

Crosstalk of CXCR7 with CXCR3 and CXCR4

Although CXCR7 shares ligands with CXCR3 and CXCR4, downstream signaling is frequently independent, even when these receptors are co-expressed by the same cell. For instance, while CXCR7 co-expresses with CXCR3 or CXCR4 in CRPC, breast cancer, colon cancer, and lung cancer cell lines, activities appear independent (Puchert et al. 2019; Rafiei et al. 2019). In contrast, CXCR7 and CXCR4 co-overexpressed in colorectal cancer cells form heterodimers that induce histone demethylation and promote transcription of oncogenes and inflammatory factors, leading to tumorigenesis (Song et al. 2019). Double knockout of CXCR7 and CXCR4 in TNBC model mice resulted in significantly stronger suppression of tumor cell proliferation, migration, and invasion than either CXCR7 or CXCR4 single knockout (Yang et al. 2019).

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

7.5

25

Antagonist Development Targeting CXCR7

There have been a series of reports on CXCR7-targeted antagonists, such as CCX771 and CCX662 as well as their prototypes CCX733 and CCX754 (Bachelerie et al. 2014). The competitive antagonist CCX771 was found to reduce activation of CXCR7-related downstream signals and subsequently inhibit tumor proliferation and angiogenesis (Qian et al. 2018; Luo et al. 2018). Further, CCX771 diminished EGFR expression in breast cancer cells both in vitro and in vivo (Qian et al. 2018). Combined use of CCX771 and enzalutamide effectively reduced activation of EGFR, Akt, and VEGF in prostate cancer xenografts (Luo et al. 2018). The antagonist CCX733 simultaneously suppressed Akt and ERK signaling pathways, resulting in significantly reduced CXCR7-induced EMT and thus greater therapeutic efficacy against bladder cancer compared to combined use of the ERK inhibitor U0126 and Akt inhibitor LY294002 (Hao et al. 2012). It has been reported that CXCR7 is susceptible to CXCR4 regulation (Wang et al. 2008), and the CXCR4 antagonist AMD3100 also interacts with CXCR7 (Kalatskaya et al. 2009). However, combination therapy using enzalutamide plus AMD 3100 appeared to be less effective than enzalutamide plus CCX771 for inhibition of prostate cancer cell growth and migration (Luo et al. 2018). Finally, CCX662 has been applied for treatment of glioblastoma multiforme (Walters et al. 2014). However, no further information on CCX662 has been reported since 2014.

8 Concluding Remarks Chemokine receptors are expressed by both cancerous cells and immune cells to regulate tumor growth, invasion, metastasis, and angiogenesis through a variety of chemical signals in the tumor microenvironment. Intensive research over the past two decades has revealed the detailed structure, function, and signaling mechanisms of many chemokine receptors. The human seven-transmembrane CXCRs are mid-sized proteins of 350–415 amino acids and 39,280–45,523 Da (Table 1). However, resolving the molecular structures of the transmembrane CXCRs has proven challenging, and to date only the full structures of CXCR1 (Park et al. 2012) and CXCR4 (Wu et al. 2010) have been revealed. This lack of protein structure has greatly hindered investigations on molecular signaling mechanisms and the development of selective chemical modulators such as antagonists. Nonetheless, evidence to date indicates that CXCRs are involved in a variety of tumorigenic processes, including cell proliferation, migration, invasion, metastasis, and angiogenesis, mediated by numerous downstream signaling pathways (Fig. 6), and thus are promising biomarkers and molecular targets for anti-tumor therapy. Elevated expression of CXCRs in tumor cells is usually associated with tumor progression, while CXCR+ tumor-infiltrating immune cells can be anti- or pro-tumorigenic (Table 2). For example, CXCR1+ CD4 Treg cells, CXCR2+

26

Y. Xun et al.

Fig. 6 Network diagram representing the associations between CXCR1–CXCR7 and different signaling pathways

neutrophils, CXCR1+ and CXCR2+ MDSC cells, CXCR3+ Treg cells, CXCR3+ B-cells, CXCR4+ MDSC cells, CXCR4+ CD4+ T-cells, CXCR5+ MDSC cells, and CXCR5+ CD4+ T-cells are generally associated with tumor progression, while CXCR3+ NK cells, CXCR3+ dendritic cells, CXCR4+ NK cells, CXCR5+ CD8+ T-cells, CXCR5+ Tfh cells, and CXCR6+ NK cells are usually reported to have antitumor effects. Studies by Yu (Yu and Zhang 2019) and Lofroos (Lofroos et al. 2017) examining the frequencies of gastric cancer cells expressing CXCR1–7 and lymphocytes expressing CXCR3–6 in colorectal cancer revealed substantial variation in the proportions of different T-cells (CD4+ and CD8+) and receptors in the same tumor type. These findings indicate that the expression profile of different CXCRs, as well as the recruitment of CXCRs+ tumor cells and CXCRs+ TILs into the tumor microenvironment, is highly regulated. However, detailed molecular mechanism of such regulation is not fully understood. Among CXCR members, CXCR4 is the most intensively studied for its functions in tumor regulation, and the CXCR4 antagonist AMD3100 is the only CXCRtargeted drug approved for clinical use (De Clercq 2019). In contrast, only a few anti-tumor antagonists have been developed for CXCR7, and no antagonists yet exist for CXCR5 or CXCR6 due to the lack of protein structural information, despite numerous studies implicating these receptors in tumorigenic processes (Table 3). It is noteworthy that CXCL16 has both transmembrane and secreted forms with

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

27

opposing functions in tumor progression (Gutwein et al. 2009; Lang et al. 2017). CXCL16 may even act in an “inverse signaling” mechanism, where sCXCL16 directly interacts with TM-CXCL16 to induce downstream signaling without involving CXCR6 activity (Hattermann et al. 2016). Therefore, caution is required when considering the clinical use of antagonists targeting the CXCR6–CXCL16 axis, as this may lead to unpredictable outcomes among tumor types. Similarly, CXCR3A and CXCR3B have both pro- and anti-tumor effects, respectively, so application of the nonselective. CXCR3 antagonist AMG487 may suppress or promote tumor growth depending on context (Puchert et al. 2019). Further investigation may be required when using AMG487 to treat tumors expressing both CXCR3A and CXCR3B. CXCR–CXCL signaling affects a variety of immune cells, leading to a strong rationale to develop translational treatment combining the CXCR targeting with radio-, chemo-, or immunotherapy (Table 4). Some CXCR antagonists have been shown to improve sensitivity of tumor cells to radiotherapy or chemotherapy drugs and reduce tumor recurrence compared with single agent administration (Ning et al. 2012; Reeves et al. 2017; Zhou et al. 2018a; Lecavalier-Barsoum et al. 2018; Fahham et al. 2012; Beider et al. 2013; Bockorny et al. 2020). CXCR antagonist together with immunotherapy also provides novel anticancer strategies by interfering and reshaping the tumor microenvironment. Combination of CXCR2 or CXCR4 antagonist with PD-1–PD-L1 blockade has been shown to promote anti-tumor immunity in various cancer types (Steele et al. 2016; Jiang et al. 2019b; Seo et al. 2019; Zeng et al. 2019), mainly by enhancing T-cell infiltration and removing or inhibiting immunosuppressive cells such as Treg cells and MDSC infiltration, creating an anti-tumor microenvironment. To perform such combination therapies, it is important to understand the function of CXCR–CXCL signaling in immune cells and the effects of CXCR blockade on the immune response to tumor cells. Subtle associations have been discovered among CXCRs. For example, CXCR3 was reported to elevate surface expression of CXCR4 on colorectal cancer cells and enhance CXCR4 activity by forming a heterodimer, thereby promoting cancer invasion (Jin et al. 2018). A similar interaction was reported between CXCR7 and CXCR4 in colorectal cancer progression, with coupling to CXCR7 enhancing the pro-tumor activity of CXCR4 (Song et al. 2019). This implies that application of a CXCR3 antagonist, CXCR7 antagonists, dual antagonist targeting CXCR3 and CXCR4, or a dual antagonist targeting CXCR7 and CXCR4 may be a more effective way to suppress tumorigenesis than a CXCR4 antagonist alone. In addition, CXCR7 is subject to CXCR4 regulation (Wang et al. 2008), so the CXCR4 antagonist AMD3100 also acts on CXCR7 as an allosteric agonist (Kalatskaya et al. 2009) but with less efficiency for suppressing tumor migration (Luo et al. 2018). Furthermore, CXCL12 not only inhibits the expression of CXCR4 by negative feedback but also inhibits the expression of CXCR5 (Chan et al. 2003). While this helps keep the internal environment stable, it also affects the clinical efficacy of anti-tumor antagonists. In the development of CXCR antagonist-based therapies, it is also important to consider that autocrine and paracrine signaling mechanisms in the tumor

28

Y. Xun et al.

Table 4 Combination of CXC receptor antagonist and other drugs for anti-tumor treatment

Targets CXCR1/ CXCR2

Antagonist SCH527123

Combinatory drugs SCH-479833 Oxaliplatin Pembrolizumab

Reparixin

CXCR4

AZD5069 AMD3100

Phase 2

Anti-PD-1 antibody Paclitaxel

PDAC

Both

Ovarian cancer

In vitro

Radiotherapy

TNBC, cervical cancer

Both

Anti-PD-1 antibody Anti-PD-1 antibody

PDAC

In vitro

Ovarian cancer, osteosarcoma

Both

Metastatic pancreatic cancer (NCT04177810) Hand and neck cancer (NCT04058145) Relapsed multiple myeloma (NCT00903968) NSCLC

Phase 2

(Jiang et al. 2019b; Zeng et al. 2019) –

Phase 2



Phase 2



Both

(Fahham et al. 2012)

NHL

Both

Pancreatic cancer (NCT02826486)

Phase 2

(Beider et al. 2013) (Bockorny et al. 2020)

Paclitaxel

Cemiplimab

Pembrolizumab

Bortezomib

BKT140

Both

Ref. (Varney et al. 2011) (Ning et al. 2012) –

NSCLC, CRPC, and colorectal cancer (NCT03473925) Metastatic breast cancer (NCT02370238) Gastric cancer

5-fluorouracil CXCR2

Targeted cancer types Liver metastasis of colon cancer Colon cancer

In vitro/ in vivo, or clinical studies In vitro

Radiotherapy, cisplatin, or paclitaxel Rituximab Pembrolizumab

Phase 2

Both

(Wang et al. 2016; Schott et al. 2017) (Wang et al. 2016) (Steele et al. 2016) (Reeves et al. 2017) (Zhou et al. 2018a; LecavalierBarsoum et al. 2018) (Seo et al. 2019)

(continued)

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

29

Table 4 (continued)

Targets

CXCR7

Antagonist AMD070

Combinatory drugs Nivolumab

Targeted cancer types Renal cell carcinoma (NCT02923531)

Pembrolizumab

Advanced melanoma (NCT02823405) Metastatic breast cancer (NCT01837095) Relapsed glioblastoma multiforme (NCT02765165) Prostate cancer

POL6326

Eribulin

USL-311

Lomustine

CCX771

Enzalutamide

In vitro/ in vivo, or clinical studies Phase 2

Ref. –

Phase 1



Phase 1



Phase 2



Both

(Luo et al. 2018)

microenvironment may give rise to divergent outcomes (Wendel et al. 2008; Chen et al. 2018). Moreover, some CXCR antagonists have strong side effects, such as cardiotoxicity, so development of safer alternatives is required (Gimenez-Arnau et al. 2007). Development of antagonists with milder toxicity and both greater efficiency and receptor specificity is essential. In conclusion, the effects of CXCR signaling on tumor progression warrant continued study to define novel therapeutic targets, including anti-tumor CXCR antagonists. Acknowledgments This study was supported by National Key Research and Development Program (No.2018YFA0902702) and National Science and Technology Major Project (No.2018ZX10731301-003), National Natural Science Foundation of China (No.81570202), Guangdong Medical Research Fund (No.A2019396), basic and applied basic research project of Guangdong Province (No.2019A1515110495), and Research Start-up Foundation for Lingnan Scholar of Foshan University (directed by Fang Liu). Conflict of Interest The authors declare no conflicts of interest regarding this report.

References Al-Toub M, Almohawes M, Vishnubalaji R, Alfayez M, Aldahmash A, Kassem M et al (2019) CXCR7 signaling promotes breast cancer survival in response to mesenchymal stromal stem cell-derived factors. Cell Death Discov 5:87 Bachelerie F, Ben-Baruch A, Burkhardt AM, Combadiere C, Farber JM, Graham GJ et al (2014) International Union of Basic and Clinical Pharmacology. [corrected]. LXXXIX. Update on the extended family of chemokine receptors and introducing a new nomenclature for atypical chemokine receptors. Pharmacol Rev 66(1):1–79

30

Y. Xun et al.

Bai M, Zheng Y, Liu H, Su B, Zhan Y, He H (2017) CXCR5(+) CD8(+) T cells potently infiltrate pancreatic tumors and present high functionality. Exp Cell Res 361(1):39–45 Baugher PJ, Richmond A (2008) The carboxyl-terminal PDZ ligand motif of chemokine receptor CXCR2 modulates post-endocytic sorting and cellular chemotaxis. J Biol Chem 283 (45):30868–30878 Becker JH, Gao Y, Soucheray M, Pulido I, Kikuchi E, Rodriguez ML et al (2019) CXCR7 reactivates ERK signaling to promote resistance to EGFR kinase inhibitors in NSCLC. Cancer Res 79(17):4439–4452 Beider K, Begin M, Abraham M, Wald H, Weiss ID, Wald O et al (2011) CXCR4 antagonist 4F-benzoyl-TN14003 inhibits leukemia and multiple myeloma tumor growth. Exp Hematol 39 (3):282–292 Beider K, Ribakovsky E, Abraham M, Wald H, Weiss L, Rosenberg E et al (2013) Targeting the CD20 and CXCR4 pathways in non-hodgkin lymphoma with rituximab and high-affinity CXCR4 antagonist BKT140. Clin Cancer Res 19(13):3495–3507 Beider K, Darash-Yahana M, Blaier O, Koren-Michowitz M, Abraham M, Wald H et al (2014) Combination of imatinib with CXCR4 antagonist BKT140 overcomes the protective effect of stroma and targets CML in vitro and in vivo. Mol Cancer Ther 13(5):1155–1169 Benhadjeba S, Edjekouane L, Sauve K, Carmona E, Tremblay A (2018) Feedback control of the CXCR7/CXCL11 chemokine axis by estrogen receptor alpha in ovarian cancer. Mol Oncol 12 (10):1689–1705 Bi H, Zhang Y, Wang S, Fang W, He W, Yin L et al (2019) Interleukin-8 promotes cell migration via CXCR1 and CXCR2 in liver cancer. Oncol Lett 18(4):4176–4184 Bleul CC, Fuhlbrigge RC, Casasnovas JM, Aiuti A, Springer TA (1996) A highly efficacious lymphocyte chemoattractant, stromal cell-derived factor 1 (SDF-1). J Exp Med 184 (3):1101–1109 Bockorny B, Semenisty V, Macarulla T, Borazanci E, Wolpin BM, Stemmer SM et al (2020) BL-8040, a CXCR4 antagonist, in combination with pembrolizumab and chemotherapy for pancreatic cancer: the COMBAT trial. Nat Med 26(6):878–885 Bronger H, Karge A, Dreyer T, Zech D, Kraeft S, Avril S et al (2017) Induction of cathepsin B by the CXCR3 chemokines CXCL9 and CXCL10 in human breast cancer cells. Oncol Lett 13 (6):4224–4230 Burger JA, Stewart DJ, Wald O, Peled A (2011) Potential of CXCR4 antagonists for the treatment of metastatic lung cancer. Expert Rev Anticancer Ther 11(4):621–630 Cambien B, Karimdjee BF, Richard-Fiardo P, Bziouech H, Barthel R, Millet MA et al (2009) Organ-specific inhibition of metastatic colon carcinoma by CXCR3 antagonism. Br J Cancer 100(11):1755–1764 Cha Z, Qian G, Zang Y, Gu H, Huang Y, Zhu L et al (2017) Circulating CXCR5+CD4+ T cells assist in the survival and growth of primary diffuse large B cell lymphoma cells through interleukin 10 pathway. Exp Cell Res 350(1):154–160 Chan CC, Shen D, Hackett JJ, Buggage RR, Tuaillon N (2003) Expression of chemokine receptors, CXCR4 and CXCR5, and chemokines, BLC and SDF-1, in the eyes of patients with primary intraocular lymphoma. Ophthalmology 110(2):421–426 Chang Y, Zhou L, Xu L, Fu Q, Yang Y, Lin Z et al (2017) High expression of CXC chemokine receptor 6 associates with poor prognosis in patients with clear cell renal cell carcinoma. Urol Oncol 35(12):675 e17–675 e24 Chatterjee S, Behnam Azad B, Nimmagadda S (2014) The intricate role of CXCR4 in cancer. Adv Cancer Res 124:31–82 Chen Y, Teng F, Wang G, Nie Z (2016) Overexpression of CXCR7 induces angiogenic capacity of human hepatocellular carcinoma cells via the AKT signaling pathway. Oncol Rep 36 (4):2275–2281 Chen F, Yin S, Niu L, Luo J, Wang B, Xu Z et al (2018) Expression of the chemokine receptor CXCR3 correlates with dendritic cell recruitment and prognosis in gastric cancer. Genet Test Mol Biomarkers 22(1):35–42

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

31

Chen F, Yuan J, Yan H, Liu H, Yin S (2019a) Chemokine receptor CXCR3 correlates with decreased M2 macrophage infiltration and favorable prognosis in gastric cancer. Biomed Res Int 2019:6832867 Chen L, Ouyang J, Wienand K, Bojarczuk K, Hao Y, Chapuy B et al (2019b) CXCR4 upregulation is an indicator of sensitivity to B-cell receptor/PI3K blockade and a potential resistance mechanism in B-cell receptor-dependent diffuse large B-cell lymphomas. Haematologica 105 (5):1361–1368 Chen IX, Chauhan VP, Posada J, Ng MR, Wu MW, Adstamongkonkul P et al (2019c) Blocking CXCR4 alleviates desmoplasia, increases T-lymphocyte infiltration, and improves immunotherapy in metastatic breast cancer. Proc Natl Acad Sci U S A 116(10):4558–4566 Chheda ZS, Sharma RK, Jala VR, Luster AD, Haribabu B (2016) Chemoattractant receptors BLT1 and CXCR3 regulate antitumor immunity by facilitating CD8+ T cell migration into tumors. J Immunol 197(5):2016–2026 Chu F, Li HS, Liu X, Cao J, Ma W, Ma Y et al (2019) CXCR5(+)CD8(+) T cells are a distinct functional subset with an antitumor activity. Leukemia 33(11):2640–2653 Chung B, Esmaeili AA, Gopalakrishna-Pillai S, Murad JP, Andersen ES, Kumar Reddy N et al (2017) Human brain metastatic stroma attracts breast cancer cells via chemokines CXCL16 and CXCL12. NPJ Breast Cancer 3:6 Ciuca FI, Marasescu PC, Matei M, Florescu AM, Margaritescu C, Demetrian AD et al (2019) The prognostic value of CXCR4, MMP-2 and MMP-9 in tongue squamous carcinoma. Romanian J Morphol Embryol 60(1):59–68 Colvin RA, Campanella GS, Sun J, Luster AD (2004) Intracellular domains of CXCR3 that mediate CXCL9, CXCL10, and CXCL11 function. J Biol Chem 279(29):30219–30227 De Clercq E (2019) Mozobil(R) (Plerixafor, AMD3100), 10 years after its approval by the US Food and Drug Administration. Antivir Chem Chemother 27:2040206619829382 Desurmont T, Skrypek N, Duhamel A, Jonckheere N, Millet G, Leteurtre E et al (2015) Overexpression of chemokine receptor CXCR2 and ligand CXCL7 in liver metastases from colon cancer is correlated to shorter disease-free and overall survival. Cancer Sci 106 (3):262–269 Di Mitri D, Mirenda M, Vasilevska J, Calcinotto A, Delaleu N, Revandkar A et al (2019) Re-education of tumor-associated macrophages by CXCR2 blockade drives senescence and tumor inhibition in advanced prostate cancer. Cell Rep 28(8):2156–68 e5 Ding Y, Shen J, Zhang G, Chen X, Wu J, Chen W (2015) CD40 controls CXCR5-induced recruitment of myeloid-derived suppressor cells to gastric cancer. Oncotarget 6 (36):38901–38911 Donzella GA, Schols D, Lin SW, Este JA, Nagashima KA, Maddon PJ et al (1998) AMD3100, a small molecule inhibitor of HIV-1 entry via the CXCR4 co-receptor. Nat Med 4(1):72–77 Du J, Zhang X, Han J, Man K, Zhang Y, Chu ES et al (2017) Pro-inflammatory CXCR3 impairs mitochondrial function in experimental non-alcoholic steatohepatitis. Theranostics 7 (17):4192–4203 Du W, Lu C, Zhu X, Hu D, Chen X, Li J et al (2019) Prognostic significance of CXCR4 expression in acute myeloid leukemia. Cancer Med 8(15):6595–6603 Duan Z, Gao J, Zhang L, Liang H, Huang X, Xu Q et al (2015) Phenotype and function of CXCR5 +CD45RA-CD4+ T cells were altered in HBV-related hepatocellular carcinoma and elevated serum CXCL13 predicted better prognosis. Oncotarget 6(42):44239–44253 Eikawa S, Ohue Y, Kitaoka K, Aji T, Uenaka A, Oka M et al (2010) Enrichment of Foxp3+ CD4 regulatory T cells in migrated T cells to IL-6- and IL-8-expressing tumors through predominant induction of CXCR1 by IL-6. J Immunol 185(11):6734–6740 El Haibi CP, Sharma PK, Singh R, Johnson PR, Suttles J, Singh S et al (2010) PI3Kp110-, Src-, FAK-dependent and DOCK2-independent migration and invasion of CXCL13-stimulated prostate cancer cells. Mol Cancer 9:85

32

Y. Xun et al.

Erin N, Nizam E, Tanriover G, Koksoy S (2015) Autocrine control of MIP-2 secretion from metastatic breast cancer cells is mediated by CXCR2: a mechanism for possible resistance to CXCR2 antagonists. Breast Cancer Res Treat 150(1):57–69 Erreni M, Laghi L, Celesti G, Bianchi P, Grizzi F (2010) W1767 expression and function of the chemokine CX3CL1 and its receptor Cx3cr1 in human colorectal cancer. Gastroenterology 138 (5):S-736-S Fahham D, Weiss ID, Abraham M, Beider K, Hanna W, Shlomai Z et al (2012) In vitro and in vivo therapeutic efficacy of CXCR4 antagonist BKT140 against human non-small cell lung cancer. J Thorac Cardiovasc Surg 144(5):1167–75 e1 Furusato B, Mohamed A, Uhlen M, Rhim JS (2010) CXCR4 and cancer. Pathol Int 60(7):497–505 Gabellini C, Trisciuoglio D, Desideri M, Candiloro A, Ragazzoni Y, Orlandi A et al (2009) Functional activity of CXCL8 receptors, CXCR1 and CXCR2, on human malignant melanoma progression. Eur J Cancer 45(14):2618–2627 Gadalla R, Hassan H, Ibrahim SA, Abdullah MS, Gaballah A, Greve B et al (2019) Tumor microenvironmental plasmacytoid dendritic cells contribute to breast cancer lymph node metastasis via CXCR4/SDF-1 axis. Breast Cancer Res Treat 174(3):679–691 Gao B, Lin J, Jiang Z, Yang Z, Yu H, Ding L et al (2018) Upregulation of chemokine CXCL10 enhances chronic pulmonary inflammation in tree shrew collagen-induced arthritis. Sci Rep 8 (1):9993 Garg R, Blando JM, Perez CJ, Abba MC, Benavides F, Kazanietz MG (2017) Protein kinase C epsilon cooperates with PTEN loss for prostate tumorigenesis through the CXCL13-CXCR5 pathway. Cell Rep 19(2):375–388 Gimenez-Arnau A, Pujol RM, Ianosi S, Kaszuba A, Malbran A, Poop G et al (2007) Rupatadine in the treatment of chronic idiopathic urticaria: a double-blind, randomized, placebo-controlled multicentre study. Allergy 62(5):539–546 Goldberg-Bittman L, Neumark E, Sagi-Assif O, Azenshtein E, Meshel T, Witz IP et al (2004) The expression of the chemokine receptor CXCR3 and its ligand, CXCL10, in human breast adenocarcinoma cell lines. Immunol Lett 92(1–2):171–178 Grepin R, Guyot M, Giuliano S, Boncompagni M, Ambrosetti D, Chamorey E et al (2014) The CXCL7/CXCR1/2 axis is a key driver in the growth of clear cell renal cell carcinoma. Cancer Res 74(3):873–883 Gutwein P, Schramme A, Sinke N, Abdel-Bakky MS, Voss B, Obermuller N et al (2009) Tumoural CXCL16 expression is a novel prognostic marker of longer survival times in renal cell cancer patients. Eur J Cancer 45(3):478–489 Han XG, Du L, Qiao H, Tu B, Wang YG, Qin A et al (2015) CXCR1 knockdown improves the sensitivity of osteosarcoma to cisplatin. Cancer Lett 369(2):405–415 Hao M, Zheng J, Hou K, Wang J, Chen X, Lu X et al (2012) Role of chemokine receptor CXCR7 in bladder cancer progression. Biochem Pharmacol 84(2):204–214 Hattermann K, Bartsch K, Gebhardt HH, Mehdorn HM, Synowitz M, Schmitt AD et al (2016) “Inverse signaling” of the transmembrane chemokine CXCL16 contributes to proliferative and anti-apoptotic effects in cultured human meningioma cells. Cell Commun Signal 14(1):26 Hojo S, Koizumi K, Tsuneyama K, Arita Y, Cui Z, Shinohara K et al (2007) High-level expression of chemokine CXCL16 by tumor cells correlates with a good prognosis and increased tumorinfiltrating lymphocytes in colorectal cancer. Cancer Res 67(10):4725–4731 Hong L, Wang S, Li W, Wu D, Chen W (2018) Tumor-associated macrophages promote the metastasis of ovarian carcinoma cells by enhancing CXCL16/CXCR6 expression. Pathol Res Pract 214(9):1345–1351 Ijichi H, Chytil A, Gorska AE, Aakre ME, Bierie B, Tada M et al (2011) Inhibiting Cxcr2 disrupts tumor-stromal interactions and improves survival in a mouse model of pancreatic ductal adenocarcinoma. J Clin Invest 121(10):4106–4117 Jiang Q, Sun Y, Liu X (2019a) CXCR4 as a prognostic biomarker in gastrointestinal cancer: a metaanalysis. Biomarkers 24(6):510–516

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

33

Jiang K, Li J, Zhang J, Wang L, Zhang Q, Ge J et al (2019b) SDF-1/CXCR4 axis facilitates myeloid-derived suppressor cells accumulation in osteosarcoma microenvironment and blunts the response to anti-PD-1 therapy. Int Immunopharmacol 75:105818 Jifu E, Yan F, Kang Z, Zhu L, Xing J, Yu E (2018) CD8(+)CXCR5(+) T cells in tumor-draining lymph nodes are highly activated and predict better prognosis in colorectal cancer. Hum Immunol 79(6):446–452 Jin Y, Lang C, Tang J, Geng J, Song HK, Sun Z et al (2017a) CXCR5(+)CD8(+) T cells could induce the death of tumor cells in HBV-related hepatocellular carcinoma. Int Immunopharmacol 53:42–48 Jin JJ, Dai FX, Long ZW, Cai H, Liu XW, Zhou Y et al (2017b) CXCR6 predicts poor prognosis in gastric cancer and promotes tumor metastasis through epithelial-mesenchymal transition. Oncol Rep 37(6):3279–3286 Jin J, Zhang Z, Wang H, Zhan Y, Li G, Yang H et al (2018) CXCR3 expression in colorectal cancer cells enhanced invasion through preventing CXCR4 internalization. Exp Cell Res 371 (1):162–174 Kaifi JT, Yekebas EF, Schurr P, Obonyo D, Wachowiak R, Busch P et al (2005) Tumor-cell homing to lymph nodes and bone marrow and CXCR4 expression in esophageal cancer. J Natl Cancer Inst 97(24):1840–1847 Kalatskaya I, Berchiche YA, Gravel S, Limberg BJ, Rosenbaum JS, Heveker N (2009) AMD3100 is a CXCR7 ligand with allosteric agonist properties. Mol Pharmacol 75(5):1240 Kapur N, Mir H, Sonpavde GP, Jain S, Bae S, Lillard JW Jr et al (2019) Prostate cancer cells hyperactivate CXCR6 signaling by cleaving CXCL16 to overcome effect of docetaxel. Cancer Lett 454:1–13 Kashyap MK, Kumar D, Jones H, Amaya-Chanaga CI, Choi MY, Melo-Cardenas J et al (2016) Ulocuplumab (BMS-936564/MDX1338): a fully human anti-CXCR4 antibody induces cell death in chronic lymphocytic leukemia mediated through a reactive oxygen species-dependent pathway. Oncotarget 7(3):2809–2822 Kaunisto A, Henry WS, Montaser-Kouhsari L, Jaminet SC, Oh EY, Zhao L et al (2015) NFAT1 promotes intratumoral neutrophil infiltration by regulating IL8 expression in breast cancer. Mol Oncol 9(6):1140–1154 Kawada K, Sonoshita M, Sakashita H, Takabayashi A, Yamaoka Y, Manabe T et al (2004) Pivotal role of CXCR3 in melanoma cell metastasis to lymph nodes. Cancer Res 64(11):4010–4017 Kawada K, Hosogi H, Sonoshita M, Sakashita H, Manabe T, Shimahara Y et al (2007) Chemokine receptor CXCR3 promotes colon cancer metastasis to lymph nodes. Oncogene 26 (32):4679–4688 Ke C, Ren Y, Lv L, Hu W, Zhou W (2017) Association between CXCL16/CXCR6 expression and the clinicopathological features of patients with non-small cell lung cancer. Oncol Lett 13 (6):4661–4668 Keane MP, Belperio JA, Xue YY, Burdick MD, Strieter RM (2004) Depletion of CXCR2 inhibits tumor growth and angiogenesis in a murine model of lung cancer. J Immunol 172(5):2853–2860 Kenny PA, Lee GY, Bissell MJ (2007) Targeting the tumor microenvironment. Front Biosci 12:3468–3474 Kuhne MR, Mulvey T, Belanger B, Chen S, Pan C, Chong C et al (2013) BMS-936564/MDX1338: a fully human anti-CXCR4 antibody induces apoptosis in vitro and shows antitumor activity in vivo in hematologic malignancies. Clin Cancer Res 19(2):357–366 Kumar S, Wilkes DW, Samuel N, Blanco MA, Nayak A, Alicea-Torres K et al (2018) DeltaNp63driven recruitment of myeloid-derived suppressor cells promotes metastasis in triple-negative breast cancer. J Clin Invest 128(11):5095–5109 Lang K, Bonberg N, Robens S, Behrens T, Hovanec J, Deix T et al (2017) Soluble chemokine (C-X-C motif) ligand 16 (CXCL16) in urine as a novel biomarker candidate to identify high grade and muscle invasive urothelial carcinomas. Oncotarget 8(62):104946–104959 Lasagni L, Francalanci M, Annunziato F, Lazzeri E, Giannini S, Cosmi L et al (2003) An alternatively spliced variant of CXCR3 mediates the inhibition of endothelial cell growth

34

Y. Xun et al.

induced by IP-10, Mig, and I-TAC, and acts as functional receptor for platelet factor 4. J Exp Med 197(11):1537–1549 Lecavalier-Barsoum M, Chaudary N, Han K, Koritzinsky M, Hill R, Milosevic M (2018) Targeting the CXCL12/CXCR4 pathway and myeloid cells to improve radiation treatment of locally advanced cervical cancer. Int J Cancer 143(5):1017–1028 Lee YS, Choi I, Ning Y, Kim NY, Khatchadourian V, Yang D et al (2012) Interleukin-8 and its receptor CXCR2 in the tumour microenvironment promote colon cancer growth, progression and metastasis. Br J Cancer 106(11):1833–1841 Lepore F, D’Alessandro G, Antonangeli F, Santoro A, Esposito V, Limatola C et al (2018) CXCL16/CXCR6 axis drives microglia/macrophages phenotype in physiological conditions and plays a crucial role in glioma. Front Immunol 9:2750 Li L, Xu L, Yan J, Zhen ZJ, Ji Y, Liu CQ et al (2015) CXCR2-CXCL1 axis is correlated with neutrophil infiltration and predicts a poor prognosis in hepatocellular carcinoma. J Exp Clin Cancer Res 34:129 Li H, Chen Y, Xu N, Yu M, Tu X, Chen Z et al (2017a) AMD3100 inhibits brain-specific metastasis in lung cancer via suppressing the SDF-1/CXCR4 axis and protecting blood-brain barrier. Am J Transl Res 9(12):5259–5274 Li W, Ding Q, Ding Y, Lu L, Wang X, Zhang Y et al (2017b) Oroxylin A reverses the drug resistance of chronic myelogenous leukemia cells to imatinib through CXCL12/CXCR7 axis in bone marrow microenvironment. Mol Carcinog 56(3):863–876 Li Z, Chen G, Ding L, Wang Y, Zhu C, Wang K et al (2019a) Increased survival by pulmonary treatment of established lung metastases with dual STAT3/CXCR4 inhibition by siRNA nanoemulsions. Mol Ther 27(12):2100–2110 Li X, Wang X, Li Z, Zhang Z, Zhang Y (2019b) Chemokine receptor 7 targets the vascular endothelial growth factor via the AKT/ERK pathway to regulate angiogenesis in colon cancer. Cancer Med 8(11):5327–5340 Li N, Xu H, Ou Y, Feng Z, Zhang Q, Zhu Q et al (2019c) LPS-induced CXCR7 expression promotes gastric Cancer proliferation and migration via the TLR4/MD-2 pathway. Diagn Pathol 14(1):3 Li S, Fong KW, Gritsina G, Zhang A, Zhao JC, Kim J et al (2019d) Activation of MAPK signaling by CXCR7 leads to enzalutamide resistance in prostate cancer. Cancer Res 79(10):2580–2592 Liao YX, Fu ZZ, Zhou CH, Shan LC, Wang ZY, Yin F et al (2015) AMD3100 reduces CXCR4mediated survival and metastasis of osteosarcoma by inhibiting JNK and Akt, but not p38 or Erk1/2, pathways in in vitro and mouse experiments. Oncol Rep 34(1):33–42 Lin L, Han MM, Wang F, Xu LL, Yu HX, Yang PY (2014) CXCR7 stimulates MAPK signaling to regulate hepatocellular carcinoma progression. Cell Death Dis 5:e1488 Liu L, Sun H, Wu S, Tan H, Sun Y, Liu X et al (2019a) IL17A promotes CXCR2dependent angiogenesis in a mouse model of liver cancer. Mol Med Rep 20(2):1065–1074 Liu B, Song S, Setroikromo R, Chen S, Hu W, Chen D et al (2019b) CX chemokine receptor 7 contributes to survival of KRAS-mutant non-small cell lung cancer upon loss of epidermal growth factor receptor. Cancers 11(4):455 Lo MC, Yip TC, Ngan KC, Cheng WW, Law CK, Chan PS et al (2013) Role of MIF/CXCL8/ CXCR2 signaling in the growth of nasopharyngeal carcinoma tumor spheres. Cancer Lett 335 (1):81–92 Lofroos AB, Kadivar M, Resic Lindehammer S, Marsal J (2017) Colorectal cancer-infiltrating T lymphocytes display a distinct chemokine receptor expression profile. Eur J Med Res 22(1):40 Lu G, Wu Y, Jiang Y, Wang S, Hou Y, Guan X et al (2013) Structural insights into neutrophilic migration revealed by the crystal structure of the chemokine receptor CXCR2 in complex with the first PDZ domain of NHERF1. PLoS One 8(10):e76219 Lu CL, Guo J, Gu J, Ge D, Hou YY, Lin ZW et al (2014) CXCR4 heterogeneous expression in esophageal squamous cell cancer and stronger metastatic potential with CXCR4-positive cancer cells. Dis Esophagus 27(3):294–302

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

35

Luo Y, Azad AK, Karanika S, Basourakos SP, Zuo X, Wang J et al (2018) Enzalutamide and CXCR7 inhibitor combination treatment suppresses cell growth and angiogenic signaling in castration-resistant prostate cancer models. Int J Cancer 142(10):2163–2174 Luo N, Chen DD, Liu L, Li L, Cheng ZP (2019) CXCL12 promotes human ovarian cancer cell invasion through suppressing ARHGAP10 expression. Biochem Biophys Res Commun 518 (3):416–422 Ma X, Norsworthy K, Kundu N, Rodgers WH, Gimotty PA, Goloubeva O et al (2009) CXCR3 expression is associated with poor survival in breast cancer and promotes metastasis in a murine model. Mol Cancer Ther 8(3):490–498 Ma Y, Xu X, Luo M (2017) CXCR6 promotes tumor cell proliferation and metastasis in osteosarcoma through the Akt pathway. Cell Immunol 311:80–85 MacManus CF, Pettigrew J, Seaton A, Wilson C, Maxwell PJ, Berlingeri S et al (2007) Interleukin8 signaling promotes translational regulation of cyclin D in androgen-independent prostate cancer cells. Mol Cancer Res 5(7):737–748 Maekawa S, Iwasaki A, Shirakusa T, Kawakami T, Yanagisawa J (2008) Association between the expression of chemokine receptors CCR7 and CXCR3, and lymph node metastatic potential in lung adenocarcinoma. Oncol Rep 19(6):1461–1468 Manjavachi MN, Quintao NL, Campos MM, Deschamps IK, Yunes RA, Nunes RJ et al (2010) The effects of the selective and non-peptide CXCR2 receptor antagonist SB225002 on acute and long-lasting models of nociception in mice. Eur J Pain 14(1):23–31 Matsuo Y, Campbell PM, Brekken RA, Sung B, Ouellette MM, Fleming JB et al (2009) K-Ras promotes angiogenesis mediated by immortalized human pancreatic epithelial cells through mitogen-activated protein kinase signaling pathways. Mol Cancer Res 7(6):799–808 Meijer J, Zeelenberg IS, Sipos B, Roos E (2006) The CXCR5 chemokine receptor is expressed by carcinoma cells and promotes growth of colon carcinoma in the liver. Cancer Res 66 (19):9576–9582 Mir H, Kaur G, Kapur N, Bae S, Lillard JW Jr, Singh S (2019) Higher CXCL16 exodomain is associated with aggressive ovarian cancer and promotes the disease by CXCR6 activation and MMP modulation. Sci Rep 9(1):2527 Morimoto M, Matsuo Y, Koide S, Tsuboi K, Shamoto T, Sato T et al (2016) Enhancement of the CXCL12/CXCR4 axis due to acquisition of gemcitabine resistance in pancreatic cancer: effect of CXCR4 antagonists. BMC Cancer 16:305 Mossanen JC, Kohlhepp M, Wehr A, Krenkel O, Liepelt A, Roeth AA et al (2019) CXCR6 inhibits hepatocarcinogenesis by promoting natural killer T- and CD4(+) T-cell-dependent control of senescence. Gastroenterology 156(6):1877–89 e4 Muller A, Homey B, Soto H, Ge N, Catron D, Buchanan ME et al (2001) Involvement of chemokine receptors in breast cancer metastasis. Nature 410(6824):50–56 Muller N, Michen S, Tietze S, Topfer K, Schulte A, Lamszus K et al (2015) Engineering NK cells modified with an EGFRvIII-specific chimeric antigen receptor to overexpress CXCR4 improves immunotherapy of CXCL12/SDF-1alpha-secreting glioblastoma. J Immunother 38(5):197–210 Najjar YG, Rayman P, Jia X, Pavicic PG Jr, Rini BI, Tannenbaum C et al (2017) Myeloid-derived suppressor cell subset accumulation in renal cell carcinoma parenchyma is associated with intratumoral expression of IL1beta, IL8, CXCL5, and Mip-1alpha. Clin Cancer Res 23 (9):2346–2355 Nakajima C, Mukai T, Yamaguchi N, Morimoto Y, Park WR, Iwasaki M et al (2002) Induction of the chemokine receptor CXCR3 on TCR-stimulated T cells: dependence on the release from persistent TCR-triggering and requirement for IFN-gamma stimulation. Eur J Immunol 32 (6):1792–1801 Naumann U, Cameroni E, Pruenster M, Mahabaleshwar H, Raz E, Zerwes HG et al (2010) CXCR7 functions as a scavenger for CXCL12 and CXCL11. PLoS One 5(2):e9175 Ning Y, Labonte MJ, Zhang W, Bohanes PO, Gerger A, Yang D et al (2012) The CXCR2 antagonist, SCH-527123, shows antitumor activity and sensitizes cells to oxaliplatin in preclinical colon cancer models. Mol Cancer Ther 11(6):1353–1364

36

Y. Xun et al.

Nywening TM, Belt BA, Cullinan DR, Panni RZ, Han BJ, Sanford DE et al (2018) Targeting both tumour-associated CXCR2(+) neutrophils and CCR2(+) macrophages disrupts myeloid recruitment and improves chemotherapeutic responses in pancreatic ductal adenocarcinoma. Gut 67 (6):1112–1123 Ogawa R, Yamamoto T, Hirai H, Hanada K, Kiyasu Y, Nishikawa G et al (2019) Loss of SMAD4 promotes colorectal cancer progression by recruiting tumor-associated neutrophils via the CXCL1/8-CXCR2 axis. Clin Cancer Res 25(9):2887–2899 Oghumu S, Varikuti S, Terrazas C, Kotov D, Nasser MW, Powell CA et al (2014) CXCR3 deficiency enhances tumor progression by promoting macrophage M2 polarization in a murine breast cancer model. Immunology 143(1):109–119 Oja AE, Piet B, van der Zwan D, Blaauwgeers H, Mensink M, de Kivit S et al (2018) Functional heterogeneity of CD4(+) tumor-infiltrating lymphocytes with a resident memory phenotype in NSCLC. Front Immunol 9:2654 Oliveira Frick V, Rubie C, Ghadjar P, Faust SK, Wagner M, Graber S et al (2011) Changes in CXCL12/CXCR4-chemokine expression during onset of colorectal malignancies. Tumour Biol 32(1):189–196 Palladino P, Portella L, Colonna G, Raucci R, Saviano G, Rossi F et al (2012) The N-terminal region of CXCL11 as structural template for CXCR3 molecular recognition: synthesis, conformational analysis, and binding studies. Chem Biol Drug Des 80(2):254–265 Panjideh H, Muller G, Koch M, Wilde F, Scheu S, Moldenhauer G et al (2014) Immunotherapy of B-cell non-Hodgkin lymphoma by targeting the chemokine receptor CXCR5 in a preclinical mouse model. Int J Cancer 135(11):2623–2632 Parameswaran R, Yu M, Lim M, Groffen J, Heisterkamp N (2011) Combination of drug therapy in acute lymphoblastic leukemia with a CXCR4 antagonist. Leukemia 25(8):1314–1323 Park SH, Das BB, Casagrande F, Tian Y, Nothnagel HJ, Chu M et al (2012) Structure of the chemokine receptor CXCR1 in phospholipid bilayers. Nature 491(7426):779–783 Peng ZH, Kopecek J (2014) HPMA copolymer CXCR4 antagonist conjugates substantially inhibited the migration of prostate cancer cells. ACS Macro Lett 3(12):1240–1243 Pradelli E, Karimdjee-Soilihi B, Michiels JF, Ricci JE, Millet MA, Vandenbos F et al (2009) Antagonism of chemokine receptor CXCR3 inhibits osteosarcoma metastasis to lungs. Int J Cancer 125(11):2586–2594 Proost P, Schutyser E, Menten P, Struyf S, Wuyts A, Opdenakker G et al (2001) Amino-terminal truncation of CXCR3 agonists impairs receptor signaling and lymphocyte chemotaxis, while preserving antiangiogenic properties. Blood 98(13):3554–3561 Puchert M, Obst J, Koch C, Zieger K, Engele J (2019) CXCL11 promotes tumor progression by the biased use of the chemokine receptors CXCR3 and CXCR7. Cytokine 125:154809 Purohit A, Varney M, Rachagani S, Ouellette MM, Batra SK, Singh RK (2016) CXCR2 signaling regulates KRAS(G(1)(2)D)-induced autocrine growth of pancreatic cancer. Oncotarget 7 (6):7280–7296 Qi XW, Xia SH, Yin Y, Jin LF, Pu Y, Hua D et al (2014) Expression features of CXCR5 and its ligand, CXCL13 associated with poor prognosis of advanced colorectal cancer. Eur Rev Med Pharmacol Sci 18(13):1916–1924 Qian T, Liu Y, Dong Y, Zhang L, Dong Y, Sun Y et al (2018) CXCR7 regulates breast tumor metastasis and angiogenesis in vivo and in vitro. Mol Med Rep 17(3):3633–3639 Qiao Y, Zhang C, Li A, Wang D, Luo Z, Ping Y et al (2018) IL6 derived from cancer-associated fibroblasts promotes chemoresistance via CXCR7 in esophageal squamous cell carcinoma. Oncogene 37(7):873–883 Rafiei S, Gui B, Wu J, Liu XS, Kibel AS, Jia L (2019) Targeting the MIF/CXCR7/AKT signaling pathway in castration-resistant prostate cancer. Mol Cancer Res 17(1):263–276 Randhawa S, Cho BS, Ghosh D, Sivina M, Koehrer S, Muschen M et al (2016) Effects of pharmacological and genetic disruption of CXCR4 chemokine receptor function in B-cell acute lymphoblastic leukaemia. Br J Haematol 174(3):425–436

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

37

Reeves PM, Abbaslou MA, Kools FRW, Poznansky MC (2017) CXCR4 blockade with AMD3100 enhances Taxol chemotherapy to limit ovarian cancer cell growth. Anticancer Drugs 28 (9):935–942 Rezakhaniha B, Dormanesh B, Pirasteh H, Yahaghi E, Masoumi B, Ziari K et al (2016) Immunohistochemical distinction of metastases of renal cell carcinoma with molecular analysis of overexpression of the chemokines CXCR2 and CXCR3 as independent positive prognostic factors for the tumorigenesis. IUBMB Life 68(8):629–633 Romero-Moreno R, Curtis KJ, Coughlin TR, Cristina Miranda-Vergara M, Dutta S, Natarajan A et al (2019) The CXCL5/CXCR2 axis is sufficient to promote breast cancer colonization during bone metastasis. Nat Commun 10(1):4404 Rong G, Kang H, Wang Y, Hai T, Sun H (2013) Candidate markers that associate with chemotherapy resistance in breast cancer through the study on Taxotere-induced damage to tumor microenvironment and gene expression profiling of carcinoma-associated fibroblasts (CAFs). PLoS One 8(8):e70960 Saahene RO, Wang J, Wang ML, Agbo E, Song H (2019) The role of CXC chemokine ligand 4/CXC chemokine receptor 3-B in breast cancer progression. Biotech Histochem 94(1):53–59 Saintigny P, Massarelli E, Lin S, Ahn YH, Chen Y, Goswami S et al (2013) CXCR2 expression in tumor cells is a poor prognostic factor and promotes invasion and metastasis in lung adenocarcinoma. Cancer Res 73(2):571–582 Salazar N, Munoz D, Kallifatidis G, Singh RK, Jorda M, Lokeshwar BL (2014) The chemokine receptor CXCR7 interacts with EGFR to promote breast cancer cell proliferation. Mol Cancer 13:198 Saxena R, Wang Y, Mier JW (2019) CXCR4 inhibition modulates the tumor microenvironment and retards the growth of B16-OVA melanoma and Renca tumors. Melanoma Res 30(1):14–25 Scala S, Ottaiano A, Ascierto PA, Cavalli M, Simeone E, Giuliano P et al (2005) Expression of CXCR4 predicts poor prognosis in patients with malignant melanoma. Clin Cancer Res 11 (5):1835–1841 Schiffer L, Worthmann L, Haller H, Schiffer M (2015) CXCL13 as a new biomarker of systemic lupus erythematosus and lupus nephritis – from bench to bedside? Clin Exp Immunol 179(1):85 Schinke C, Giricz O, Li W, Shastri A, Gordon S, Barreyro L et al (2015) IL8-CXCR2 pathway inhibition as a therapeutic strategy against MDS and AML stem cells. Blood 125 (20):3144–3152 Schott AF, Goldstein LJ, Cristofanilli M, Ruffini PA, McCanna S, Reuben JM et al (2017) Phase Ib pilot study to evaluate reparixin in combination with weekly paclitaxel in patients with HER-2negative metastatic breast cancer. Clin Cancer Res 23(18):5358–5365 Seo YD, Jiang X, Sullivan KM, Jalikis FG, Smythe KS, Abbasi A et al (2019) Mobilization of CD8 (+) T cells via CXCR4 blockade facilitates PD-1 checkpoint therapy in human pancreatic cancer. Clin Cancer Res 25(13):3934–3945 Shah D, Osipo C (2016) Cancer stem cells and HER2 positive breast cancer: the story so far. Genes Dis 3(2):114–123 Shang FM, Li J (2018) A small-molecule antagonist of CXCR1 and CXCR2 inhibits cell proliferation, migration and invasion in melanoma via PI3K/AKT pathway. Med Clin 152(11):425–430 Shi A, Shi H, Dong L, Xu S, Jia M, Guo X et al (2017) CXCR7 as a chemokine receptor for SDF-1 promotes gastric cancer progression via MAPK pathways. Scand J Gastroenterol 52 (6–7):745–753 Shin SY, Nam JS, Lim Y, Lee YH (2010) TNFalpha-exposed bone marrow-derived mesenchymal stem cells promote locomotion of MDA-MB-231 breast cancer cells through transcriptional activation of CXCR3 ligand chemokines. J Biol Chem 285(40):30731–30740 Shrivastava MS, Hussain Z, Giricz O, Shenoy N, Polineni R, Maitra A et al (2014) Targeting chemokine pathways in esophageal adenocarcinoma. Cell Cycle 13(21):3320–3327 Singh RK, Lokeshwar BL (2011) The IL-8-regulated chemokine receptor CXCR7 stimulates EGFR signaling to promote prostate cancer growth. Cancer Res 71(9):3268–3277

38

Y. Xun et al.

Singh S, Varney M, Singh RK (2009a) Host CXCR2-dependent regulation of melanoma growth, angiogenesis, and experimental lung metastasis. Cancer Res 69(2):411–415 Singh S, Sadanandam A, Nannuru KC, Varney ML, Mayer-Ezell R, Bond R et al (2009b) Smallmolecule antagonists for CXCR2 and CXCR1 inhibit human melanoma growth by decreasing tumor cell proliferation, survival, and angiogenesis. Clin Cancer Res 15(7):2380–2386 Singh S, Singh R, Sharma PK, Singh UP, Rai SN, Chung LW et al (2009c) Serum CXCL13 positively correlates with prostatic disease, prostate-specific antigen and mediates prostate cancer cell invasion, integrin clustering and cell adhesion. Cancer Lett 283(1):29–35 Singh R, Kapur N, Mir H, Singh N, Lillard JW Jr, Singh S (2016) CXCR6-CXCL16 axis promotes prostate cancer by mediating cytoskeleton rearrangement via Ezrin activation and alphavbeta3 integrin clustering. Oncotarget 7(6):7343–7353 Song ZY, Wang F, Cui SX, Gao ZH, Qu XJ (2019) CXCR7/CXCR4 heterodimer-induced histone demethylation: a new mechanism of colorectal tumorigenesis. Oncogene 38(9):1560–1575 Steele CW, Karim SA, Leach JDG, Bailey P, Upstill-Goddard R, Rishi L et al (2016) CXCR2 inhibition profoundly suppresses metastases and augments immunotherapy in pancreatic ductal adenocarcinoma. Cancer Cell 29(6):832–845 Sueoka H, Hirano T, Uda Y, Iimuro Y, Yamanaka J, Fujimoto J (2014) Blockage of CXCR2 suppresses tumor growth of intrahepatic cholangiocellular carcinoma. Surgery 155(4):640–649 Sun L, Clavijo PE, Robbins Y, Patel P, Friedman J, Greene S et al (2019) Inhibiting myeloidderived suppressor cell trafficking enhances T cell immunotherapy. JCI Insight 4(7):e126853 Takama Y, Miyagawa S, Yamamoto A, Firdawes S, Ueno T, Ihara Y et al (2011) Effects of a calcineurin inhibitor, FK506, and a CCR5/CXCR3 antagonist, TAK-779, in a rat small intestinal transplantation model. Transpl Immunol 25(1):49–55 Tazzyman S, Barry ST, Ashton S, Wood P, Blakey D, Lewis CE et al (2011) Inhibition of neutrophil infiltration into A549 lung tumors in vitro and in vivo using a CXCR2-specific antagonist is associated with reduced tumor growth. Int J Cancer 129(4):847–858 Thelen M, Thelen S (2008) CXCR7, CXCR4 and CXCL12: an eccentric trio? J Neuroimmunol 198 (1–2):9–13 Trotta T, Costantini S, Colonna G (2009) Modelling of the membrane receptor CXCR3 and its complexes with CXCL9, CXCL10 and CXCL11 chemokines: putative target for new drug design. Mol Immunol 47(2–3):332–339 Tulotta C, Stefanescu C, Beletkaia E, Bussmann J, Tarbashevich K, Schmidt T et al (2016) Inhibition of signaling between human CXCR4 and zebrafish ligands by the small molecule IT1t impairs the formation of triple-negative breast cancer early metastases in a zebrafish xenograft model. Dis Model Mech 9(2):141–153 Uchida D, Kuribayashi N, Kinouchi M, Sawatani Y, Shimura M, Mori T et al (2018) Effect of a novel orally bioavailable CXCR4 inhibitor, AMD070, on the metastasis of oral cancer cells. Oncol Rep 40(1):303–308 Varney ML, Singh S, Li A, Mayer-Ezell R, Bond R, Singh RK (2011) Small molecule antagonists for CXCR2 and CXCR1 inhibit human colon cancer liver metastases. Cancer Lett 300 (2):180–188 Wagsater D, Dimberg J (2004) Expression of chemokine receptor CXCR6 in human colorectal adenocarcinomas. Anticancer Res 24(6):3711–3714 Wagsater D, Hugander A, Dimberg J (2004) Expression of CXCL16 in human rectal cancer. Int J Mol Med 14(1):65–69 Wald O, Izhar U, Amir G, Avniel S, Bar-Shavit Y, Wald H et al (2006) CD4+CXCR4highCD69+ T cells accumulate in lung adenocarcinoma. J Immunol 177(10):6983–6990 Walenkamp AME, Lapa C, Herrmann K, Wester HJ (2017) CXCR4 ligands: the next big hit? J Nucl Med 58(Suppl 2):77S–82S Walters MJ, Ebsworth K, Berahovich RD, Penfold ME, Liu SC, Al Omran R et al (2014) Inhibition of CXCR7 extends survival following irradiation of brain tumours in mice and rats. Br J Cancer 110(5):1179–1188

CXC Chemokine Receptors in the Tumor Microenvironment and an Update of. . .

39

Wang B, Hendricks DT, Wamunyokoli F, Parker MI (2006) A growth-related oncogene/CXC chemokine receptor 2 autocrine loop contributes to cellular proliferation in esophageal cancer. Cancer Res 66(6):3071–3077 Wang J, Shiozawa Y, Wang J, Wang Y, Jung Y, Pienta KJ et al (2008) The role of CXCR7/RDC1 as a chemokine receptor for CXCL12/SDF-1 in prostate cancer. J Biol Chem 283(7):4283–4294 Wang J, Hu W, Wang K, Yu J, Luo B, Luo G et al (2016) Repertaxin, an inhibitor of the chemokine receptors CXCR1 and CXCR2, inhibits malignant behavior of human gastric cancer MKN45 cells in vitro and in vivo and enhances efficacy of 5-fluorouracil. Int J Oncol 48(4):1341–1352 Wang C, Chen W, Shen J (2018a) CXCR7 targeting and its major disease relevance. Front Pharmacol 9:641 Wang M, Yang X, Wei M, Wang Z (2018b) The role of CXCL12 axis in lung metastasis of colorectal cancer. J Cancer 9(21):3898–3903 Wei Y, Lao XM, Xiao X, Wang XY, Wu ZJ, Zeng QH et al (2019) Plasma cell polarization to the immunoglobulin G phenotype in hepatocellular carcinomas involves epigenetic alterations and promotes hepatoma progression in mice. Gastroenterology 156(6):1890–904 e16 Wendel M, Galani IE, Suri-Payer E, Cerwenka A (2008) Natural killer cell accumulation in tumors is dependent on IFN-gamma and CXCR3 ligands. Cancer Res 68(20):8437–8445 Winkler AE, Brotman JJ, Pittman ME, Judd NP, Lewis JS Jr, Schreiber RD et al (2011) CXCR3 enhances a T-cell-dependent epidermal proliferative response and promotes skin tumorigenesis. Cancer Res 71(17):5707–5716 Wu B, Chien EY, Mol CD, Fenalti G, Liu W, Katritch V et al (2010) Structures of the CXCR4 chemokine GPCR with small-molecule and cyclic peptide antagonists. Science 330 (6007):1066–1071 Wu Y, Tian L, Xu Y, Zhang M, Xiang S, Zhao J et al (2018) CXCR7 silencing inhibits the migration and invasion of human tumor endothelial cells derived from hepatocellular carcinoma by suppressing STAT3. Mol Med Rep 18(2):1644–1650 Xu JM, Weng MZ, Song FB, Chen JY, Zhang JY, Wu JY et al (2014) Blockade of the CXCR6 signaling inhibits growth and invasion of hepatocellular carcinoma cells through inhibition of the VEGF expression. Int J Immunopathol Pharmacol 27(4):553–561 Xu H, Lin F, Wang Z, Yang L, Meng J, Ou Z et al (2018a) CXCR2 promotes breast cancer metastasis and chemoresistance via suppression of AKT1 and activation of COX2. Cancer Lett 412:69–80 Xu M, Jiang H, Wang H, Liu J, Liu B, Guo Z (2018b) SB225002 inhibits prostate cancer invasion and attenuates the expression of BSP, OPN and MMP2. Oncol Rep 40(2):726–736 Xu L, Liang Z, Li S, Ma J (2018c) Signaling via the CXCR5/ERK pathway is mediated by CXCL13 in mice with breast cancer. Oncol Lett 15(6):9293–9298 Xu S, Tang J, Wang C, Liu J, Fu Y, Luo Y (2019) CXCR7 promotes melanoma tumorigenesis via Src kinase signaling. Cell Death Dis 10(3):191 Yang S, Wang B, Guan C, Wu B, Cai C, Wang M et al (2011) Foxp3+IL-17+ T cells promote development of cancer-initiating cells in colorectal cancer. J Leukoc Biol 89(1):85–91 Yang C, Zheng W, Du W (2016) CXCR3A contributes to the invasion and metastasis of gastric cancer cells. Oncol Rep 36(3):1686–1692 Yang M, Zeng C, Li P, Qian L, Ding B, Huang L et al (2019) Impact of CXCR4 and CXCR7 knockout by CRISPR/Cas9 on the function of triple-negative breast cancer cells. Onco Targets Ther 12:3849–3858 Yi L, Zhou X, Li T, Liu P, Hai L, Tong L et al (2019) Notch1 signaling pathway promotes invasion, self-renewal and growth of glioma initiating cells via modulating chemokine system CXCL12/ CXCR4. J Exp Clin Cancer Res 38(1):339 Yoon MS, Pham CT, Phan MT, Shin DJ, Jang YY, Park MH et al (2016) Irradiation of breast cancer cells enhances CXCL16 ligand expression and induces the migration of natural killer cells expressing the CXCR6 receptor. Cytotherapy 18(12):1532–1542 Yu C, Zhang Y (2019) Characterization of the prognostic values of CXCR family in gastric cancer. Cytokine 123:154785

40

Y. Xun et al.

Yu PF, Huang Y, Han YY, Lin LY, Sun WH, Rabson AB et al (2017) TNFalpha-activated mesenchymal stromal cells promote breast cancer metastasis by recruiting CXCR2(+) neutrophils. Oncogene 36(4):482–490 Yung MM, Tang HW, Cai PC, Leung TH, Ngu SF, Chan KK et al (2018) GRO-alpha and IL-8 enhance ovarian cancer metastatic potential via the CXCR2-mediated TAK1/NFkappaB signaling cascade. Theranostics 8(5):1270–1285 Zeng Y, Li B, Liang Y, Reeves PM, Qu X, Ran C et al (2019) Dual blockade of CXCL12-CXCR4 and PD-1-PD-L1 pathways prolongs survival of ovarian tumor-bearing mice by prevention of immunosuppression in the tumor microenvironment. FASEB J 33(5):6596–6608 Zhang Q, Sun L, Yin L, Ming J, Zhang S, Luo W et al (2013) CCL19/CCR7 upregulates heparanase via specificity protein-1 (Sp1) to promote invasion of cell in lung cancer. Tumour Biol 34 (5):2703–2708 Zhao L, Lim SY, Gordon-Weeks AN, Tapmeier TT, Im JH, Cao Y et al (2013) Recruitment of a myeloid cell subset (CD11b/Gr1 mid) via CCL2/CCR2 promotes the development of colorectal cancer liver metastasis. Hepatology 57(2):829–839 Zhao W, Xu Y, Xu J, Wu D, Zhao B, Yin Z et al (2015) Subsets of myeloid-derived suppressor cells in hepatocellular carcinoma express chemokines and chemokine receptors differentially. Int Immunopharmacol 26(2):314–321 Zheng Z, Cai Y, Chen H, Chen Z, Zhu D, Zhong Q et al (2018) CXCL13/CXCR5 axis predicts poor prognosis and promotes progression through PI3K/AKT/mTOR pathway in clear cell renal cell carcinoma. Front Oncol 8:682 Zhou SL, Zhou ZJ, Hu ZQ, Li X, Huang XW, Wang Z et al (2015) CXCR2/CXCL5 axis contributes to epithelial-mesenchymal transition of HCC cells through activating PI3K/Akt/GSK-3beta/ snail signaling. Cancer Lett 358(2):124–135 Zhou H, Wu J, Wang T, Zhang X, Liu D (2016) CXCL10/CXCR3 axis promotes the invasion of gastric cancer via PI3K/AKT pathway-dependent MMPs production. Biomed Pharmacother 82:479–488 Zhou KX, Xie LH, Peng X, Guo QM, Wu QY, Wang WH et al (2018a) CXCR4 antagonist AMD3100 enhances the response of MDA-MB-231 triple-negative breast cancer cells to ionizing radiation. Cancer Lett 418:196–203 Zhou Y, Guo L, Sun H, Xu J, Ba T (2018b) CXCR5(+) CD8 T cells displayed higher activation potential despite high PD-1 expression, in tumor-involved lymph nodes from patients with thyroid cancer. Int Immunopharmacol 62:114–119 Zhu G, Yan HH, Pang Y, Jian J, Achyut BR, Liang X et al (2015a) CXCR3 as a molecular target in breast cancer metastasis: inhibition of tumor cell migration and promotion of host anti-tumor immunity. Oncotarget 6(41):43408–43419 Zhu Z, Zhang X, Guo H, Fu L, Pan G, Sun Y (2015b) CXCL13-CXCR5 axis promotes the growth and invasion of colon cancer cells via PI3K/AKT pathway. Mol Cell Biochem 400 (1–2):287–295 Zhu WB, Zhao ZF, Zhou X (2019) AMD3100 inhibits epithelial-mesenchymal transition, cell invasion, and metastasis in the liver and the lung through blocking the SDF-1alpha/CXCR4 signaling pathway in prostate cancer. J Cell Physiol 234(7):11746–11759

Rev Physiol Biochem Pharmacol (2020) 178: 41–82 https://doi.org/10.1007/112_2020_39 © The Author(s), under exclusive licence to Springer Nature Switzerland AG 2020 Published online: 14 August 2020

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal Dystrophies Darpan Malhotra and Joseph R. Casey

Contents 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 The Cornea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Stroma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Descemet’s Membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Endothelium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 Roles of Corneal Endothelium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Supplying Corneal Nutrition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Maintenance of Corneal Transparency by the Endothelial Pump . . . . . . . . . . . . . . . . . . . . . 3 Corneal Dystrophies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 Endothelial Corneal Dystrophies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1 Fuchs Endothelial Corneal Dystrophy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Congenital Hereditary Endothelial Dystrophy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Harboyan Syndrome (HS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Endothelial Corneal Dystrophies Genes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 Molecular Mechanisms of Endothelial Corneal Dystrophies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 Sex and Environmental Influences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 RNA Toxicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 Oxidative Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4 Mitochondrial Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5 Epithelial-Mesenchymal Transition (EMT) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6 Cell Adhesion Defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7 ER Stress and the Unfolded Protein Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

D. Malhotra Department of Biochemistry, University of Alberta, Edmonton, AB, Canada Membrane Protein Disease Research Group, University of Alberta, Edmonton, AB, Canada J. R. Casey (*) Department of Biochemistry, University of Alberta, Edmonton, AB, Canada Membrane Protein Disease Research Group, University of Alberta, Edmonton, AB, Canada Department of Physiology, University of Alberta, Edmonton, AB, Canada Department of Ophthalmology and Visual Science, University of Alberta, Edmonton, AB, Canada e-mail: [email protected]

42 42 44 44 45 47 47 47 48 49 49 50 52 52 56 57 57 59 60 63 63 66

42

D. Malhotra and J. R. Casey

5.8 Water Pump Defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.9 Altered Gene Expression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

68 70 70 71

Abstract The cornea, the eye’s outermost layer, protects the eye from the environment. The cornea’s innermost layer is an endothelium separating the stromal layer from the aqueous humor. A central role of the endothelium is to maintain stromal hydration state. Defects in maintaining this hydration can impair corneal clarity and thus visual acuity. Two endothelial corneal dystrophies, Fuchs Endothelial Corneal Dystrophy (FECD) and Congenital Hereditary Endothelial Dystrophy (CHED), are blinding corneal diseases with varied clinical presentation in patients across different age demographics. Recessive CHED with an early onset (typically age: 0–3 years) and dominantly inherited FECD with a late onset (age: 40–50 years) have similar phenotypes, although caused by defects in several different genes. A range of molecular mechanisms have been proposed to explain FECD and CHED pathology given the involvement of multiple causative genes. This critical review provides insight into the proposed molecular mechanisms underlying FECD and CHED pathology along with common pathways that may explain the link between the defective gene products and provide a new perspective to view these genetic blinding diseases. Keywords Cell adhesion · CHED · COL8A2 · Corneal dystrophy · Corneal endothelium · Descemet’s membrane · FECD · Molecular mechanism · SLC4A11 · TCF4

1 Introduction 1.1

The Cornea

The cornea, the outer portion of the eye, is convex, transparent, avascular, with two principal functions: (1) primary structural, protective barrier for the rest of the eye and (2) providing a significant fraction of the eye’s focusing power. The cornea is the most densely innervated tissue of the body and most of its nerves are sensory (Labetoulle et al. 2019). Human cornea consists of five layers: three cellular layers: epithelium, stroma and endothelium, and two acellular connective tissue structures: Bowman’s layer and Descemet’s membrane (DM) (Fig. 1). This review centers on two specific posterior corneal dystrophies (PCDs), caused by defects in the endothelium. We will thus only discuss the endothelium and adjacent regions contributing to PCDs.

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

43

Choroid Vitreous chamber Vitreous humor Sclera Retina Conjunctiva Optic Nerve Posterior chamber Anterior chamber Aqueous Humor

A

Iris Lens Cornea Ciliary body

B Epithelium Bowman’s Layer

Stroma

Keratocytes

Descemet’s Membrane

20 µm

Endothelium

Fig. 1 Structure of the human eye and the cornea. (a) Schematic representation of the human eye with outside on the right. The outermost layer consists of cornea and the sclera, which is covered by a transparent mucous layer of conjunctiva. The middle layer consists of the lens (that receives the light and focuses it to the retina in the inner layer), ciliary body, iris, and the choroid. The inner layer consists of the retina, the sensory region that receives the light and transmits the signal to the brain through the optic nerve. (b) Stained cross-section of the cornea’s five layers, courtesy of Dr. Mathieu Thériault, University of Laval. Epithelium is the outermost layer of the cornea. Going toward the interior is the Bowman’s layer, the stroma, DM, and the endothelium

44

1.2

D. Malhotra and J. R. Casey

Stroma

Stroma, the central corneal layer which is approximately 450-μm thick in humans, comprises about 90% of the total corneal volume (Meek and Knupp 2015). Major components of stroma are collagen fibers, composed of collagen I and V, which form heterodimers with a narrow diameter. These fibers are surrounded by proteoglycans (chondroitin sulfate and keratan sulfate), which maintain corneal hydration (see section “Maintenance of corneal transparency by the endothelial pump”). These collagen fibers precisely arrange into parallel bundles called fibrils that further pack as parallel layers (lamellae) to sustain stromal architecture and transparency. Stroma is comprised of 200–250 lamellae arranged at right angle to the fibers in adjacent lamellae, resulting in reduced light scattering and corneal transparency (Meek and Boote 2004; Torricelli and Wilson 2014). Keratocytes, the resident cells of the stroma, secrete collagens and glycosamines and are located throughout stroma. Stromal avascularity critically contributes to corneal transparency. Moreover, corneal architecture and its hydration state are essential to maintain its refractive index and clarity.

1.3

Descemet’s Membrane

DM is a basement membrane, composed of extracellular matrix (ECM) proteins secreted by the corneal endothelium on which endothelial cells adhere. DM’s collagenous and non-collagenous proteins have a more complex organization than other basement membranes in the body. DM, first described by Descemet in 1758 (Johnson et al. 1982), is secreted throughout life by the endothelium. DM has an anterior banded layer and a posterior non-banded layer. In the anterior banded layer, collagen fibrils have a lattice-like organization with continuous banding pattern at 110 nm intervals. Unlike other basement membranes in which collagen IV is common, collagen VIII is the specific major structural component of DM. Other DM proteins include collagen I, collagen IV, collagen XII, collagen XVIII, fibronectin (FNC), laminins, osteonectin, and versican (Schlotzer-Schrehardt et al. 2011). While FNC and vitronectin concentrate on the stromal side of DM, other proteins are prominent facing the endothelium. Two collagen VIII proteins in DM, collagen VIII alpha chain 1 (COL8A1) and collagen VIII alpha chain 2 (COL8A2), form trimers of two COL8A1 and one COL8A2 that further arrange into hexagonal lattices in the DM (Kapoor et al. 1986). DM provides a structural support for the endothelium (Eghrari et al. 2015).

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

1.4

45

Endothelium

Corneal endothelium, a single flat monolayer of highly specialized cells [corneal endothelial cells (CEC)], adheres to DM (Fig. 1). CEC, which are 5 μm deep and 20 μm in diameter, arrange in hexagonal lattices forming a honeycomb-like mosaic. Humans have a significant reserve of CEC at birth, about 5,000–7,000 cells/mm2 (Fig. 2). CEC density declines rapidly, by 45% in the first year and then it continues to decrease until the mid-twenties (Wilson and Roper-Hall 1982; Sherrard et al. 1987). CEC density declines more gradually between 20 and 80 years of age, at a constant rate of 0.5% per year (Tuft and Coster 1990; Sanchis-Gimeno et al. 2005). The endothelium maintains corneal hydration state, which is critical for clear vision. CEC have ion-transport systems that form an “endothelial pump,” continuously moving fluid from stroma to aqueous humor to maintain stroma hydration state at 78% (% water in stroma by weight) (Fig. 3). Homeostatic control of this corneal deturgescence is essential to maintain precise collagen fibril organization, essential for clear vision (Willoughby et al. 2010). Since CEC do not proliferate (Joyce 2012), cell damage leads to irreplaceable CEC loss. CEC do, however, enlarge (polymegathism), change shape (pleomorphism), and migrate to fill the space left by cell death to maintain an intact Healthy FECD CHED

CEC density/mm2

6000 5000 4000 3000 2000

Symptomatic Threshold

1000 0

0

10

20

30

40

50

60

70

80

90 100

Age (years) Fig. 2 CEC density declines with age in healthy individuals, FECD and CHED patients. CEC density profiles from several studies were analyzed to present cell density with age in healthy individuals, FECD and CHED patients (Wilson and Roper-Hall 1982; Sherrard et al. 1987; Syed et al. 2017; McLaren et al. 2014; Iovino et al. 2018; Sultana et al. 2007; Kenyon and Antine 1971; Abdellah et al. 2019; Arici et al. 2014; Duman et al. 2016; Ewete et al. 2016; Angmo et al. 2018; Mittal et al. 2011). CEC density of approximately 6,000 cells per mm2 at birth declines rapidly in the first few years reaching a steady rate in adulthood. Decline in CEC density below a critical threshold (symptomatic threshold) (Syed et al. 2017; McLaren et al. 2014; Iovino et al. 2018) may compromise the endothelial monolayer integrity which further compromises the endothelial pump, resulting in corneal edema. FECD patients see a faster decline in CEC density approaching the critical threshold faster than healthy individuals. CHED patients are born with 5–10 times lower CEC density than healthy individuals, leaving them close to the symptomatic threshold

46

D. Malhotra and J. R. Casey

Epithelium Bowman’s Layer

Collagen Lamellae

Stroma

Keratocytes Osmotic Leak

Descemet’s Membrane

Endothelial Pump

Endothelium

Aqueous Humor

Stroma

H2O, + H+/OH- H

Na+ 2HCO3- Na+

NBCe1

SLC4A11

NHE1

H+

Basolateral H+ Lactate

MCT1

Na+

Kir2.1

NKCC1

K+

3Na+

Na-K ATPase 2K+

HCO3AE2 Cl-

TJ

Cl- or HCO3-

CACC1

Aqueous Humor

CFTR

H 2O

AQP1

H+

Lactate

Endothelial Pump

Osmotic Leak

H+

K+ 2Cl-

TJ

MCT4

Apical

Fig. 3 The Corneal Endothelial Pump maintains corneal fluid balance and transparency. Upper panel presents cornea structure with five distinct layers. Stroma features collagen lamellae organized in a distinct fashion to prevent light scattering, and keratocytes with vital roles in stromal biology. The osmotic leak represents the endothelial leaky barrier, which allows nutrient diffusion to the stroma. Fluid osmotically leaks into the stroma and is pumped back to the aqueous humor by the endothelial pump as shown. The lower panel is a zoomed view of the CEC with basolateral side facing the stroma and the apical side facing the aqueous humor. Proteins that contribute to the endothelial pump include basolateral components NBCe1, NHE1, SLC4A11, MCT1, NKCC1, AE2, Kr2.1 and Na+-K+ ATPase and apical proteins CACC1, CFTR, AQP1, and MCT4. TJ represents tight junctions. HCO3 transport by basolateral NBCe1 and AE2 and lactate transport by MCT1 into the cell drives fluid secretion. Apical CACC1 and CFTR move HCO3 and MCT4 transports lactate across the apical surface to drive fluid movement into the aqueous humor. Basolateral SLC4A11 and apical AQP1 form a water permeation pathway from stroma to aqueous humor. Adapted from (Bonanno 2012; Vilas et al. 2013)

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

47

endothelial monolayer. With constant loss in number, increased polymegathism and pleomorphism arise in CEC, accompanied by reduced ability of CEC to deturgesce the cornea. Decline in CEC count below 500–700 cells/mm2 disturbs the endothelial monolayer integrity (Fig. 2), compromises the endothelial pump, and causes corneal edema (DelMonte and Kim 2011). The focus of this review is on endothelial corneal dystrophies (ECDs) that arise from compromised endothelial physiology.

2 Roles of Corneal Endothelium The cornea forms an optically clear protective layer by precise organization of collagen stromal fibrils to ensure light transmission to the retina. While stroma is the unique structural component of the cornea, the endothelium crucially maintains healthy corneal physiology.

2.1

Supplying Corneal Nutrition

Corneal avascularity is required for corneal transparency, but lack of blood supply presents a challenge to provide nutrients to stromal keratocytes and corneal epithelial cells. The endothelium, therefore, provides a pathway to provide nutrients to these cells from the aqueous humor. Corneal epithelium receives oxygen, calcium, magnesium, potassium, and sodium chloride through the tears (Bourne 2003). In contrast, other nutrients, including glucose, are virtually absent in tear fluid. The adjacent limbal vasculature provides about 20% of the glucose demand to the cornea, yet ablation of this source does not disrupt corneal physiology (Sweeney et al. 1998). Aqueous humor is the main source of glucose and other essential nutrients. This supply is achieved by the “leaky” nature of the endothelium. Corneal endothelium acts as a partially leaky barrier to allow diffusion of these solutes from the aqueous humor into the stroma through the tight junctions (Fig. 3) (Bonanno 2012). Blocking this supply route induces stromal and epithelial degeneration (Bourne 2003).

2.2

Maintenance of Corneal Transparency by the Endothelial Pump

Stromal hydration is maintained by several factors, including endothelial tight junctions, endothelial pump function, and the ability of CEC to accommodate for the continual cell loss by altering their shape and size. CEC are polarized cells with their basolateral surface facing DM and apical surface toward aqueous humor.

48

D. Malhotra and J. R. Casey

Movement of material between CEC (paracellular pathway) is limited by the connections between cells (Barry et al. 1995; Petroll et al. 1999). The endothelial “pump” is a two-step process to move excess fluid from stroma to aqueous humor against its osmotic gradient (Fig. 3), which requires substantial energy expenditure (Bonanno 2012). Solutes are moved against their electrochemical gradients to promote the exit of water along a local osmotic gradient. CEC proteins include SLC4A11, Na+/K+-ATPase, Na+/H+ exchanger 1 (NHE1), Cl/ HCO3 anion exchanger 2 (SLC4A2 or AE2), electrogenic Na+/HCO3 cotransporter (SLC4A4 or NBCe1), Na+:K+:2Cl cotransporter (NKCC1), cystic fibrosis transmembrane conductance regulator (CFTR), calcium-activated chloride channels (CaCC1), monocarboxylate transporters 1 and 4 (MCT1 and 4), and aquaporin 1 (AQP1) (Vilas et al. 2013; Jalimarada et al. 2013; Bonanno 2003). Models have been proposed to explain the endothelial pump mechanism (Bonanno 2003, 2012; Jalimarada et al. 2013). Movement of ions across the basolateral and apical surface of CEC may generate localized osmotic gradients, sometimes driven by negative membrane potential, which leads to the movement of fluid from the stroma back to aqueous humor (Bonanno 2003, 2012; Fischbarg 2012; Schey et al. 2014). SLC4A11, at the basolateral surface of CEC (facing stroma), facilitates the direct movement of water into the cells, which could be further moved into the aqueous humor by apical AQP1 (facing aqueous humor) (Vilas et al. 2013). The combined leaky barrier function and the fluid pump, referred to as the pump-leak mechanism, is essential to maintain healthy corneal physiology and transparency. Failures of CEC pumping mechanism give rise to stromal edema and visual haziness by distorting the stromal collagen fibril array (Meek and Boote 2004).

3 Corneal Dystrophies Corneal dystrophies are classified by their anatomical location in the layer where the defect lies: (1) superficial or anterior corneal dystrophies caused by defects in the epithelium, basal lamina or the Bowman’s layer, (2) stromal corneal dystrophies caused by the defects in the corneal stroma, and (3) posterior or endothelial corneal dystrophies caused by defects in the DM or the endothelium. Most follow a Mendelian pattern of inheritance [(autosomal dominant (AD), autosomal recessive (AR), or X-linked (XLD)] and have a variable clinical presentation with different ages of onset (Vincent 2014). This review focuses on diseases arising from defects in the DM or the endothelium and hence the ECD. More particularly, this review centers on two ECD: Congenital Hereditary Endothelial Dystrophy (OMIM #217700) and Fuchs Endothelial Corneal Dystrophy (OMIM #136800). Patients present with corneal edema, variable corneal opacity, and compromised visual acuity (Klintworth 2009). In some CHED patients (see “Harboyan syndrome” below) hearing deficits are present and hearing loss occurs at a high frequency in FECD patients than the general population (Stehouwer et al. 2011). Symptoms beyond cornea and hearing have not been reported in corneal dystrophy patients.

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

49

4 Endothelial Corneal Dystrophies ECDs are characterized by defects in the corneal endothelium or the DM secreted by the endothelium. Disruptions of endothelial monolayer compromise the endothelial pump, resulting in excessive fluid in the stroma, causing corneal edema and a decline in visual acuity. The three prominent diseases discussed in this review are Fuchs Endothelial Corneal Dystrophy, Congenital Hereditary Endothelial Dystrophy, and Harboyan Syndrome (HS) (Table 1).

4.1

Fuchs Endothelial Corneal Dystrophy

FECD arises due to a complex combination of genetic factors with two different ages of onset in patients. Early onset happens around the second to third decade of life and is well characterized, but rare. Late onset occurs around the fourth to fifth decade and accounts for the majority of patients, with heterogeneous genetics and physiological etiology (Biswas et al. 2001) (Table 1). FECD, which arises due to defects in the corneal endothelium, has a regionally varied prevalence of 4–9% in individuals over 40 years of age (Soh et al. 2020) and is one of the leading causes for corneal transplants globally (Feizi 2018). FECD arises from AD inheritance of several Table 1 Genes established to cause or implicated in CHED and FECD Endothelial corneal dystrophy CHED/FECD

Gene SLC4A11

Gene function H2O/H+, NH3 transport, cell adhesion

CHED/FECD

MPDZ

FECD

SLC4A11

Protein–protein interactions stabilization of intercellular connections H2O/H+, NH3 transport, cell adhesion

FECD

COL8A2

FECD

COL17A1

FECD

LAMC1

FECD FECD FECD FECD

ATPB1 LOXHD1 TCF4 ZEB1 (TCF8) AGBL1 DMPK

FECD FECD

Collagen-Descemet’s membrane component Collagen-Descemet’s membrane component Laminin-Descemet’s membrane component Na+/K+-ATPase subunit Protein trafficking Transcription factor Transcription factor Glutamate decarboxylase Dystrophia myotonia protein kinase

Possible disease mechanisms Water pump/mitochondria ROS/cell adhesion Water pump Water pump/mitochondria ROS/cell adhesion Cell adhesion/ER stress Cell adhesion/ER stress Cell adhesion/ER stress Water pump/ER stress ER stress RNA toxicity RNA toxicity RNA toxicity RNA toxicity

50

D. Malhotra and J. R. Casey

different genes (see below). Age of onset and degree of penetrance may be influenced by environmental factors and the specific disease allele inherited and is characterized by a progressive decline in CEC density, CEC polymegathism, and pleomorphism in patients (Feizi 2018; EBAA 2017) (Fig. 4). In FECD patients, CEC density declines until the endothelial pump fails (symptomatic threshold), which compromises corneal deturgescence (Fig. 2). Stromal fluid accumulation increases corneal thickness and cloudiness, which compromises visual clarity. Patients also have bead-like deposits of ECM on DM, called guttae. Non-confluent guttae on the peripheral DM are normal with progressing age (Hassall–Henle warts), but FECD patients have guttae in the central cornea that enlarge to grow confluent as the disease progresses (Eghrari and Gottsch 2010) (Fig. 4). FECD typically progresses in four stages: Stage I is asymptomatic but non-confluent guttae are present in the central cornea. In stage II, guttae become confluent with loss of CEC, polymegathism, and pleomorphism. In stage III, the endothelial pump is compromised, and corneal edema ensues. By stage IV, corneal edema severity causes corneal scarring, diminishing visual acuity (Feizi 2018). With the compromised endothelial pump, stromal fluid accumulation produces a bluegray stromal haze. Eventually, this fluid accumulation increases corneal thickness and gives the cornea a ground glass-like appearance. Ultimately, excess fluid may accumulate between epithelial cells and sub-epithelium. These bubbles burst through the epithelium causing painful corneal erosions, bullous keratopathy (BK). Symptoms of epithelial edema subsequently reduce but visual acuity continues to decline as the epithelium secretes connective tissue to repair itself (Klintworth 2009). FECD begins in the central cornea and radially progresses to peripheral cornea with extensive DM wrinkling. In advanced FECD, abnormalities arise in all corneal layers, but consistent changes are observed in DM and the endothelium. CEC are progressively lost over the guttae. Abnormal CEC are significantly larger in size, have widened intercellular spaces, swollen mitochondria, and dilated rough endoplasmic reticulum (ER). Some CEC also develop fibroblast-like morphology. DM thickness increases to about fourfold due to excessive ECM deposition. Size and confluence of guttae also increase in the central DM where they may protrude into the anterior chamber (Elhalis et al. 2010).

4.2

Congenital Hereditary Endothelial Dystrophy

CHED is a rare ECD with symptoms present in infancy or within the first few years of life, although sometimes symptoms are not reported until age 5–10 (Brejchova et al. 2019). CHED has an AR mode of inheritance and is thus most prevalent in countries with high incidence of consanguineous marriages (Vithana et al. 2006). CHED is characterized by corneal opacification in infancy with twofold to threefold increase in corneal thickness. Patient corneas have a ground glass-like appearance and present with extensive stromal edema. Patients are born with significantly decreased CEC density, with a fibrotic morphology. CHED patients also have

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

A

B

C

D

51

Fig. 4 Clinical presentation of CEC loss in FECD and CHED patients. Representative images of specular microscopy examination of ECD patient eyes reveal the change in CEC shape, density and appearance of DM. (a) Healthy corneal endothelium with hexagonal shaped CEC arranged tightly in monolayer. (b, c) Loss of CEC is visible along with polymegathism, pleomorphism, and exposed underlying DM (black). (d) Significant CEC loss in advanced stages of FECD exposes the underlying DM (black), which disrupts corneal endothelium integrity and the endothelial pump. (Procured and modified with consent from Konan Medical, CA, USA)

52

D. Malhotra and J. R. Casey

thickened DM, but guttae are not usually evident. DM has a normal anterior banded layer and a posterior non-banded layer, but with an additional 20–40 nm thick fibrous layer at the posterior surface consisting of a mixture of ECM-like fibrous material. CHED is not a progressive disease with symptoms that remain stable throughout the patient’s life (Klintworth 2009).

4.3

Harboyan Syndrome (HS)

HS or Corneal Dystrophy and Perceptive Deafness (CDPD) is a sensorineural hearing disorder associated with some cases of CHED, although HS is also seen as a manifestation of CHED. Infants with HS are born with a compromised visual acuity, but they develop progressive hearing deficits usually around 10–15 years of age (Desir and Abramowicz 2008). A single pedigree revealed that CHED patients eventually develop HS, although their hearing is not routinely examined, which may be due to lack of information and infrastructure in developing countries (Siddiqui et al. 2014). In some cases, parents of CHED patients are affected by FECD and are carriers of disease alleles, indicating that homozygous inheritance of FECD alleles causes CHED, but a single allele may suffice to cause FECD (Kim et al. 2015). The presence of all three diseases in the same pedigree suggests changes in clinical practices to monitor for HS in all CHED patients and for FECD in their parents (Siddiqui et al. 2014). The lack of reports of symptoms beyond vision and hearing may reflect the prevalence of CHED/HS patients in developing countries where parents may not report non-corneal symptoms due to lack of information/infrastructure. More thorough investigations in patients are needed to detect possible disease manifestation beyond vision and hearing.

4.4

Endothelial Corneal Dystrophies Genes

Amongst ECD, FECD has an AD mode of transmission, although sporadic cases are more prevalent. Multiple genetic and environmental factors underlie the FECD pathology. The genetic basis of FECD is heterogeneous with incomplete penetrance and variable expressivity. Mutations in FECD (Vedana et al. 2016; Weiss et al. 2015) have been found in genes (Table 1) including: COL8A2 (Biswas et al. 2001), SLC4A11 (Vithana et al. 2008), TCF4 (Baratz et al. 2010), ZEB1 (formerly called TCF8) (Riazuddin et al. 2010), AGBL1 (Riazuddin et al. 2013a), LOXHD1 (Riazuddin et al. 2012), and MPDZ (Moazzeni et al. 2019). Genome-wide association studies (GWAS) implicated KANK4, LAMC1, and ATPB1 in the pathogenesis of FECD (Afshari et al. 2017). Thus far, CHED has been found to arise only from AR inheritance of SLC4A11 (Vithana et al. 2006) and possibly MPDZ (Moazzeni

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

53

et al. 2019) mutations. How can the pathology of FECD and CHED be rationalized by the involvement of this set of genes? COL8A2 COL8A2 codes for collagen VIIIA2 protein which is the principal component of DM and secreted by CEC. It associates with collagen VIIIA1 to form heterotrimers organized into highly ordered three-dimensional arrays in DM (Greenhill et al. 2000). Two COL8A2 mutations, p.Leu450Trp (Gottsch et al. 2005) and p.Gln455Lys (Biswas et al. 2001), cause abnormal intracellular accumulation and distort DM arrangement. Knock-in mouse models of both Col8a2 mutations mimic the early onset FECD phenotype in mice (Meng et al. 2013; Jun et al. 2012). COL8A2 mutations also cause a type of posterior polymorphous corneal dystrophy (PPCD) (OMIM #609140 (Biswas et al. 2001). Since this finding has not been confirmed, PPCD has similarities to FECD, and members of the index family had FECD, this finding may represent a special presentation of FECD rather than PPCD (Biswas et al. 2001). SLC4A11 SLC4A11 is a large (891 amino acids, Accession no. NP_114423.1) multi-domain integral membrane protein encoded by SLC4A11 gene. Identified in 2001 as BTR1 or Bicarbonate Transporter 1, SLC4A11 belongs to the SLC4 bicarbonate transporter gene family (Parker et al. 2001). SLC4A11 is expressed in most body tissues with high expression in cornea, cerebral cortex, epididymis, gallbladder, small intestine, testes, kidneys, breast, cerebellum, lungs, ovaries, pancreas, and stomach (THPA 2019). SLC4A11 localizes at the basolateral surface of CEC where its extracellular region faces the DM (Vilas et al. 2013; Jalimarada et al. 2013). It is one of the most abundantly expressed proteins of the corneal endothelium (Frausto et al. 2014; Chng et al. 2013). Three N-terminal variants of human SLC4A11 have been reported in the genome database, due to differential splicing of the mRNA: SLC4A11 Variant 1 (v1) which encodes for a 918 amino acid protein (NCBI Reference Sequence: NP_001167561.1), SLC4A11 Variant 2 (v2) which encodes for an 891 amino acid splice form 2 (NP_114423), and SLC4A11 Variant 3 (v3) which encodes for an 875 amino acid splice form 3 (NP_001167560). These variants have also been respectively called SLC4A11A, -B, and -C (Kao et al. 2015). These variants differ in their N-termini by 20–60 amino acids. SLC4A11 v2 was the first variant to be cloned in 2001 when the protein was discovered (Parker et al. 2001) and is the most widely used variant in in vitro studies. In human cornea SLC4A11 v1 is not expressed, whereas v2 is highly expressed relative to v3 (Malhotra et al. 2019a; Hara et al. 2019). Multiple transport roles have been attributed to SLC4A11. The plant ortholog of SLC4A11, BOR1, transports borate, and an initial functional assessment indicated that the human protein also transports borate (Park et al. 2004; Takano et al. 2002). Borate cross-links vicinal diol components of cell wall and is critical to their structural integrity in bacteria, plants, and fungi (O'Neill et al. 2004). Mammals do not have cell wall and insignificant to no borate in their blood. Hence, the borate transport function of SLC4A11 is unclear from a physiological standpoint. Later studies found that indeed SLC4A11 is not a borate transporter (Vilas et al. 2013;

54

D. Malhotra and J. R. Casey

Jalimarada et al. 2013). SLC4A11 is a member of the SLC4 family, whose other members are all bicarbonate transporters, yet SLC4A11 does not transport bicarbonate (Vilas et al. 2013; Jalimarada et al. 2013). Human SLC4A11 v2 was reported to function as an amiloride-sensitive Na+-OH (H+) transporter (Ogando et al. 2013). Evidence is converging to indicate that both mouse and human SLC4A11 act as amiloride-insensitive Na+-independent electrogenic H+ (OH) permeation pathways (Myers et al. 2016; Kao et al. 2020). Moreover, SLC4A11 H+ transport function is regulated by pH (Quade et al. 2020). SLC4A11 also facilitates NH3 transport either alone (Loganathan et al. 2016) or by cotransport with H+ (Kao et al. 2020; Zhang et al. 2015). SLC4A11 also facilitates movement of water (Vilas et al. 2013). In addition to its membrane transport function, SLC4A11 was recently found to serve as an attachment site between CEC and DM (Fig. 5) (Malhotra et al. 2019b).

Basolateral

Integrins

SLC4A11

Descemet’s Membrane

COL8A1

COL8A2

COL1

COLIV

Vitronectin

Tenasin

Fibronectins

Laminins

Tight Junction

Fig. 5 Endothelial cell adhesion to Descemet’s membrane. CEC cell–cell and cell-DM attachment maintains the endothelial monolayer and efficient endothelial pump. SLC4A11 and integrins promote CEC-DM attachment and tight junctions maintain cell–cell interactions and endothelial polarity

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

55

Until recently SLC4A11 was the only gene reported as responsible for CHED (Vithana et al. 2006). Now MPDZ mutations (see below) have been suggested as responsible for one case of CHED (Moazzeni et al. 2019). While SLC4A11 mutations cause CHED when inherited recessively, it appears that carrier parents may also develop FECD (Kim et al. 2015; Chaurasia et al. 2020). TCF4 TCF4 or transcription factor 4 encodes the transcription factor 4 that belongs to the E-protein family of class I basic helix-loop-helix (bHLH) transcription factors (Baratz et al. 2010). A CTG (cytosine–thymine–guanine) trinucleotide repeat expansion in the third intron of TCF4 causes the majority of FECD (Baratz et al. 2010). ZEB1 Zinc-finger E-box binding Homeobox 1 protein (ZEB1) is a transcription factor, also known as TCF8 (transcription factor 8). Five late-onset FECD-causing mutations have been identified in TCF8 (Riazuddin et al. 2010). Interestingly, a different set of ZEB1 mutations cause a disease similar to FECD, AD posterior polymorphous corneal dystrophy (OMIM #609141), which has been attributed to ZEB1 haploinsufficiency (Krafchak et al. 2005). AGBL1 AGBL1 mutations have been reported to cause some cases of FECD (Riazuddin et al. 2013a; Zhang et al. 2019). ATP/GTP binding-protein like 1 (AGBL1) is a metallo-carboxypeptidase that mediates removal of glutamyl groups from polyglutamylated proteins. The most notable biological effect of protein glutamylation is to destabilize tubulin (Wloga et al. 2017). The link between tubulin and mechanisms of ECDs is unclear. The role of AGBL1 mutations in FECD should be viewed cautiously as only three cases have been identified. Two of these mutations are predicted to have low pathogenicity, yet one is protein truncation mutant. LOXHD1 LOXHD1 gene (Lipoxygenase Homology Domain-containing 1 protein) is a rare FECD gene whose mutations also cause hearing loss (Riazuddin et al. 2012). This cytosolic protein, which has received too little attention, is composed of 15 PLAT (polycystin/lipoxygenase/α-toxin) domains, which likely are Ca++ binding sites (Grillet et al. 2009). PLAT domains have a role in plasma membrane targeting of client proteins via protein–protein interactions (Grillet et al. 2009). Five LOXHD1 mutations were recently identified to also cause autosomal recessive non-syndromic hearing loss (Bai et al. 2020). KANK4 KANK4 gene encodes KN motif and ankyrin repeat domain containing protein 4 or KANK4 which is a cytoplasmic protein involved in actin polymerization and control of cytoskeletal formation (Kakinuma et al. 2009). A genome wide association study (GWAS) identified KANK4 as an FECD locus, but whether it represents a modifier gene or a disease locus is not yet established (Afshari et al. 2017). LAMC1 LAMC1 encodes Laminin gamma-1 protein (LAMC1), an ECM protein of DM. GWAS identified LAMC1 as an FECD locus, but no point mutations have yet been identified (Afshari et al. 2017).

56

D. Malhotra and J. R. Casey

ATPB1 ATPB1 encodes the beta subunit of the Na+/K+ ATPase. GWAS identified ATBP1 as a locus in FECD, however, its correlation to the disease pathophysiology remains unclear (Afshari et al. 2017). Since the Na+/K+ ATPase is central to the maintenance of the membrane potential and the Na+ and K+ gradients across the plasma membrane of mammalian cells, ATPB1 defects could readily be understood to affect the water pump function of the endothelium. COL17A1 COL17A1 encodes collagen 17A1. Homozygous p.Pro1185Leu mutation has been reported in one FECD patient (Zhang et al. 2019). MPDZ Multi-PDZ domain protein (MPDZ, also known as MUPP1) was identified as homozygously mutated in one CHED patient who lacked any SLC4A11 defect (Moazzeni et al. 2019). One parent of the child with CHED carried an MPDZ mutation and had FECD, suggesting that MPDZ mutations cause CHED when homozygous and FECD when heterozygous (Moazzeni et al. 2019). MPDZ is a cytosolic scaffold protein containing 13 PDZ protein–protein interaction domains that interact with proteins, including those of the tight junction (Feldner et al. 2017). Expressed in the choroid plexus, some MPDZ mutations cause autosomal recessive non-syndromic hydrocephalus (HYC2, OMIM #615219) (Feldner et al. 2017). The mpdz/ mouse has increased paracellular permeability arising from defective intercellular junctions in the choroid plexus (Yang et al. 2019).

5 Molecular Mechanisms of Endothelial Corneal Dystrophies Complex pathophysiology and genetics challenge our understanding of molecular mechanisms underlying ECD. In healthy corneal physiology, endothelial monolayer integrity, its association with the DM (Fig. 5), and pumping function (Fig. 3) are essential to prevent ECD. Factors that compromise the endothelial pump can explain symptoms of corneal edema. Healthy individuals experience declining CEC density at an average rate of 0.5% per year (Wilson and Roper-Hall 1982). When cell density falls below 1,000–1,200 cells/mm2, corneal edema (symptomatic threshold) may develop (Fig. 2) (Syed et al. 2017; McLaren et al. 2014; Iovino et al. 2018). Molecular mechanisms that increase the rate of CEC loss will thus accelerate the development of ECD. In addition, defects in genes integral to the endothelial fluid pump may also promote ECD development. This section highlights the underlying molecular mechanisms that disrupt the endothelial integrity and pump function to explain ECD symptoms.

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

5.1

57

Sex and Environmental Influences

Women have a threefold increased risk of FECD for reasons that are unclear (Musch et al. 2011). Amongst the other factors, smoking has been linked to increased guttae in patients. Smoking in female FECD patients increases the risk of progressing to advanced stages, perhaps due to increased oxidative stress (see below) (Zhang et al. 2013). Diabetes in FECD patients also increases central corneal thickness, which is independent of FECD. FECD patients have a fourfold increase in myocardial infarction, angina pectoris, and cardiac insufficiency compared to control individuals, but the underlying reason remains unexplained (Vedana et al. 2016).

5.2

RNA Toxicity

In FECD, about 70% of the disease burden rests on mutations of the transcription factor, TCF4 (Vasanth et al. 2015). The third intron of TCF4 contains a repeated DNA base sequence of CTG that is prone to expansion. When the size of the CTG array expands to beyond 40, odds of developing FECD rise significantly (Mootha et al. 2015; Soh et al. 2019) and FECD severity correlates with the CTG repeat expansion size in FECD patients (Soliman et al. 2015). The proposed disease mechanism was first established for Myotonic dystrophy type 1 (DM1). CUG-expanded RNA transcripts bind and sequester the RNA splicing regulator, MBNL1 (Mankodi et al. 2001). In the case of FECD and TCF4, this sequestration leads to microscopically detectable foci of MBNL1 protein with CUG-expanded TCF4 transcript (Du et al. 2015) (Fig. 6). Altered splicing of MBNL1 client transcripts arises from insufficiency of MBNL1, which has been detected in corneas from FECD patients (Du et al. 2015). Thus, CTG repeat expansion ultimately may cause disease from defective splicing and expression of critical genes in the endothelium (Fig. 6). Additional complexity arises with the observation that MBNL2 protein, related to MBNL1, is also sequestered by CUG-expanded TCF4 transcripts (Zarouchlioti et al. 2018). Over-expression of MBNL1, however, fails to rescue mice with a repeat expansion in their DMPK1 gene and mRNA splicing remains defective (Yadava et al. 2019). This observation suggests that sequestration of MBNL1 does not fully explain the basis for DM1 and by extension for TCF4 in FECD; other factors may be required in combination with MBNL1 to rescue the splicing defects. The authors emphasize that MBNL1, binding over 2,400 target RNAs, has important roles in RNA metabolism (RNA translation, localization, mRNA decay) and binds spliced RNAs, indicating functions beyond being a splice effector (Yadava et al. 2019). An additional explanation is provided by the observation that splicing factor, MBNL2, is sequestered by CUG-expanded transcripts (Zarouchlioti et al. 2018), so some toxicity likely arises from MBNL2 sequestration in addition to MBNL1. Taken together, MBNL1 has complex interactions with RNAs that may require other

58

D. Malhotra and J. R. Casey

Basolateral

Protein-mRNA aggregates MBNL1/2

CUG CU G C

UG

Fig. 6 TCF4 RNA toxicity. Trinucleotide CUG repeats in the third intron of TCF4 mRNA sequester MBNL1/2, an RNA splicing regulator, which causes accumulation of microscopically detectable foci of mRNA-protein complexes in the CEC nuclei. Loss of MBNL1/2 function due to sequestration could affect the transcriptional splicing of client transcripts. Inability to clear MBNL1/ 2-mRNA complexes causes RNA toxicity that may cause cell death. Halted transcriptional maturation of MBNL1 client transcripts could irreversibly trigger pathways that can additionally contribute to CEC death. This disturbs the endothelial pump and a healthy endothelial physiology leading to FECD onset

factors and binding to CUG-expanded transcripts adversely affects these roles. In addition, inability to clear toxic MBNL-RNA complexes may trigger diseasecausing pathways that cannot be reversed by MBNL overexpression. DM1 arises from CTG repeat expansion in the 30 -untranslated region of the dystrophia myotonia protein kinase gene (DMPK) (Mootha et al. 2017). As is the case with TCF4, the DMPK transcript associates with MBNL1, giving rise to mis-splicing of some MBNL1 target genes. Interestingly, FECD symptoms are tenfold more prevalent in DM1 patients than in the general population (Mootha et al. 2017). Thus, (1) DMPK defects can cause FECD-like corneal symptoms

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

59

(Winkler et al. 2018) and (2) sequestration of MBNL1 leading to aberrant expression of its client genes is the molecular mechanism underlying both TCF4 and DMPK mutations in FECD. FECD mutations in AGBL1 include one nonsense mutation, p.Arg1028X and one missense mutation p.Cys990Ser (Riazuddin et al. 2013b), but their roles in FECD pathology are unclear. AGBL1 binds TCF4 and both identified mutations affect AGBL1-TCF4 interaction. Thus, alterations of the AGBL1-TCF4 complex may contribute to FECD through the same pathway as TCF4 (Riazuddin et al. 2013b). Whereas wild-type AGBL1 localizes to the cytoplasm, truncated p.Arg1028X protein traffics to the nucleus (Riazuddin et al. 2013b). Similarly, ZEB1 (TCF8) and TCF4 modulate each other’s expression (SanchezTillo et al. 2015) and are hypothesized to be a part of the same pathway that causes FECD. A defective step in the transcription cascade containing TCF4, TCF8, and AGBL1 may lead to accumulation of toxic RNA-protein complexes to explain CEC loss in FECD corneas.

5.3

Oxidative Stress

The late onset of FECD suggests disease mechanisms requiring prolonged low-level insults. Indeed, increased rates of CEC loss will lead to FECD. Because of the cornea’s high metabolic activity and exposure to the aerobic atmosphere and ultraviolet light, CEC are especially vulnerable to oxidative damage. Contribution of oxidative mechanisms to CEC health and development of FECD have recently been well summarized (Jurkunas 2018; Vallabh et al. 2017). Oxidative stress in FECD patients is underscored by increased levels of damaged mitochondrial DNA and dysregulation of antioxidant protective pathways (Jurkunas et al. 2010). SLC4A11 is implicated in resistance to oxidative stress. SLC4A11 mutations leave cells more vulnerable to oxidative insults (Roy et al. 2015). Cells with SLC4A11 mutations that cause SLC4A11 protein ER-retention had elevated reactive oxygen species (ROS) levels and impaired mitochondrial function. Antioxidant signalling pathway triggered by nuclear factor erythroid-2 related factor 2 (Nrf2) was impaired by SLC4A11 depletion (Guha et al. 2017). Nrf2 is tightly regulated by KEAP1 (Kelch-like ECH associated protein 1) under physiological conditions where Nrf2 is held in the cytoplasm and constantly degraded by KEAP1-mediated ubiquitination (Kensler et al. 2007; Sun et al. 2007). Under oxidative stress, protein DJ-1 (also known as PARK7 or Parkinsonism-associated deglycase) prevents NRF2-KEAP1 interaction and regulates its nuclear translocation where it activates the antioxidant defense genes (Liu et al. 2014; Bitar et al. 2012). NRF2 (Jurkunas et al. 2010), DJ-1 (Bitar et al. 2012), and SLC4A11 (Gottsch et al. 2003) are downregulated in FECD. TCF4, in osteoblasts, is suggested to be involved in antioxidant defense (Yang et al. 2013) and absence of TCF4 triggers apoptosis (Forrest et al. 2013), suggesting potential transcript regulation of DJ-1 and NRF2 by TCF4 to regulate healthy CEC physiology. Loss of TCF4 expression in most

60

D. Malhotra and J. R. Casey

cases of FECD can lead to loss of antioxidant defense pathways, which may increase ROS in cells leading to apoptosis. Oxidative stress provides a compelling explanation for cell death, which will contribute to loss of CEC density leading to FECD. Nrf2 is also downregulated in CHED (Guha et al. 2017), which suggests direct or indirect feedback regulation of antioxidative pathway by SLC4A11 as CHED is only caused by SLC4A11 mutations. That said, none of the identified FECD or CHED genes (above) have a clear role in oxidative mechanisms. This suggests that factors causing oxidative stress will contribute importantly to these diseases, but most likely by an environmental, or background factor or as secondary effects resulting from other mechanisms described that alter CEC gene expression.

5.4

Mitochondrial Function

Potential mitochondrial roles in FECD progression have recently been reviewed (Miyai 2018). The CEC water pump function is highly energy intensive, making these cells among the body’s most energy-demanding (Chng et al. 2013). High energy metabolism inevitably causes higher levels of damaging free radical species as a fraction of electrons leak from the respiratory electron transport chain (Miyai 2018). Free radicals can damage nuclear and mitochondrial DNA. Cells may respond to mitochondrial DNA damage by activating the programmed cell death pathway (Miyai 2018), which may underpin some CEC loss observed in ECD. Evidence for mitochondrial dysfunction in FECD patient corneas is, however, somewhat indirect. Menadione, which generates free radicals in cells, thus mimicking the effects of electron leakage in mitochondrial function, induced mitochondrial DNA damage and cytochrome C release (an apoptotic marker) in immortalized CEC models (Halilovic et al. 2016). Cultured, immortalized CEC, derived from FECD patients had lower levels of mitochondria and reduced mitochondrial fitness (as measured by mitochondrial membrane potential) than cells derived from control subjects (Benischke et al. 2017). This would be consistent with oxidative damage culminating in reduced mitochondrial fitness. Potentially then, CEC from FECD patients are poised to undergo apoptosis upon additional oxidative damage. Consistent with this, a comparison of CEC from control and FECD patients revealed increased activation of mitophagy pathway in FECD (Miyai et al. 2019). An additional role for mitochondria in corneal dystrophies centers on SLC4A11 (Ogando et al. 2019). Corneal endothelial cells from slc4a11/ mice display elevations of superoxide, mitochondrial depolarization, and higher rates of apoptosis than controls (Ogando et al. 2019). This could explain the increased rate of cell loss in patients with SLC4A11 mutations. In this model (Fig. 7), SLC4A11 functions in both plasma membrane and inner mitochondrial membrane as a H+ channel, moving H+ from the extracellular medium into the mitochondrial matrix, following the plasma membrane and mitochondrial membrane potentials (Ogando et al. 2019). SLC4A11-mediated H+ flux in this manner would degrade the mitochondrial membrane potential and thus decrease ATP production by F0-F1-ATPase. In the context

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

H

61

+

Basolateral

H

+

Mitochondrion

Δψm

H

+

ROS

NH3 NH3

Gln Glu

α-KG

TCA

Fig. 7 Compromised mitochondrial function. CEC use glutaminolysis to produce much of their ATP: Glutamine (Gln) is deaminated to glutamic acid (Glu), which in turn is deaminated to alphaketoglutarate (α-KG), which feeds into the tricarboxylic acid cycle (TCA). Ammonia (NH3) stimulates reactive oxygen species (ROS) production by some component of the mitochondrial electron transport chain (MET), which can induce apoptotic cell death. SLC4A11 (green square) has been proposed to localize to the mitochondrial inner membrane, where it may mediate H+ influx that would slightly depolarize the mitochondrial membrane potential (ΔΨm), resulting in decreased ROS production (Ogando et al. 2019). SLC4A11 could also normally contribute to NH3 efflux from the inner mitochondrial membrane, or plasma membrane. SLC4A11 function may thus reduce ROS production either by reducing matrix NH3 levels or by slightly reducing ΔΨm. ROS levels thus may rise upon loss of SLC4A11 function, leading to apoptotic cell death

62

D. Malhotra and J. R. Casey

of CEC, however, the heavy use of glutaminolysis for energy production (Zhang et al. 2017) gives rise to mitochondrial hyperpolarization and high NH3 production. NH3 stimulates ROS production by some component of the mitochondrial electron transport chain (Murthy et al. 2001). SLC4A11 is proposed to serve a protective function by reducing ROS production (Ogando et al. 2019). This could occur either by reducing the mitochondrial membrane potential by SLC4A11-mediated transport of H+ into the mitochondrial matrix or by transport of NH3 out of the matrix. Indeed, ROS production is reduced in glutamine-exposed cells expressing SLC4A11 in comparison to cells lacking SLC4A11 (Ogando et al. 2019). Further, supporting a link of SLC4A11 to mitochondria, HEK293 cells expressing WT SLC4A11 were less prone to accumulate ROS and to undergo apoptosis than those expressing corneal dystrophy mutants of SLC4A11 (Roy et al. 2015). Finally, in immortalized cultured human CEC, knockdown of SLC4A11 expression led to increased apoptotic cell death (Liu et al. 2012). Taken together, loss of SLC4A11 function in FECD patients increases their CEC ROS load, which may lead to an increased rate of CEC loss via cell death. The data supporting the role of SLC4A11 in protecting against glutaminolysisassociated cell death are compelling (Ogando et al. 2019), but some caution needs to be exercised in details of the mitochondrial mechanism. Although evidence has been presented for an inner mitochondrial membrane localization for SLC4A11, none of the protein’s splicing variants contain an N-terminal mitochondrial targeting sequence (MTS) recognized by an online analysis (Fukasawa et al. 2015). Further, full sequence analysis (Kumar et al. 2018) does not predict a mitochondrial localization for SLC4A11. Lacking a recognizable N-terminal MTS, SLC4A11 would need to target to the inner mitochondrial membrane via the mitochondrial carrier family pathway, which has been observed for carrier proteins (SLC25) with six transmembrane segments (Ogunbona and Claypool 2019). SLC4A11 with 14 transmembrane segments (Badior et al. 2017) would be extraordinarily hydrophobic to be able to reach the inner membrane. Finally, SLC4A11 does not appear in the list of 1,098 proteins present in mitochondria (Pagliarini et al. 2008), from tissues including kidney where SLC4A11 is expressed (Groeger et al. 2010). Targeting of SLC4A11 to mitochondria may be signaled by high glutamine metabolism or by a CEC-specific factor, but the reported localization of the protein to IMM will require additional study. In addition, a role of SLC4A11 in preventing glutaminolysis-associated cell death could occur through mechanisms that do not require IMM localization. The crux of the glutaminolysis-cell death model is that mitochondrial matrix NH3 promotes ROS production. Removal of NH3 from the matrix would thus reduce ROS and reduce cell death. SLC4A11 acts as either an NH3 or NH3/H+ transporter (Loganathan et al. 2016; Zhang et al. 2015; Kao et al. 2019). Both the mitochondrial membrane potential and H+ gradient (matrix is more alkaline than cytosol) disfavor SLC4A11 transport of NH3/H+ out of the matrix. Similarly, SLC4A11 could not transport NH3/H+ out of the cell as the plasma membrane potential is negative. SLC4A11 does localize to the CEC plasma membrane, where it could facilitate NH3 efflux from the cell. In so doing, it would reduce the cytosolic NH3 concentration,

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

63

which would increase the magnitude of the NH3 concentration gradient from the mitochondrial matrix to the cytosol. Thus, observations of reduced glutaminolysisassociated cell death could also be explained by SLC4A11 plasma membrane NH3 transport function, in addition to the proposed role of SLC4A11 in H+ movement into the mitochondrial matrix (Ogando et al. 2019).

5.5

Epithelial-Mesenchymal Transition (EMT)

Cell polarity and tight cell–cell contacts are essential for proper functioning of the corneal endothelium. These epithelial cell features are lost when cells undergo EMT (Lavin and Tiwari 2020). Analysis of RNA from corneas affected by FECD revealed enrichment of genes promoting EMT (Cui et al. 2018). Further, ZEB1 promotes EMT possibly through its role in downregulating expression of E-cadherin (Iliff et al. 2012). Finally, TCF4 promotes EMT and MBNL1 splicing is important in EMT (Du et al. 2015; Iliff et al. 2012), suggesting that some part of TCF4 pathology arises through promotion of EMT.

5.6

Cell Adhesion Defects

Cell attachment to a basement membrane is a key feature of epithelia, like the corneal endothelium. Cell adhesion sites connect the inside of the cell to the extracellular environment and transduce signals to maintain a healthy physiology (Gumbiner 1996). Cell adhesion complexes include: CAM or receptors, ECM proteins, and the cytoplasmic/ancillary/peripheral membrane proteins. CAMs, which link cytosolic proteins and ECM, are integral membrane proteins in five families: integrins, immunoglobulin cell adhesion molecules (IgCAMs), selectins, cadherins, and proteoglycans (Murray et al. 1999). ECM proteins that constitute the basement membrane to which the cells adhere are usually secreted by these cells themselves. ECM proteins include collagens, laminins, fibronectins (FNCs), and proteoglycans that assemble into macromolecular arrays to create a defined basement membrane arrangement (Khalili and Ahmad 2015). Genes encoding ECM proteins are upregulated in corneas from FECD patients (Du et al. 2015). ECM proteins interact with their plasma membrane receptors to hold the cells tightly in place (Fig. 5). Cytosolic/peripheral membrane proteins connect cytosolic CAM regions to the cytoskeleton to transduce signals to the cytoplasm. Loss of epithelial cell anchorage from ECM triggers programmed cell death called anoikis, a physiological defense mechanism in multi-cellular organisms that prevents growth of detached cells on distal matrices (Gilmore 2005). Thus, cell death observed in corneal dystrophies could arise from defects in cell adhesion. Attachment of CEC to DM is critical to maintain the endothelial barrier and thus corneal transparency. Integrins in the corneal endothelium contribute to CEC-DM

64

D. Malhotra and J. R. Casey

attachment (Stepp 2006; Carter 2009; Goyer et al. 2018). Integrin αV (ITGAV), Integrin β3 (ITGB3), and integrin β5 (ITGB5), the corneal endothelial integrins, are heterodimers of αVβ3 and αVβ5. They interact with collagens, FNC, and laminins in DM by binding to their Arg-Gly-Asp (RGD) motifs (Stepp 2006). α4β1 and α6β1, discovered recently in CEC, interact with common basement membrane proteins through their large extracellular domains (Goyer et al. 2018; TCGA 2019; Hall et al. 1990; Mould et al. 1994). ECD patients show progressive CEC loss suggesting that defective cell adhesion could be one of the potential underlying causes of the disease (Fig. 8). In FECD, DM thickens and guttae formed from extracellular matrix secreted by CEC. This could be interpreted as an adaptive response to close gaps left by detached CEC to maintain endothelial integrity or that cells sense diminished adhesion to DM and attempt to compensate by secretion of DM components to provide a matrix for cell attachment. Indeed, CEC from FECD patients secrete higher levels of extracellular matrix proteins, including collagens (Xia et al. 2016) and fibronectin (Goyer et al. 2018). A prominent FECD gene is COL8A2 (Biswas et al. 2001), encoding a key component of DM. Normally, COL8A2 forms triple helices, which FECD point mutations are predicted to compromise (Moschos et al. 2019). Unfortunately, since COL8A2 is a large, post-translationally modified, helical, secreted protein, the effects of mutations have not been fully assessed. Col8a2L450W/L450W and col8a2Q455K/Q455K mice, however, exhibit CEC loss due to decrease in DM stiffness as a consequence of mis-organization of collagen fibrils (Leonard et al. 2019). Mis-organization of COL8A2 protein and biomechanical changes in DM could compromise CEC attachment, with consequences for cell survival, although there is no experimental evidence yet to support the idea. Interestingly, a second DM component, laminin γ1 chain (LAMC1), was identified in a GWAS as a FECD gene (Afshari et al. 2017). Specific LAMC1 mutations have not yet been reported, making assessment of the role of LAMC1 in corneal dystrophies somewhat speculative. A role of LAMC1 in cell adhesion has, however, been studied in the context of tumor cell adhesion to basement membranes (Sun et al. 2018). Interestingly, LAMC1 has been connected as an extracellular component of the PI3K/Akt signaling pathway (Sun et al. 2018), in which KANK4 acts as an intracellular effector to regulate actin filament polymerization (Kakinuma et al. 2009). KANK4 also was identified in a GWAS as an FECD gene (Afshari et al. 2017), thus implicating the cell adhesion-cytoskeletal axis as a potential disease mechanism in FECD. Both LAMC1 and KANK4 are associated with tumor phenotypes, promoting cell growth, but in the same vein defects in these proteins could adversely affect CEC growth leading to the increased rates of cell death found in FECD. Recent identification of COL17A1 mutation in FECD (Zhang et al. 2019) is interesting and points toward a cell adhesion defect. As a collagen, COL17A1 likely contributes to DM. The reported site of mutation is located on the COL17A1 surface (Zhang et al. 2019), where the defect is more likely to affect protein–protein contacts than protein folding. Other COL17A1 mutations cause epithelial recurrent erosion

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

65

Loss of Cell Adhesion

Healthy Endothelium

Adhesion Defect

DM Oversecretion

Folding Defect

Diseased Endothelium

Adhesion Mutants

Folding Mutants

COL8A2

COL8A1

Gutatta

Fig. 8 Defective SLC4A11-mediated cell adhesion in FECD and CHED. SLC4A11 contributes to CEC adhesion to DM components (Malhotra et al. 2019b). Mutations in SLC4A11 that affect its cell adhesion role (adhesion mutants) or protein folding that cause ER-retention (folding mutants) contribute to the compromised CEC-DM attachment. CEC, in a failed attempt to retain attachment to DM, may oversecrete DM components which leads to increases in DM thickness and continuous deposition of DM components can lead to guttae on the posterior surface of DM. Compromised SLC4A11 adhesion role, increase in DM thickness and guttae can eventually contribute to loss of CEC in diseased corneas. The severity of the effect of SLC4A11 mutations may contribute to increased rate of cell loss (as in CHED) or delayed CEC detachment along with other physiological factors (as in FECD) that affect CEC-DM physiology

dystrophy (MIM #122400), which could be explained by an epithelial cell adhesion defect. Recently, a role of SLC4A11 in CEC adhesion to DM has been described (Malhotra et al. 2019b). SLC4A11 is among the CEC’s most abundant proteins (Frausto et al. 2014; Chng et al. 2013) and localizes to the basolateral surface of CEC, facing DM (Vilas et al. 2013). Further, SLC4A1 (Band 3, AE1), a member of

66

D. Malhotra and J. R. Casey

the SLC4 family with SLC4A11, is an abundant erythrocyte membrane protein with a key role in attachment to cytoskeletal proteins in addition to its transport role (Cordat and Casey 2009), suggesting that SLC4 proteins may have membrane adhesion roles apart from transport. Through its large third extracellular loop (EL3), SLC4A11 promotes adhesion of cultured HEK293 cells and human CEC to culture dishes coated with DM extracts (Malhotra et al. 2019b). Transfection of HEK293 cells with SLC4A11 or αV and β3 integrins promoted adhesion to DM to the same degree. Further, an antibody against SLC4A11 EL3 blocked about 25% of cell adhesion to DM by primary cultures of human CEC (Malhotra et al. 2019b). Together, this suggests a significant role of SLC4A11 in maintaining CEC-DM attachment (Fig. 8). SLC4A11 potentially interacts with DM components, COL8A2 and COL8A1. Four FECD-causing mutations in SLC4A11-EL3 compromised the proteins’ ability to bind DM components, without affecting SLC4A11 cell surface trafficking or ability to facilitate a water flux (Malhotra et al. 2019b). Increased rates of cell loss found in FECD patients with SLC4A11 mutations may thus be explained by a reduced strength of CEC-DM adhesion, leading to cell death through anoikis. The majority of CHED-causing SLC4A11 mutations compromise cell surface abundance of SLC4A11 due to its misfolding and subsequent ER-retention (Alka and Casey 2018). CHED-causing SLC4A11 mutations also compromised SLC4A11mediated adhesion to DM (Malhotra et al. 2019b). Reduced abundance or absence of SLC4A11 at CEC surface may compromise adhesion, explaining onset of CHED during infancy. Transcriptional profiling of neural crest cells (NCC), the embryonic cells that migrate into the cleft between surface ectoderm and lens vesicle to form the endothelium, shows SLC4A11 expression (Lumb et al. 2017). CHED may be a developmental disorder due to defective NCC adhesion during migration absence of SLC4A11. Indeed, CHED patients are born with a tenfold lower CEC density compared to healthy infants further consistent with prenatal defects (Sultana et al. 2007; Kenyon and Antine 1971). This strongly suggests cell adhesion as a defective mechanism in FECD and CHED that can explain CEC loss in patients.

5.7

ER Stress and the Unfolded Protein Response

A consequence of mutations can be defective folding of the corresponding mutated protein. Proteins synthesized in association with the endoplasmic reticulum (ER), integral membrane proteins, and secretory proteins, may be retained in the ER upon misfolding. Prolonged ER retention provides time for misfolded proteins to achieve a native structure. Ultimately, if ER-associated proteins fail to fold, they can accumulate or be targeted for degradation. Accumulation of misfolded proteins in ER can induce the unfolded protein response (UPR) in cells, which can lead to apoptosis (Almanza et al. 2019). Through this mechanism, mutations of corneal dystrophy genes could induce CEC loss. Indeed, increased levels of apoptotic cell death are a feature of FECD endothelium (Borderie et al. 2000) and markers of UPR are elevated in FECD patient corneas (Engler et al. 2010). Thus, corneal dystrophy

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

67

gene products that pass through the endoplasmic reticulum (membrane proteins or secretory proteins) are candidates for causing disease through UPR/apoptosis. These include: LAMC1, ATPB1, SLC4A11, and COL8A2. ATPB1 and LAMC1 were identified in a GWAS screen as FECD loci (Afshari et al. 2017) but point mutants have not been found in the genes. So, it is impossible to assess their potential pathogenic mechanism of their mutations. DM component, COL8A2, is secreted by CEC and its mutations cause some FECD cases (Riazuddin et al. 2009). COL8A2 point mutations have been proposed to cause the protein to accumulate in ER prior to secretion, leading to UPR-associated cell death (Moschos et al. 2019). Most significantly, a transgenic knock-in mouse of the col8a2 p.Gln455Lys FECD allele revealed ER-swelling and increased apoptotic markers (Jun et al. 2012). This provides strong evidence for UPR-induced apoptosis as the disease mechanism in the case of one COL8A2 FECD mutation. LOXHD1 does not have any predicted transmembrane segments and the wildtype protein is cytosolic (Riazuddin et al. 2012), meaning that LOXHD1 does not pass through the ER. Fifteen missense mutations have been identified in LOXHD1, localizing at the protein surface, suggesting a role in protein–protein interactions. Analysis of the explanted cornea of an individual with an p.Arg547Cys mutation revealed increased levels of LOXHD1 expression and evidence of stained protein aggregates (Riazuddin et al. 2012). This would be consistent with protein misfolding leading to LOXHD1 accumulation. Since LOXHD1 is a polyvalent protein binder, with 15 PLAT domains, mutant LOXHD1 could potentially cause aberrant clustering of its client membrane proteins, potentially causing them to accumulate in the ER, leading to UPR. LOXHD1 is associated with late-onset FECD, suggesting that the effect of LOXHD1 results in a slight increase in cell death and late-onset disease. Extensive studies of SLC4A11 point mutants revealed folding defects in 20% of FECD-causing mutants and 61% of CHED-causing mutants (Alka and Casey 2018). SLC4A11 would thus appear to be a candidate to cause disease through the UPR/apoptotic route. The case for increased cell death arising from UPR is not, however, strong. Indeed, knock-down of SLC4A11 expression in human corneal endothelial cells increases apoptotic cell death (Liu et al. 2012), suggesting that SLC4A11 prevents cell death. Expression of misfolded, intracellular-retained SLC4A11 mutants was associated with increased apoptotic cell death, but via increased sensitivity to oxidative stress (Roy et al. 2015). Arguing against increased cell death associated with expression of SLC4A11 mutants, no increase of apoptotic or necrotic markers was observed in cells expressing misfolded SLC4A11 in a cell culture model (Loganathan and Casey 2014). Recent data suggesting that SLC4A11 localizes to the mitochondrial inner membrane (Ogando et al. 2019) lead to the possibility that misfolded SLC4A11 induces cell death via the mitochondrial precursor overaccumulation stress pathway (Coyne and Chen 2019). The observation that slc4a11/ mice manifest corneal dystrophy symptoms (Vilas et al. 2013; Groeger et al. 2010; Lopez et al. 2009), however, strongly suggests that corneal dystrophy arises from loss of SLC4A11 function rather than a toxic gain of function, as would be the case if SLC4A11 mutants induce UPR/apoptosis.

68

D. Malhotra and J. R. Casey

5.8

Water Pump Defects

The principal symptom for corneal dystrophy patients is corneal haziness, secondary to corneal edema. Disease mechanisms increasing rates of CEC loss will impair the endothelial water pump, as a result of increased water leakage from impaired continuity of the endothelial layer and because fewer cells will be available to pump fluid out of the stroma. Pleomorphism and polymegathism also reduce the ability of the endothelium to deturgesce the cornea through endothelial pump (DelMonte and Kim 2011; Feizi 2018). In addition, defects in the molecular components of the corneal fluid pump may explain corneal dystrophy in some patients. The endothelial water pump mechanism has been well elaborated (Bonanno 2012) (Fig. 3). While the components of the pump are generally agreed upon, how they work together remains uncertain. The corneal edema observed upon treatment with the Na+/K+-ATPase inhibitor, ouabain, argues persuasively for a central role of this protein (Fig. 9) (Geroski et al. 1984). It also argues that FECD associated with ATPB1 defects are affected by the loss of the protein’s function. While AQP1, found

Stromal Edema Fluid accumulation

Distorted lamellae arrangement

H2O H+

NBCe1

NHE1

SLC4A11

MCT1

Kir2.1

NKCC1

Osmotic Leak

Na-K ATPase 2K+

TJ

TJ

H 2O

CACC1

CFTR

AQP1

Endothelial Pump

3Na+

AE2

MCT4

Aqueous Humor

Fig. 9 Endothelial pump defects in FECD and CHED. An array of proteins contributes to corneal endothelial pump (Fig. 3), which help in maintaining corneal transparency. Defective fluid transport due to SLC4A11 mutations affects the movement of excessive fluid from stroma to aqueous humor via AQP1. Accumulation of fluid in stroma presents as edema while it also distorts the lamellar organization which further compromises visual clarity in FECD and CHED. ATP1B1, a subunit of sodium–potassium ATPase which also contributes to endothelial pump, has also been linked to FECD in GWAS. However, no mutations have been reported that affect the transport role of the protein in FECD or CHED

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

69

in apical CEC surface, is the only AQP protein of CEC (Vilas et al. 2013; Hamann et al. 1998), its role in maintenance of stromal deturgescence appears dispensable as aqp1/ mice have thinned corneal stroma (Thiagarajah and Verkman 2002). Complicating this interpretation, these mice also manifest slowed recovery from corneal swelling induced by hypoosmotic challenge at the corneal epithelial surface (Thiagarajah and Verkman 2002), which suggests a role of AQP1 in recovery from edema. The strongest argument that AQP1 is not essential for water pump function is the lack of corneal phenotype in patients lacking AQP1 in spite of their profound urinary concentrating defect (King et al. 2001). Amongst the corneal dystrophy genes, only two stand out as potentially involved in the endothelial pump: SLC4A11 and ATP1B1. ATP1B1 encodes the beta subunit of the Na+/K+-ATPase. As discussed above, a critical role of Na+/K+-ATPase in maintenance of stromal fluid balance is established, so defects in the protein could readily explain corneal symptoms. Thus far, however, ATP1B1 has only been identified as an FECD gene in a GWAS, which identified a single nucleotide polymorphism (SNPEffect 2019), rs1200114, in an intergenic region near ATP1B1 (Afshari et al. 2017). Since the other gene flanking the SNP is not expressed in cornea, the linkage implies involvement of ATP1B1. No mutations have yet been identified in the ATP1B1 gene of corneal dystrophy patients. Given the critical role of Na+/K+-ATPase in maintaining plasma membrane potential, and Na+, and K+ gradients, mutations profoundly altering the protein’s function would likely manifest in many tissues beyond cornea. SLC4A11 contributes to maintenance of stromal deturgescence by two mechanisms: (1) It provides a water flux pathway from stroma back into endothelial cells and (2) SLC4A11’s H+ transport function has been proposed to have a critical role in the endothelial water pump by countering the alkalinizing action of HCO3 accumulation (Myers et al. 2016). The only characterized mutation that compromises the water flux function without affecting cell surface abundance is p.Arg125His (Vilas et al. 2013). Other SLC4A11 mutations that affect protein trafficking will compromise cell surface abundance of SLC4A11, which in turn will compromise the water flux function in plasma membrane (Loganathan and Casey 2014). Thirty out of 55 missense mutations compromise cell surface trafficking of the protein (Alka and Casey 2018; Loganathan and Casey 2014; Chiu et al. 2015). SLC4A11 defects can compromise the endothelial pump, resulting in corneal edema. The accumulated fluid distorts the arrangement of stromal collagen lamellae resulting in increased scattering of light and compromised visual acuity. Indeed, slc4a11/ mice develop corneal edema and increased central corneal thickness, resembling the clinical presentations of ECD in patients (Vilas et al. 2013; Han et al. 2013). Recent recognition that MPDZ mutations may be responsible for CHED when homozygous and FECD when heterozygous (Moazzeni et al. 2019) suggests an additional component of the endothelial water pump that may cause FECD and CHED. MPDZ mutations cause hydrocephalus (Feldner et al. 2017). In turn hydrocephalus arises from over-accumulation of cerebrospinal fluid, whose production is controlled by the choroid plexus, an endothelium with some similarities to the corneal endothelium. Careful analysis of mpdz-null mice revealed loss of integrity

70

D. Malhotra and J. R. Casey

of the connections between the cells of the choroid plexus (Yang et al. 2019), giving rise to fluid leak between these cells. If a similar phenomenon happens in the corneal endothelium, then increased leakage of fluid from aqueous humor into stroma would counter the normal pumping action of the corneal endothelial cells. In spite of effective corneal endothelial water pump, stromal edema would ensue because the pump could not overcome the rate of fluid leakage through this paracellular pathway.

5.9

Altered Gene Expression

Products of some ECD-causing genes modulate the expression of other ECD genes. ZEB1 is a transcriptional activator and repressor of several genes of Wnt and TCF4/β-catenin pathway (Aigner et al. 2007; Sanchez-Tillo et al. 2011). ZEB1 binds to the SLC4A11 promoter at its transcription start site to modulate its expression (Zhang and Aldave 2018). TCF4 interacts with ZEB1, converting ZEB1 from a transcriptional repressor to an activator (Sanchez-Tillo et al. 2015). The association between TCF4 and ZEB1 suggests that TCF4 may regulate SLC4A11 expression. ZEB1 acts as a transcription suppressor in the absence of TCF4, suggesting that SLC4A11 may be downregulated under reduced TCF4 levels. Interestingly, SAGE analysis from FECD corneas (which were probably defective in TCF4, since 70% of FECD arises from TCF4 mutations) revealed downregulated SLC4A11 as a common FECD feature (Gottsch et al. 2003). Other CAMs, including NRXN1 (Forrest et al. 2012), CNTNAP2 (Forrest et al. 2012), and integrins (Plantefaber and Hynes 1989) are also regulated by TCF4. TCF4 also binds to non-coding regulatory regions of COL8A2, KANK4, ATP1B1, and LAMC1, suggesting downstream transcriptional regulation (Zhang and Aldave 2018). Reduced ATP1B1 RNA levels are found in FECD corneas (Jalimarada et al. 2013). ECD-causing ZEB1 mutation also leads to downregulation of COL8A2 (Lechner et al. 2013), which further suggests transcriptional regulation of other ECD genes by TCF4 and ZEB1. Together, these data suggest that aberrant gene expression caused by TCF4 and TCF8 deficits could underlay some pathophysiology of corneal dystrophies.

6 Conclusions Endothelial corneal dystrophies are complex, arising from a combination of environmental and genetic factors. All are marked by a decreased density of CEC, giving rise to defective cornea stromal deturgescence, which in turn disrupts collagen fibril packing and greatly impairs vision. Since CEC cannot proliferate, our CEC density can only decline through life. Factors that increase the rate of cell loss will hasten the onset of ECD, marked by the stage when the endothelium can no longer maintain the required level of stromal dehydration. Thus, environmental factors increasing cell

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

71

death, including oxidative stress, interplay with genetic predispositions, culminating in either early onset CHED or requiring decades to ensue in the case of FECD. Endothelial dystrophy genes point toward molecular mechanisms of disease that include RNA toxicity associated with deficiencies in the MBNL1 RNA processing pathway, oxidative stress, mitochondrial dysfunction, the unfolded protein response/ apoptosis associated with misfolded proteins, ammonia toxicity, cell adhesion defects leading to anoikis and defects in the endothelial water pump. As outlined in the review, some genes may cause CHED or FECD through more than one of these mechanisms. Acknowledgments We thank the many researchers who have contributed to this rich field and apologize to those whose work we were unable to cite. We thank Dr. Mathieu Thériault, University of Laval, for providing images used in this manuscript. Canadian Institutes of Health Research supports research in the JRC laboratory.

References Abdellah MM, Ammar HG, Anbar M, Mostafa EM, Farouk MM, Sayed K, Alsmman AH, Elghobaier MG (2019) Corneal endothelial cell density and morphology in healthy egyptian eyes. J Ophthalmol 2019:6370241. https://doi.org/10.1155/2019/6370241 Afshari NA, Igo RP Jr, Morris NJ, Stambolian D, Sharma S, Pulagam VL, Dunn S, Stamler JF, Truitt BJ, Rimmler J, Kuot A, Croasdale CR, Qin X, Burdon KP, Riazuddin SA, Mills R, Klebe S, Minear MA, Zhao J, Balajonda E, Rosenwasser GO, Baratz KH, Mootha VV, Patel SV, Gregory SG, Bailey-Wilson JE, Price MO, Price FW Jr, Craig JE, Fingert JH, Gottsch JD, Aldave AJ, Klintworth GK, Lass JH, Li YJ, Iyengar SK (2017) Genome-wide association study identifies three novel loci in Fuchs endothelial corneal dystrophy. Nat Commun 8:14898. https://doi.org/10.1038/ncomms14898 Aigner K, Dampier B, Descovich L, Mikula M, Sultan A, Schreiber M, Mikulits W, Brabletz T, Strand D, Obrist P, Sommergruber W, Schweifer N, Wernitznig A, Beug H, Foisner R, Eger A (2007) The transcription factor ZEB1 (deltaEF1) promotes tumour cell dedifferentiation by repressing master regulators of epithelial polarity. Oncogene 26(49):6979–6988. https://doi.org/ 10.1038/sj.onc.1210508 Alka K, Casey JR (2018) Molecular phenotype of SLC4A11 missense mutants: setting the stage for personalized medicine in corneal dystrophies. Hum Mutat 39(5):676–690. https://doi.org/10. 1002/humu.23401 Almanza A, Carlesso A, Chintha C, Creedican S, Doultsinos D, Leuzzi B, Luis A, McCarthy N, Montibeller L, More S, Papaioannou A, Puschel F, Sassano ML, Skoko J, Agostinis P, de Belleroche J, Eriksson LA, Fulda S, Gorman AM, Healy S, Kozlov A, Munoz-Pinedo C, Rehm M, Chevet E, Samali A (2019) Endoplasmic reticulum stress signalling - from basic mechanisms to clinical applications. FEBS J 286(2):241–278. https://doi.org/10.1111/febs. 14608 Angmo D, Selvan H, Behera AK, Suman PK (2018) Unilateral corneal edema in young: a diagnostic dilemma. Indian J Ophthalmol 66(11):1612–1614. https://doi.org/10.4103/ijo.IJO_ 564_18 Arici C, Arslan OS, Dikkaya F (2014) Corneal endothelial cell density and morphology in healthy Turkish eyes. J Ophthalmol 2014:852624. https://doi.org/10.1155/2014/852624

72

D. Malhotra and J. R. Casey

Badior KE, Alka K, Casey JR (2017) SLC4A11 three-dimensional homology model rationalizes corneal dystrophy-causing mutations. Hum Mutat 38(3):279–288. https://doi.org/10.1002/ humu.23152 Bai X, Zhang C, Zhang F, Xiao Y, Jin Y, Wang H, Xu L (2020) Five novel mutations in LOXHD1 gene were identified to cause autosomal recessive nonsyndromic hearing loss in four Chinese families. Biomed Res Int 2020:1685974. https://doi.org/10.1155/2020/1685974 Baratz KH, Tosakulwong N, Ryu E, Brown WL, Branham K, Chen W, Tran KD, Schmid-Kubista KE, Heckenlively JR, Swaroop A, Abecasis G, Bailey KR, Edwards AO (2010) E2-2 protein and Fuchs's corneal dystrophy. N Engl J Med 363(11):1016–1024. https://doi.org/10.1056/ NEJMoa1007064 Barry PA, Petroll WM, Andrews PM, Cavanagh HD, Jester JV (1995) The spatial organization of corneal endothelial cytoskeletal proteins and their relationship to the apical junctional complex. Invest Ophthalmol Vis Sci 36(6):1115–1124 Benischke AS, Vasanth S, Miyai T, Katikireddy KR, White T, Chen Y, Halilovic A, Price M, Price F Jr, Liton PB, Jurkunas UV (2017) Activation of mitophagy leads to decline in Mfn2 and loss of mitochondrial mass in Fuchs endothelial corneal dystrophy. Sci Rep 7(1):6656. https://doi. org/10.1038/s41598-017-06523-2 Biswas S, Munier FL, Yardley J, Hart-Holden N, Perveen R, Cousin P, Sutphin JE, Noble B, Batterbury M, Kielty C, Hackett A, Bonshek R, Ridgway A, McLeod D, Sheffield VC, Stone EM, Schorderet DF, Black GC (2001) Missense mutations in COL8A2, the gene encoding the alpha2 chain of type VIII collagen, cause two forms of corneal endothelial dystrophy. Hum Mol Genet 10(21):2415–2423 Bitar MS, Liu C, Ziaei A, Chen Y, Schmedt T, Jurkunas UV (2012) Decline in DJ-1 and decreased nuclear translocation of Nrf2 in Fuchs endothelial corneal dystrophy. Invest Ophthalmol Vis Sci 53(9):5806–5813. https://doi.org/10.1167/iovs.12-10119 Bonanno JA (2003) Identity and regulation of ion transport mechanisms in the corneal endothelium. Prog Retin Eye Res 22(1):69–94 Bonanno JA (2012) Molecular mechanisms underlying the corneal endothelial pump. Exp Eye Res 95(1):2–7. https://doi.org/10.1016/j.exer.2011.06.004 Borderie VM, Baudrimont M, Vallee A, Ereau TL, Gray F, Laroche L (2000) Corneal endothelial cell apoptosis in patients with Fuchs’ dystrophy. Invest Ophthalmol Vis Sci 41(9):2501–2505 Bourne WM (2003) Biology of the corneal endothelium in health and disease. Eye (Lond) 17 (8):912–918. https://doi.org/10.1038/sj.eye.6700559 Brejchova K, Dudakova L, Skalicka P, Dobrovolny R, Masek P, Putzova M, Moosajee M, Tuft SJ, Davidson AE, Liskova P (2019) IPSC-derived corneal endothelial-like cells act as an appropriate model system to assess the impact of SLC4A11 variants on pre-mRNA splicing. Invest Ophthalmol Vis Sci 60(8):3084–3090. https://doi.org/10.1167/iovs.19-26930 Carter RT (2009) The role of integrins in corneal wound healing. Vet Ophthalmol 12(Suppl 1):2–9. https://doi.org/10.1111/j.1463-5224.2009.00726.x Chaurasia S, Ramappa M, Annapurna M, Kannabiran C (2020) Coexistence of congenital hereditary endothelial dystrophy and fuchs endothelial corneal dystrophy associated with SLC4A11 mutations in affected families. Cornea 39(3):354–357. https://doi.org/10.1097/ico. 0000000000002183 Chiu AM, Mandziuk JJ, Loganathan SK, Alka K, Casey JR (2015) High throughput assay identifies glafenine as a corrector for the folding defect in corneal dystrophy-causing mutants of SLC4A11. Invest Ophthalmol Vis Sci 56(13):7739–7753. https://doi.org/10.1167/iovs.1517802 Chng Z, Peh GS, Herath WB, Cheng TY, Ang HP, Toh KP, Robson P, Mehta JS, Colman A (2013) High throughput gene expression analysis identifies reliable expression markers of human corneal endothelial cells. PLoS One 8(7):e67546. https://doi.org/10.1371/journal.pone.0067546

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

73

Cordat E, Casey JR (2009) Bicarbonate transport in cell physiology and disease. Biochem J 417 (2):423–439. https://doi.org/10.1042/BJ20081634 Coyne LP, Chen XJ (2019) Consequences of inner mitochondrial membrane protein misfolding. Mitochondrion 49:46–55. https://doi.org/10.1016/j.mito.2019.06.001 Cui Z, Zeng Q, Guo Y, Liu S, Wang P, Xie M, Chen J (2018) Pathological molecular mechanism of symptomatic late-onset Fuchs endothelial corneal dystrophy by bioinformatic analysis. PLoS One 13(5):e0197750. https://doi.org/10.1371/journal.pone.0197750 DelMonte DW, Kim T (2011) Anatomy and physiology of the cornea. J Cataract Refract Surg 37 (3):588–598. https://doi.org/10.1016/j.jcrs.2010.12.037 Desir J, Abramowicz M (2008) Congenital hereditary endothelial dystrophy with progressive sensorineural deafness (Harboyan syndrome). Orphanet J Rare Dis 3:28. https://doi.org/10. 1186/1750-1172-3-28 Du J, Aleff RA, Soragni E, Kalari K, Nie J, Tang X, Davila J, Kocher JP, Patel SV, Gottesfeld JM, Baratz KH, Wieben ED (2015) RNA toxicity and missplicing in the common eye disease fuchs endothelial corneal dystrophy. J Biol Chem 290(10):5979–5990. https://doi.org/10.1074/jbc. M114.621607 Duman R, Tok Cevik M, Gorkem Cevik S, Duman R, Perente I (2016) Corneal endothelial cell density in healthy Caucasian population. Saudi J Ophthalmol 30(4):236–239. https://doi.org/10. 1016/j.sjopt.2016.10.003 EBAA (2017) 2016 Eye bank statistical report. Eye bank Association of America, Washington Eghrari AO, Gottsch JD (2010) Fuchs' corneal dystrophy. Expert Rev Ophthalmol 5(2):147–159. https://doi.org/10.1586/eop.10.8 Eghrari AO, Riazuddin SA, Gottsch JD (2015) Overview of the cornea: structure, function, and development. Prog Mol Biol Transl Sci 134:7–23. https://doi.org/10.1016/bs.pmbts.2015.04. 001 Elhalis H, Azizi B, Jurkunas UV (2010) Fuchs endothelial corneal dystrophy. Ocul Surf 8 (4):173–184 Engler C, Kelliher C, Spitze AR, Speck CL, Eberhart CG, Jun AS (2010) Unfolded protein response in fuchs endothelial corneal dystrophy: a unifying pathogenic pathway? Am J Ophthalmol 149 (2):194–202.e192. https://doi.org/10.1016/j.ajo.2009.09.009 Ewete T, Ani EU, Alabi AS (2016) Normal corneal endothelial cell density in Nigerians. Clin Ophthalmol 10:497–501. https://doi.org/10.2147/OPTH.S97070 Feizi S (2018) Corneal endothelial cell dysfunction: etiologies and management. Ther Adv Ophthalmol 10:2515841418815802. https://doi.org/10.1177/2515841418815802 Feldner A, Adam MG, Tetzlaff F, Moll I, Komljenovic D, Sahm F, Bäuerle T, Ishikawa H, Schroten H, Korff T, Hofmann I, Wolburg H, von Deimling A, Fischer A (2017) Loss of Mpdz impairs ependymal cell integrity leading to perinatal-onset hydrocephalus in mice. EMBO Mol Med 9(7):890–905. https://doi.org/10.15252/emmm.201606430 Fischbarg J (2012) Water channels and their roles in some ocular tissues. Mol Asp Med 33 (5–6):638–641. https://doi.org/10.1016/j.mam.2012.07.016 Forrest M, Chapman RM, Doyle AM, Tinsley CL, Waite A, Blake DJ (2012) Functional analysis of TCF4 missense mutations that cause Pitt-Hopkins syndrome. Hum Mutat 33(12):1676–1686. https://doi.org/10.1002/humu.22160 Forrest MP, Waite AJ, Martin-Rendon E, Blake DJ (2013) Knockdown of human TCF4 affects multiple signaling pathways involved in cell survival, epithelial to mesenchymal transition and neuronal differentiation. PLoS One 8(8):e73169. https://doi.org/10.1371/journal.pone.0073169 Frausto RF, Wang C, Aldave AJ (2014) Transcriptome analysis of the human corneal endothelium. Invest Ophthalmol Vis Sci 55(12):7821–7830. https://doi.org/10.1167/iovs.14-15021 Fukasawa Y, Tsuji J, Fu SC, Tomii K, Horton P, Imai K (2015) MitoFates: improved prediction of mitochondrial targeting sequences and their cleavage sites. Mol Cell Proteomics 14 (4):1113–1126. https://doi.org/10.1074/mcp.M114.043083

74

D. Malhotra and J. R. Casey

Geroski DH, Kies JC, Edelhauser HF (1984) The effects of ouabain on endothelial function in human and rabbit corneas. Curr Eye Res 3(2):331–338. https://doi.org/10.3109/ 02713688408997217 Gilmore AP (2005) Anoikis. Cell Death Differ 12(Suppl 2):1473–1477. https://doi.org/10.1038/sj. cdd.4401723 Gottsch JD, Bowers AL, Margulies EH, Seitzman GD, Kim SW, Saha S, Jun AS, Stark WJ, Liu SH (2003) Serial analysis of gene expression in the corneal endothelium of Fuchs’ dystrophy. Invest Ophthalmol Vis Sci 44(2):594–599 Gottsch JD, Zhang C, Sundin OH, Bell WR, Stark WJ, Green WR (2005) Fuchs corneal dystrophy: aberrant collagen distribution in an L450W mutant of the COL8A2 gene. Invest Ophthalmol Vis Sci 46(12):4504–4511. https://doi.org/10.1167/iovs.05-0497 Goyer B, Theriault M, Gendron SP, Brunette I, Rochette PJ, Proulx S (2018) Extracellular matrix and integrin expression profiles in Fuchs endothelial corneal dystrophy cells and tissue model. Tissue Eng Part A 24(7–8):607–615. https://doi.org/10.1089/ten.TEA.2017.0128 Greenhill NS, Ruger BM, Hasan Q, Davis PF (2000) The alpha1(VIII) and alpha2(VIII) collagen chains form two distinct homotrimeric proteins in vivo. Matrix Biol 19(1):19–28 Grillet N, Schwander M, Hildebrand MS, Sczaniecka A, Kolatkar A, Velasco J, Webster JA, Kahrizi K, Najmabadi H, Kimberling WJ, Stephan D, Bahlo M, Wiltshire T, Tarantino LM, Kuhn P, Smith RJ, Muller U (2009) Mutations in LOXHD1, an evolutionarily conserved stereociliary protein, disrupt hair cell function in mice and cause progressive hearing loss in humans. Am J Hum Genet 85(3):328–337. https://doi.org/10.1016/j.ajhg.2009.07.017 Groeger N, Froehlich H, Maier H, Olbrich A, Kostin S, Braun T, Boettger T (2010) SLC4A11 prevents osmotic imbalance leading to corneal endothelial dystrophy, deafness, and polyuria. J Biol Chem 285:14467–14474. https://doi.org/10.1074/jbc.M109.094680 Guha S, Chaurasia S, Ramachandran C, Roy S (2017) SLC4A11 depletion impairs NRF2 mediated antioxidant signaling and increases reactive oxygen species in human corneal endothelial cells during oxidative stress. Sci Rep 7(1):4074. https://doi.org/10.1038/s41598-017-03654-4 Gumbiner BM (1996) Cell adhesion: the molecular basis of tissue architecture and morphogenesis. Cell 84(3):345–357 Halilovic A, Schmedt T, Benischke AS, Hamill C, Chen Y, Santos JH, Jurkunas UV (2016) Menadione-induced DNA damage leads to mitochondrial dysfunction and fragmentation during rosette formation in Fuchs endothelial corneal dystrophy. Antioxid Redox Signal 24 (18):1072–1083. https://doi.org/10.1089/ars.2015.6532 Hall DE, Reichardt LF, Crowley E, Holley B, Moezzi H, Sonnenberg A, Damsky CH (1990) The alpha 1/beta 1 and alpha 6/beta 1 integrin heterodimers mediate cell attachment to distinct sites on laminin. J Cell Biol 110(6):2175–2184. https://doi.org/10.1083/jcb.110.6.2175 Hamann S, Zeuthen T, La Cour M, Nagelhus EA, Ottersen OP, Agre P, Nielsen S (1998) Aquaporins in complex tissues: distribution of aquaporins 1-5 in human and rat eye. Am J Phys 274(5):C1332–C1345. https://doi.org/10.1152/ajpcell.1998.274.5.C1332 Han SB, Ang HP, Poh R, Chaurasia SS, Peh G, Liu J, Tan DT, Vithana EN, Mehta JS (2013) Mice with a targeted disruption of slc4a11 model the progressive corneal changes of congenital hereditary endothelial dystrophy. Invest Ophthalmol Vis Sci 54(9):6179–6189. https://doi.org/ 10.1167/iovs.13-12089 Hara S, Tsujikawa M, Kawasaki S, Nishida K (2019) Homeostasis of SLC4A11 protein is mediated by endoplasmic reticulum-associated degradation. Exp Eye Res 188:107782. https://doi.org/10. 1016/j.exer.2019.107782 Iliff BW, Riazuddin SA, Gottsch JD (2012) The genetics of Fuchs’ corneal dystrophy. Expert Rev Ophthalmol 7(4):363–375. https://doi.org/10.1586/eop.12.39 Iovino C, Fossarello M, Giannaccare G, Pellegrini M, Braghiroli M, Demarinis G, Napoli PE (2018) Corneal endothelium features in Fuchs' endothelial corneal dystrophy: a preliminary 3D anterior segment optical coherence tomography study. PLoS One 13(11):e0207891. https://doi. org/10.1371/journal.pone.0207891

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

75

Jalimarada SS, Ogando DG, Vithana EN, Bonanno JA (2013) Ion transport function of SLC4A11 in corneal endothelium. Invest Ophthalmol Vis Sci 54(6):4330–4340. https://doi.org/10.1167/ iovs.13-11929 Johnson DH, Bourne WM, Campbell RJ (1982) The ultrastructure of Descemet’s membrane. I. Changes with age in normal corneas. Arch Ophthalmol 100(12):1942–1947 Joyce NC (2012) Proliferative capacity of corneal endothelial cells. Exp Eye Res 95(1):16–23. https://doi.org/10.1016/j.exer.2011.08.014 Jun AS, Meng H, Ramanan N, Matthaei M, Chakravarti S, Bonshek R, Black GC, Grebe R, Kimos M (2012) An alpha 2 collagen VIII transgenic knock-in mouse model of Fuchs endothelial corneal dystrophy shows early endothelial cell unfolded protein response and apoptosis. Hum Mol Genet 21(2):384–393. https://doi.org/10.1093/hmg/ddr473 Jurkunas UV (2018) Fuchs endothelial corneal dystrophy through the prism of oxidative stress. Cornea 37(Suppl 1):S50–s54. https://doi.org/10.1097/ico.0000000000001775 Jurkunas UV, Bitar MS, Funaki T, Azizi B (2010) Evidence of oxidative stress in the pathogenesis of fuchs endothelial corneal dystrophy. Am J Pathol 177(5):2278–2289. https://doi.org/10. 2353/ajpath.2010.100279 Kakinuma N, Zhu Y, Wang Y, Roy BC, Kiyama R (2009) Kank proteins: structure, functions and diseases. Cell Mol Life Sci 66(16):2651–2659. https://doi.org/10.1007/s00018-009-0038-y Kao L, Azimov R, Abuladze N, Newman D, Kurtz I (2015) Human SLC4A11-C functions as a DIDS-stimulatable H+(OH) permeation pathway: partial correction of R109H mutant transport. Am J Physiol Cell Physiol 308(2):C176–C188. https://doi.org/10.1152/ajpcell.00271.2014 Kao L, Azimov R, Shao XM, Abuladze N, Newman D, Zhekova H, Noskov S, Pushkin A, Kurtz I (2019) SLC4A11 function: evidence for H+(OH) and NH3-H+ transport. Am J Physiol Cell Physiol. https://doi.org/10.1152/ajpcell.00425.2019 Kao L, Azimov R, Shao XM, Abuladze N, Newman D, Zhekova H, Noskov S, Pushkin A, Kurtz I (2020) SLC4A11 function: evidence for H+(OH) and NH3-H+ transport. Am J Physiol Cell Physiol 318(2):C392–c405. https://doi.org/10.1152/ajpcell.00425.2019 Kapoor R, Bornstein P, Sage EH (1986) Type VIII collagen from bovine Descemet’s membrane: structural characterization of a triple-helical domain. Biochemistry 25(13):3930–3937 Kensler TW, Wakabayashi N, Biswal S (2007) Cell survival responses to environmental stresses via the Keap1-Nrf2-ARE pathway. Annu Rev Pharmacol Toxicol 47:89–116. https://doi.org/10. 1146/annurev.pharmtox.46.120604.141046 Kenyon KR, Antine B (1971) The pathogenesis of congenital hereditary endothelial dystrophy of the cornea. Am J Ophthalmol 72(4):787–795 Khalili AA, Ahmad MR (2015) A review of cell adhesion studies for biomedical and biological applications. Int J Mol Sci 16(8):18149–18184. https://doi.org/10.3390/ijms160818149 Kim JH, Ko JM, Tchah H (2015) Fuchs endothelial corneal dystrophy in a heterozygous carrier of congenital hereditary endothelial dystrophy type 2 with a novel mutation in SLC4A11. Ophthalmic Genet 36(3):284–286. https://doi.org/10.3109/13816810.2014.881510 King LS, Choi M, Fernandez PC, Cartron JP, Agre P (2001) Defective urinary concentrating ability due to a complete deficiency of aquaporin-1. N Engl J Med 345(3):175–179. https://doi.org/10. 1056/nejm200107193450304 Klintworth GK (2009) Corneal dystrophies. Orphanet J Rare Dis 4:7. https://doi.org/10.1186/17501172-4-7 Krafchak CM, Pawar H, Moroi SE, Sugar A, Lichter PR, Mackey DA, Mian S, Nairus T, Elner V, Schteingart MT, Downs CA, Kijek TG, Johnson JM, Trager EH, Rozsa FW, Mandal MN, Epstein MP, Vollrath D, Ayyagari R, Boehnke M, Richards JE (2005) Mutations in TCF8 cause posterior polymorphous corneal dystrophy and ectopic expression of COL4A3 by corneal endothelial cells. Am J Hum Genet 77(5):694–708. https://doi.org/10.1086/497348 Kumar R, Kumari B, Kumar M (2018) Proteome-wide prediction and annotation of mitochondrial and sub-mitochondrial proteins by incorporating domain information. Mitochondrion 42:11–22. https://doi.org/10.1016/j.mito.2017.10.004

76

D. Malhotra and J. R. Casey

Labetoulle M, Baudouin C, Calonge M, Merayo-Lloves J, Boboridis KG, Akova YA, Aragona P, Geerling G, Messmer EM, Benítez-Del-Castillo J (2019) Role of corneal nerves in ocular surface homeostasis and disease. Acta Ophthalmol 97(2):137–145. https://doi.org/10.1111/ aos.13844 Lavin DP, Tiwari VK (2020) Unresolved complexity in the gene regulatory network underlying EMT. Front Oncol 10:554. https://doi.org/10.3389/fonc.2020.00554 Lechner J, Dash DP, Muszynska D, Hosseini M, Segev F, George S, Frazer DG, Moore JE, Kaye SB, Young T, Simpson DA, Churchill AJ, Heon E, Willoughby CE (2013) Mutational spectrum of the ZEB1 gene in corneal dystrophies supports a genotype-phenotype correlation. Invest Ophthalmol Vis Sci 54(5):3215–3223. https://doi.org/10.1167/iovs.13-11781 Leonard BC, Jalilian I, Raghunathan VK, Wang W, Jun AS, Murphy CJ, Thomasy SM (2019) Biomechanical changes to Descemet’s membrane precede endothelial cell loss in an early-onset murine model of Fuchs endothelial corneal dystrophy. Exp Eye Res 180:18–22. https://doi.org/ 10.1016/j.exer.2018.11.021 Liu J, Seet LF, Koh LW, Venkatraman A, Venkataraman D, Mohan RR, Praetorius J, Bonanno JA, Aung T, Vithana EN (2012) Depletion of SLC4A11 causes cell death by apoptosis in an immortalized human corneal endothelial cell line. Invest Ophthalmol Vis Sci 53 (7):3270–3279. https://doi.org/10.1167/iovs.11-8724 Liu C, Chen Y, Kochevar IE, Jurkunas UV (2014) Decreased DJ-1 leads to impaired Nrf2-regulated antioxidant defense and increased UV-A-induced apoptosis in corneal endothelial cells. Invest Ophthalmol Vis Sci 55(9):5551–5560. https://doi.org/10.1167/iovs.14-14580 Loganathan SK, Casey JR (2014) Corneal dystrophy-causing SLC4A11 mutants: suitability for folding-correction therapy. Hum Mutat 35(9):1082–1091. https://doi.org/10.1002/humu.22601 Loganathan SK, Schneider HP, Morgan PE, Deitmer JW, Casey JR (2016) Functional assessment of SLC4A11, an integral membrane protein mutated in corneal dystrophies. Am J Physiol Cell Physiol 311(5):C735–C748. https://doi.org/10.1152/ajpcell.00078.2016 Lopez IA, Rosenblatt MI, Kim C, Galbraith GC, Jones SM, Kao L, Newman D, Liu W, Yeh S, Pushkin A, Abuladze N, Kurtz I (2009) Slc4a11 gene disruption in mice: cellular targets of sensorineuronal abnormalities. J Biol Chem 284(39):26882–26896. https://doi.org/10.1074/jbc. M109.008102 Lumb R, Buckberry S, Secker G, Lawrence D, Schwarz Q (2017) Transcriptome profiling reveals expression signatures of cranial neural crest cells arising from different axial levels. BMC Dev Biol 17(1):5. https://doi.org/10.1186/s12861-017-0147-z Malhotra D, Loganathan SK, Chiu AM, Lukowski CM, Casey JR (2019a) Human corneal expression of SLC4A11, a gene mutated in endothelial corneal dystrophies. Sci Rep 9(1):9681. https:// doi.org/10.1038/s41598-019-46094-y Malhotra D, Jung M, Fecher-Trost C, Lovatt M, Peh GSL, Noskov S, Mehta JS, Zimmermann R, Casey JR (2019b) Defective cell adhesion function of solute transporter, SLC4A11, in endothelial corneal dystrophies. Hum Mol Genet. https://doi.org/10.1093/hmg/ddz259 Mankodi A, Urbinati CR, Yuan QP, Moxley RT, Sansone V, Krym M, Henderson D, Schalling M, Swanson MS, Thornton CA (2001) Muscleblind localizes to nuclear foci of aberrant RNA in myotonic dystrophy types 1 and 2. Hum Mol Genet 10(19):2165–2170. https://doi.org/10.1093/ hmg/10.19.2165 McLaren JW, Bachman LA, Kane KM, Patel SV (2014) Objective assessment of the corneal endothelium in Fuchs’ endothelial dystrophy. Invest Ophthalmol Vis Sci 55(2):1184–1190. https://doi.org/10.1167/iovs.13-13041 Meek KM, Boote C (2004) The organization of collagen in the corneal stroma. Exp Eye Res 78 (3):503–512 Meek KM, Knupp C (2015) Corneal structure and transparency. Prog Retin Eye Res 49:1–16. https://doi.org/10.1016/j.preteyeres.2015.07.001 Meng H, Matthaei M, Ramanan N, Grebe R, Chakravarti S, Speck CL, Kimos M, Vij N, Eberhart CG, Jun AS (2013) L450W and Q455K col8a2 knock-in mouse models of Fuchs endothelial

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

77

corneal dystrophy show distinct phenotypes and evidence for altered autophagy. Invest Ophthalmol Vis Sci 54(3):1887–1897. https://doi.org/10.1167/iovs.12-11021 Mittal V, Mittal R, Sangwan VS (2011) Successful Descemet stripping endothelial keratoplasty in congenital hereditary endothelial dystrophy. Cornea 30(3):354–356. https://doi.org/10.1097/ ICO.0b013e3181e8441a Miyai T (2018) Fuchs endothelial corneal dystrophy and mitochondria. Cornea 37(Suppl 1):S74– s77. https://doi.org/10.1097/ico.0000000000001746 Miyai T, Vasanth S, Melangath G, Deshpande N, Kumar V, Benischke AS, Chen Y, Price MO, Price FW Jr, Jurkunas UV (2019) Activation of PINK1-parkin-mediated mitophagy degrades mitochondrial quality control proteins in fuchs endothelial corneal dystrophy. Am J Pathol 189 (10):2061–2076. https://doi.org/10.1016/j.ajpath.2019.06.012 Moazzeni H, Javadi MA, Asgari D, Khani M, Emami M, Moghadam A, Panahi-Bazaz MR, Hosseini Tehrani M, Karimian F, Hosseini B, Nekuie Moghadam T, Hassanpour H, Akbari MT, Elahi E (2019) Observation of nine previously reported and 10 non-reported SLC4A11 mutations among 20 Iranian CHED probands and identification of an MPDZ mutation as possible cause of CHED and FECD in one family. Br J Ophthalmol:e314377. https://doi.org/ 10.1136/bjophthalmol-2019-314377 Mootha VV, Hussain I, Cunnusamy K, Graham E, Gong X, Neelam S, Xing C, Kittler R, Petroll WM (2015) TCF4 triplet repeat expansion and nuclear RNA foci in Fuchs' endothelial corneal dystrophy. Invest Ophthalmol Vis Sci 56(3):2003–2011. https://doi.org/10.1167/iovs.14-16222 Mootha VV, Hansen B, Rong Z, Mammen PP, Zhou Z, Xing C, Gong X (2017) Fuchs’ endothelial corneal dystrophy and RNA foci in patients with myotonic dystrophy. Invest Ophthalmol Vis Sci 58(11):4579–4585. https://doi.org/10.1167/iovs.17-22350 Moschos MM, Diamantopoulou A, Gouliopoulos N, Droutsas K, Bagli E, Chatzistefanou K, Kitsos G, Kroupis C (2019) TCF4 and COL8A2 gene polymorphism screening in a greek population of late-onset Fuchs endothelial corneal dystrophy. In Vivo 33(3):963–971. https:// doi.org/10.21873/invivo.11565 Mould AP, Askari JA, Craig SE, Garratt AN, Clements J, Humphries MJ (1994) Integrin alpha 4 beta 1-mediated melanoma cell adhesion and migration on vascular cell adhesion molecule-1 (VCAM-1) and the alternatively spliced IIICS region of fibronectin. J Biol Chem 269 (44):27224–27230 Murray P, Frampton G, Nelson PN (1999) Cell adhesion molecules. Sticky moments in the clinic. BMJ 319(7206):332–334. https://doi.org/10.1136/bmj.319.7206.332 Murthy CR, Rama Rao KV, Bai G, Norenberg MD (2001) Ammonia-induced production of free radicals in primary cultures of rat astrocytes. J Neurosci Res 66(2):282–288. https://doi.org/10. 1002/jnr.1222 Musch DC, Niziol LM, Stein JD, Kamyar RM, Sugar A (2011) Prevalence of corneal dystrophies in the United States: estimates from claims data. Invest Ophthalmol Vis Sci 52(9):6959–6963. https://doi.org/10.1167/iovs.11-7771 Myers EJ, Marshall A, Jennings ML, Parker MD (2016) Mouse slc4a11 expressed in Xenopus oocytes is an ideally selective H+/OH conductance pathway that is stimulated by rises in intracellular and extracellular pH. Am J Physiol Cell Physiol 311(6):C945–C959. https://doi. org/10.1152/ajpcell.00259.2016 Ogando DG, Jalimarada SS, Zhang W, Vithana EN, Bonanno JA (2013) SLC4A11 is an EIPAsensitive Na+ permeable pHi regulator. Am J Physiol Cell Physiol 305(7):C716–C727. https:// doi.org/10.1152/ajpcell.00056.2013 Ogando DG, Choi M, Shyam R, Li S, Bonanno JA (2019) Ammonia sensitive SLC4A11 mitochondrial uncoupling reduces glutamine induced oxidative stress. Redox Biol 26:101260. https://doi.org/10.1016/j.redox.2019.101260 Ogunbona OB, Claypool SM (2019) Emerging roles in the biogenesis of cytochrome c oxidase for members of the mitochondrial carrier family. Front Cell Dev Biol 7:3. https://doi.org/10.3389/ fcell.2019.00003

78

D. Malhotra and J. R. Casey

O'Neill MA, Ishii T, Albersheim P, Darvill AG (2004) Rhamnogalacturonan II: structure and function of a borate cross-linked cell wall pectic polysaccharide. Annu Rev Plant Biol 55:109–139. https://doi.org/10.1146/annurev.arplant.55.031903.141750 Pagliarini DJ, Calvo SE, Chang B, Sheth SA, Vafai SB, Ong SE, Walford GA, Sugiana C, Boneh A, Chen WK, Hill DE, Vidal M, Evans JG, Thorburn DR, Carr SA, Mootha VK (2008) A mitochondrial protein compendium elucidates complex I disease biology. Cell 134 (1):112–123. https://doi.org/10.1016/j.cell.2008.06.016 Park M, Li Q, Shcheynikov N, Zeng W, Muallem S (2004) NaBC1 is a ubiquitous electrogenic Na+ coupled borate transporter essential for cellular boron homeostasis and cell growth and proliferation. Mol Cell 16(3):331–341. https://doi.org/10.1016/j.molcel.2004.09.030 Parker MD, Ourmozdi EP, Tanner MJ (2001) Human BTR1, a new bicarbonate transporter superfamily member and human AE4 from kidney. Biochem Biophys Res Commun 282 (5):1103–1109 Petroll WM, Hsu JK, Bean J, Cavanagh HD, Jester JV (1999) The spatial organization of apical junctional complex-associated proteins in feline and human corneal endothelium. Curr Eye Res 18(1):10–19 Plantefaber LC, Hynes RO (1989) Changes in integrin receptors on oncogenically transformed cells. Cell 56(2):281–290 Quade BN, Marshall A, Parker MD (2020) The pH dependence of the Slc4a11-mediated H+conductance is influenced by intracellular lysine residues and modified by disease-linked mutations. Am J Physiol Cell Physiol. https://doi.org/10.1152/ajpcell.00128.2020 Riazuddin SA, Eghrari AO, Al-Saif A, Davey L, Meadows DN, Katsanis N, Gottsch JD (2009) Linkage of a mild late-onset phenotype of Fuchs corneal dystrophy to a novel locus at 5q33.1q35.2. Invest Ophthalmol Vis Sci 50(12):5667–5671. https://doi.org/10.1167/iovs.09-3764 Riazuddin SA, Zaghloul NA, Al-Saif A, Davey L, Diplas BH, Meadows DN, Eghrari AO, Minear MA, Li YJ, Klintworth GK, Afshari N, Gregory SG, Gottsch JD, Katsanis N (2010) Missense mutations in TCF8 cause late-onset Fuchs corneal dystrophy and interact with FCD4 on chromosome 9p. Am J Hum Genet 86(1):45–53. https://doi.org/10.1016/j.ajhg.2009.12.001 Riazuddin SA, Parker DS, McGlumphy EJ, Oh EC, Iliff BW, Schmedt T, Jurkunas U, Schleif R, Katsanis N, Gottsch JD (2012) Mutations in LOXHD1, a recessive-deafness locus, cause dominant late-onset Fuchs corneal dystrophy. Am J Hum Genet 90(3):533–539. https://doi. org/10.1016/j.ajhg.2012.01.013 Riazuddin SA, Vasanth S, Katsanis N, Gottsch JD (2013a) Mutations in AGBL1 cause dominant late-onset Fuchs corneal dystrophy and alter protein-protein interaction with TCF4. Am J Hum Genet 93(4):758–764. https://doi.org/10.1016/j.ajhg.2013.08.010 Riazuddin SA, Vasanth S, Katsanis N, Gottsch JD (2013b) Mutations in AGBL1 cause dominant late-onset Fuchs corneal dystrophy and alter protein-protein interaction with TCF4. Am J Hum Genet 93(4):758–764. https://doi.org/10.1016/j.ajhg.2013.08.010 Roy S, Praneetha DC, Vendra VP (2015) Mutations in the corneal endothelial dystrophy-associated gene SLC4A11 render the cells more vulnerable to oxidative insults. Cornea 34(6):668–674. https://doi.org/10.1097/ICO.0000000000000421 Sanchez-Tillo E, de Barrios O, Siles L, Cuatrecasas M, Castells A, Postigo A (2011) Beta-catenin/ TCF4 complex induces the epithelial-to-mesenchymal transition (EMT)-activator ZEB1 to regulate tumor invasiveness. Proc Natl Acad Sci U S A 108(48):19204–19209. https://doi. org/10.1073/pnas.1108977108 Sanchez-Tillo E, de Barrios O, Valls E, Darling DS, Castells A, Postigo A (2015) ZEB1 and TCF4 reciprocally modulate their transcriptional activities to regulate Wnt target gene expression. Oncogene 34(46):5760–5770. https://doi.org/10.1038/onc.2015.352 Sanchis-Gimeno JA, Lleo-Perez A, Alonso L, Rahhal MS, Martinez Soriano F (2005) Corneal endothelial cell density decreases with age in emmetropic eyes. Histol Histopathol 20 (2):423–427. https://doi.org/10.14670/HH-20.423

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

79

Schey KL, Wang Z, Wenke JL, Qi Y (2014) Aquaporins in the eye: expression, function, and roles in ocular disease. Biochim Biophys Acta 1840(5):1513–1523. https://doi.org/10.1016/j.bbagen. 2013.10.037 Schlotzer-Schrehardt U, Bachmann BO, Laaser K, Cursiefen C, Kruse FE (2011) Characterization of the cleavage plane in Descemet's membrane endothelial keratoplasty. Ophthalmology 118 (10):1950–1957. https://doi.org/10.1016/j.ophtha.2011.03.025 Sherrard ES, Novakovic P, Speedwell L (1987) Age-related changes of the corneal endothelium and stroma as seen in vivo by specular microscopy. Eye (Lond) 1(Pt 2):197–203. https://doi.org/10. 1038/eye.1987.37 Siddiqui S, Zenteno JC, Rice A, Chacon-Camacho O, Naylor SG, Rivera-de la Parra D, Spokes DM, James N, Toomes C, Inglehearn CF, Ali M (2014) Congenital hereditary endothelial dystrophy caused by SLC4A11 mutations progresses to Harboyan syndrome. Cornea 33 (3):247–251. https://doi.org/10.1097/ICO.0000000000000041 SNPEffect (2019) ZN717_HUMAN. http://snpeffect.switchlab.org/uniprot/ZN717_HUMAN Soh YQ, Peh Swee Lim G, Htoon HM, Gong X, Mootha VV, Vithana EN, Kocaba V, Mehta JS (2019) Trinucleotide repeat expansion length as a predictor of the clinical progression of Fuchs’ endothelial corneal dystrophy. PLoS One 14(1):e0210996. https://doi.org/10.1371/journal. pone.0210996 Soh YQ, Kocaba V, Pinto M, Mehta JS (2020) Fuchs endothelial corneal dystrophy and corneal endothelial diseases: east meets west. Eye (Lond) 34(3):427–441. https://doi.org/10.1038/ s41433-019-0497-9 Soliman AZ, Xing C, Radwan SH, Gong X, Mootha VV (2015) Correlation of severity of Fuchs endothelial corneal dystrophy with triplet repeat expansion in TCF4. JAMA Ophthalmol 133 (12):1386–1391. https://doi.org/10.1001/jamaophthalmol.2015.3430 Stehouwer M, Bijlsma WR, Van der Lelij A (2011) Hearing disability in patients with Fuchs’ endothelial corneal dystrophy: unrecognized co-pathology? Clin Ophthalmol 5:1297–1301. https://doi.org/10.2147/opth.S23091 Stepp MA (2006) Corneal integrins and their functions. Exp Eye Res 83(1):3–15. https://doi.org/10. 1016/j.exer.2006.01.010 Sultana A, Garg P, Ramamurthy B, Vemuganti GK, Kannabiran C (2007) Mutational spectrum of the SLC4A11 gene in autosomal recessive congenital hereditary endothelial dystrophy. Mol Vis 13:1327–1332 Sun Z, Zhang S, Chan JY, Zhang DD (2007) Keap1 controls postinduction repression of the Nrf2mediated antioxidant response by escorting nuclear export of Nrf2. Mol Cell Biol 27 (18):6334–6349. https://doi.org/10.1128/MCB.00630-07 Sun S, Wang Y, Wu Y, Gao Y, Li Q, Abdulrahman AA, Liu XF, Ji GQ, Gao J, Li L, Wan FP, Li YQ, Gao DS (2018) Identification of COL1A1 as an invasionrelated gene in malignant astrocytoma. Int J Oncol 53(6):2542–2554. https://doi.org/10.3892/ijo.2018.4568 Sweeney DF, Xie RZ, O'Leary DJ, Vannas A, Odell R, Schindhelm K, Cheng HY, Steele JG, Holden BA (1998) Nutritional requirements of the corneal epithelium and anterior stroma: clinical findings. Invest Ophthalmol Vis Sci 39(2):284–291 Syed ZA, Tran JA, Jurkunas UV (2017) Peripheral endothelial cell count is a predictor of disease severity in advanced fuchs endothelial corneal dystrophy. Cornea 36(10):1166–1171. https:// doi.org/10.1097/ICO.0000000000001292 Takano J, Noguchi K, Yasumori M, Kobayashi M, Gajdos Z, Miwa K, Hayashi H, Yoneyama T, Fujiwara T (2002) Arabidopsis boron transporter for xylem loading. Nature 420 (6913):337–340. https://doi.org/10.1038/nature01139 TCGA (2019) The cancer genome atlas-SLC4A11. https://portal.gdc.cancer.gov/genes/ ENSG00000088836 Thiagarajah JR, Verkman AS (2002) Aquaporin deletion in mice reduces corneal water permeability and delays restoration of transparency after swelling. J Biol Chem 277(21):19139–19144. https://doi.org/10.1074/jbc.M202071200

80

D. Malhotra and J. R. Casey

THPA (2019) The human protein Atlas-SLC4A11. https://www.proteinatlas.org/ ENSG00000088836-SLC4A11/tissue Torricelli AA, Wilson SE (2014) Cellular and extracellular matrix modulation of corneal stromal opacity. Exp Eye Res 129:151–160. https://doi.org/10.1016/j.exer.2014.09.013 Tuft SJ, Coster DJ (1990) The corneal endothelium. Eye (Lond) 4(Pt 3):389–424. https://doi.org/ 10.1038/eye.1990.53 Vallabh NA, Romano V, Willoughby CE (2017) Mitochondrial dysfunction and oxidative stress in corneal disease. Mitochondrion 36:103–113. https://doi.org/10.1016/j.mito.2017.05.009 Vasanth S, Eghrari AO, Gapsis BC, Wang J, Haller NF, Stark WJ, Katsanis N, Riazuddin SA, Gottsch JD (2015) Expansion of CTG18.1 trinucleotide repeat in TCF4 is a potent driver of Fuchs' corneal dystrophy. Invest Ophthalmol Vis Sci 56(8):4531–4536. https://doi.org/10.1167/ iovs.14-16122 Vedana G, Villarreal G Jr, Jun AS (2016) Fuchs endothelial corneal dystrophy: current perspectives. Clin Ophthalmol 10:321–330. https://doi.org/10.2147/OPTH.S83467 Vilas GL, Loganathan SK, Liu J, Riau AK, Young JD, Mehta JS, Vithana EN, Casey JR (2013) Transmembrane water-flux through SLC4A11: a route defective in genetic corneal diseases. Hum Mol Genet 22(22):4579–4590. https://doi.org/10.1093/hmg/ddt307 Vincent AL (2014) Corneal dystrophies and genetics in the international committee for classification of corneal dystrophies era: a review. Clin Exp Ophthalmol 42(1):4–12. https://doi.org/10. 1111/ceo.12149 Vithana EN, Morgan P, Sundaresan P, Ebenezer ND, Tan DT, Mohamed MD, Anand S, Khine KO, Venkataraman D, Yong VH, Salto-Tellez M, Venkatraman A, Guo K, Hemadevi B, Srinivasan M, Prajna V, Khine M, Casey JR, Inglehearn CF, Aung T (2006) Mutations in sodium-borate cotransporter SLC4A11 cause recessive congenital hereditary endothelial dystrophy (CHED2). Nat Genet 38(7):755–757 Vithana EN, Morgan PE, Ramprasad V, Tan DT, Yong VH, Venkataraman D, Venkatraman A, Yam GH, Nagasamy S, Law RW, Rajagopal R, Pang CP, Kumaramanickevel G, Casey JR, Aung T (2008) SLC4A11 mutations in fuchs endothelial corneal dystrophy (FECD). Hum Mol Genet 17(5):656–666. https://doi.org/10.1093/hmg/ddm337 Weiss JS, Moller HU, Aldave AJ, Seitz B, Bredrup C, Kivela T, Munier FL, Rapuano CJ, Nischal KK, Kim EK, Sutphin J, Busin M, Labbe A, Kenyon KR, Kinoshita S, Lisch W (2015) IC3D classification of corneal dystrophies--edition 2. Cornea 34(2):117–159. https://doi.org/10.1097/ ICO.0000000000000307 Willoughby CE, Ponzin D, Ferrari S, Lobo A, Landau K, Omidi Y (2010) Anatomy and physiology of the human eye: effects of mucopolysaccharidoses disease onstructure and function–a review. Clin Exp Ophthalmol 38:2–11. https://doi.org/10.1111/j.1442-9071.2010.02363.x Wilson RS, Roper-Hall MJ (1982) Effect of age on the endothelial cell count in the normal eye. Br J Ophthalmol 66(8):513–515 Winkler NS, Milone M, Martinez-Thompson JM, Raja H, Aleff RA, Patel SV, Fautsch MP, Wieben ED, Baratz KH (2018) Fuchs’ endothelial corneal dystrophy in patients with myotonic dystrophy, type 1. Invest Ophthalmol Vis Sci 59(7):3053–3057. https://doi.org/10.1167/iovs.1723160 Wloga D, Joachimiak E, Fabczak H (2017) Tubulin post-translational modifications and microtubule dynamics. Int J Mol Sci 18(10). https://doi.org/10.3390/ijms18102207 Xia D, Zhang S, Nielsen E, Ivarsen AR, Liang C, Li Q, Thomsen K, Hjortdal JO, Dong M (2016) The ultrastructures and mechanical properties of the descement’s membrane in Fuchs endothelial corneal dystrophy. Sci Rep 6:23096. https://doi.org/10.1038/srep23096 Yadava RS, Kim YK, Mandal M, Mahadevan K, Gladman JT, Yu Q, Mahadevan MS (2019) MBNL1 overexpression is not sufficient to rescue the phenotypes in a mouse model of RNA toxicity. Hum Mol Genet 28(14):2330–2338. https://doi.org/10.1093/hmg/ddz065 Yang Y, Su Y, Wang D, Chen Y, Wu T, Li G, Sun X, Cui L (2013) Tanshinol attenuates the deleterious effects of oxidative stress on osteoblastic differentiation via Wnt/FoxO3a signaling. Oxidative Med Cell Longev 2013:351895. https://doi.org/10.1155/2013/351895

Molecular Mechanisms of Fuchs and Congenital Hereditary Endothelial Corneal. . .

81

Yang J, Simonneau C, Kilker R, Oakley L, Byrne MD, Nichtova Z, Stefanescu I, Pardeep-Kumar F, Tripathi S, Londin E, Saugier-Veber P, Willard B, Thakur M, Pickup S, Ishikawa H, Schroten H, Smeyne R, Horowitz A (2019) Murine MPDZ-linked hydrocephalus is caused by hyperpermeability of the choroid plexus. EMBO Mol Med 11(1). https://doi.org/10.15252/ emmm.201809540 Zarouchlioti C, Sanchez-Pintado B, Hafford Tear NJ, Klein P, Liskova P, Dulla K, Semo M, Vugler AA, Muthusamy K, Dudakova L, Levis HJ, Skalicka P, Hysi P, Cheetham ME, Tuft SJ, Adamson P, Hardcastle AJ, Davidson AE (2018) Antisense therapy for a common corneal dystrophy ameliorates TCF4 repeat expansion-mediated toxicity. Am J Hum Genet 102 (4):528–539. https://doi.org/10.1016/j.ajhg.2018.02.010 Zhang W, Aldave AJ (2018) Corneal endothelial dystrophies belong to the same series of disorders connected by transcriptional factors ZEB1 and TCF4/E2-2. Invest Ophthalmol Vis Sci 59 (9):1351 Zhang X, Igo RP Jr, Fondran J, Mootha VV, Oliva M, Hammersmith K, Sugar A, Lass JH, Iyengar SK, Fuchs' Genetics Multi-Center Study G (2013) Association of smoking and other risk factors with Fuchs' endothelial corneal dystrophy severity and corneal thickness. Invest Ophthalmol Vis Sci 54(8):5829–5835. https://doi.org/10.1167/iovs.13-11918 Zhang W, Ogando DG, Bonanno JA, Obukhov AG (2015) Human SLC4A11 is a novel NH3/H+ co-transporter. J Biol Chem 290(27):16894–16905. https://doi.org/10.1074/jbc.M114.627455 Zhang W, Li H, Ogando DG, Li S, Feng M, Price FW Jr, Tennessen JM, Bonanno JA (2017) Glutaminolysis is essential for energy production and ion transport in human corneal endothelium. EBioMedicine 16:292–301. https://doi.org/10.1016/j.ebiom.2017.01.004 Zhang J, Wu D, Li Y, Fan Y, Chen H, Hong J, Xu J (2019) Novel mutations associated with various types of corneal dystrophies in a han Chinese population. Front Genet 10:881. https://doi.org/ 10.3389/fgene.2019.00881

Rev Physiol Biochem Pharmacol (2020) 178: 83–110 https://doi.org/10.1007/112_2020_42 © The Author(s), under exclusive license to Springer Nature Switzerland AG 2020 Published online: 14 August 2020

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In Vivo, and Clinical Studies Alireza Fatahian, Saeed Mohammadian Haftcheshmeh, Sara Azhdari, Helaleh Kaboli Farshchi, Banafsheh Nikfar, and Amir Abbas Momtazi-Borojeni

Contents 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 Berberine Lowers Atherogenic Lipids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 In Vivo Evidence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Clinical Evidence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Mechanisms Underlying the Cholesterol-Lowering Effect of Berberine . . . . . . . . . . . . . 3 Berberine’s Effects on Atherosclerosis Lesion Progression: In Vitro Evidence . . . . . . . . . . . . 3.1 Berberine Improves Endothelial Cell Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Berberine Modulates Atherogenic Activities of Macrophages . . . . . . . . . . . . . . . . . . . . . . . .

A. Fatahian Department of Cardiology, Cardiovascular Research Center, Mazandaran University of Medical Sciences, Sari, Iran S. M. Haftcheshmeh Department of Medical Immunology, Faculty of Medicine, Mashhad University of Medical Sciences, Mashhad, Iran S. Azhdari Department of Anatomy and Embryology, School of Medicine, Bam University of Medical Sciences, Bam, Iran H. K. Farshchi Department of Horticulture, Faculty of Agriculture, Ferdowsi University of Mashhad, Mashhad, Iran B. Nikfar (*) Pars Advanced and Minimally Invasive Medical Manners Research Center, Pars Hospital, Iran University of Medical Sciences, Tehran, Iran e-mail: [email protected] A. A. Momtazi-Borojeni (*) Halal research center of IRI, FDA, Tehran, Iran Department of Medical Biotechnology, Faculty of Medicine, Mashhad University of Medical Sciences, Mashhad, Iran e-mail: [email protected]

85 86 86 89 92 94 94 95

84

A. Fatahian et al.

3.3 Berberine Blocks Proliferation and Migration of Vascular Smooth Muscle Cells . . 97 4 Berberine’s Effects on Atherosclerosis Lesion Progression: In Vivo Evidence . . . . . . . . . . . 101 5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

Abstract Elevated levels of plasma cholesterol, impaired vascular wall, and presence of inflammatory macrophages are important atherogenic risk factors contributing to atherosclerotic plaque formation and progression. The interventions modulating these risk factors have been found to protect against atherosclerosis development and to decrease atherosclerosis-related cardiovascular disorders. Nutritional approaches involving supplements followed by improving dietary habits and lifestyle have become growingly attractive and acceptable methods used to control atherosclerosis risk factors, mainly high levels of plasma cholesterol. There are a large number of studies that show berberine, a plant bioactive compound, could ameliorate atherosclerosis-related risk factors. In the present literature review, we put together this studies and provide integrated evidence that exhibits berberine has the potential atheroprotective effect through reducing increased levels of plasma cholesterol, particularly low-density lipoprotein (LDL) cholesterol (LDL-C) via LDL receptor (LDLR)-dependent and LDL receptor-independent mechanisms, inhibiting migration and inflammatory activity of macrophages, improving the functionality of endothelial cells via anti-oxidant activities, and suppressing proliferation of vascular smooth muscle cells. In conclusion, berberine can exert inhibitory effects on the atherosclerotic plaque development mainly through LDL-lowering activity and suppressing atherogenic functions of mentioned cells. As the second achievement of this review, among the signaling pathways through which berberine regulates intracellular processes, AMP-activated protein kinase (AMPK) has a central and critical role, showing that enhancing activity of AMPK pathway can be considered as a promising therapeutic approach for atherosclerosis treatment. Keywords AMP-activated protein kinase · Atherosclerosis · Berberine · Cholesterol · Endothelial cells · LDL-C · Macrophage dysfunction · Vascular smooth muscle cell

Abbreviations 30 -UTR ABCA1 AMP AMPK EMMPRIN eNOS ERK HMG-CoA

30 -untranslated region ATP-binding membrane cassette transport protein A1 Adenosine monophosphate AMP-activated protein kinase Exhibited a decreased expression of MMPs and extracellular MMP inducer endothelial nitric oxide synthase Extracellular receptor-activated kinase 3-hydroxy-3-methyl-glutaryl-coenzyme A

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

HNF1α HUVECs IL-6 JNK LDL LDL-C LDLR LOX 1 LXRα lysoPC MAP MCP-1 MIP-1α MMPs NADPH NAFLD NOD PCSK9 PCSK9 ROS SR-BI SREBP2 TG VCAM-1 VSMCs

85

hepatocytes nuclear factor 1α Human umbilical vein endothelial cells Interleukin 6 c-Jun N-terminal kinase Low-density lipoprotein LDL cholesterol LDL receptor Low-density lipoprotein receptor 1 Liver X receptor α Lysophosphatidylcholine Mitogen-activated protein Monocyte chemoattractant protein-1 Macrophage inflammatory Protein 1 alpha Matrix metalloproteinases Nicotinamide adenine dinucleotide phosphate Non-alcohol fatty liver disease Nonobese diabetic Proprotein convertase subtilisin kexin 9 Proprotein convertase subtilisin/kexin type 9 Reactive oxygen species Scavenger receptor class B type I sterol regulatory element-binding protein 2 Triglyceride Vascular cell adhesion molecule-1 Vascular smooth muscle cells

1 Introduction Berberine is an isoquinoline alkaloid presented as the principal bioactive ingredient in stem, bark, rhizome, and roots of several plants, including barberry (Berberis vulgaris), Coptis (Coptis chinensis), goldenseal (Hydrastis canadensis), tree turmeric (Berberis aristata), and Oregon grape (Berberis aquifolium) (Singh and Mahajan 2013; Srivastava et al. 2015). Evidence from traditional and modern medicine shows that berberine exerts polytrophic pharmacological effects, including lipid-lowering, anti-diabetic, anti-tumor, anti-inflammatory, anti-diarrheal, and antimicrobial activities (Ayati et al. 2017; Imanshahidi and Hosseinzadeh 2008; Li et al. 2014; Singh et al. 2010; Zhang et al. 2019). A growing body of research confirms the anti-atherogenic properties of berberine, although there is no enough report showing its direct effect on atherosclerotic plaque formation and progression. Herewith, to exhibit the possible protective effect of berberine on atherosclerotic plaque progression, the present review was aimed to gather underlying mechanisms ascribed to the anti-atherogenic effects of berberine.

86

A. Fatahian et al.

As discussed in the following sections, there are documented findings demonstrating that berberine has potential anti-atherosclerotic effects through protecting or lowering hypercholesterolemia, modulating the atherogenic activity of inflammatory macrophages, and improving abnormal functions of vascular smooth muscle and endothelial cells.

2 Berberine Lowers Atherogenic Lipids Atherogenic lipids, especially low-density lipoprotein (LDL) cholesterol (LDL-C), have a casual effect on atherosclerotic plaque development and progression (Ference et al. 2017). Various LDL-lowering medications, such as statins and proprotein convertase subtilisin/kexin type 9 (PCSK9) inhibitors, have been found to reduce the risk of atherosclerosis-related cardiovascular disorders proportional to the absolute reduction of LDL-C in numerous randomized trials (Ference et al. 2017; Ridker et al. 2017; Sabatine et al. 2017; Sahebkar and Watts 2013). Hence, an LDL-lowering agent may have the potential to exert protective and therapeutic effects against atherosclerotic plaque progression. Although statin therapy is the most commonly used approach to treat hypercholesterolemia, there might be a relatively large number of patients who are statin-resistant or statin-intolerant and unable to achieve optimal LDL-C levels despite intensive statin therapy (Toth et al. 2018; Ward et al. 2019). Therefore, exploring complementary or alternative LDL-lowering agents to prevent or treat atherosclerotic lesions is important. There are several in vivo, clinical, and mechanistic studies that show berberine can efficiently reduce increased levels of atherogenic lipids, mainly LDL-C, suggesting that this natural compound has the strong potential to protect against atherosclerotic plaque development.

2.1

In Vivo Evidence

A large number of experimental studies on rodent models of diet-induced hypercholesterolemia have been conducted to determine the cholesterol-lowering effects of the various doses and administration routes of berberine (Table 1). Since an early report introduced berberine as a new cholesterol-lowering agent (Kong et al. 2004), a growing body of experimental studies has further investigated and confirmed the therapeutic potential of berberine on hypercholesterolemia in recent years (Briand et al. 2013; Brusq et al. 2006; Chang et al. 2012; He et al. 2016; Hu and Davies 2010; Hu et al. 2012; Jia et al. 2008; Kong et al. 2004; Li et al. 2011; Wang et al. 2013, 2014; Xiao et al. 2012). As reviewed elsewhere, berberine can modulate the dysregulated lipid profile through reducing plasma levels of TC, LDL-C, and TG as well as increasing plasma HDL-C in various rodent models of hyperlipidemia (Wang and Zidichouski 2018). These results are supported by the recent animal

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

87

Table 1 In vivo lipid-lowering effect of berberine in various animal models Animal model Male ApoE/ mice

Route and time of administration Daily gavage for 13 weeks

Effect on plasma lipid profile Decreasing TC and LDL-C Increasing HDL-C

Ref. Wu et al. (2020)

Daily gavage for 2 weeks

TC (28%) LDL-C (51%)

100 mg/kg/ day

Daily gavage for 4 weeks

Significantly decreasing TC and LDL-C Significantly increasing HDL-C

30 mg/kg/ day

Daily gavage for 4 weeks

TC (10.5) LDL-C (9.8%)

50, 100, 150 mg/ kg/day 100 mg/kg/ day

Daily gavage for 8 weeks

TC (29%, 33%, 33%), non-HDLC (31%, 41%, 38%), at 50, 100, and 150 mg/kg/day, respectively No effect on both TC and non-HDL-C. but (31%) plasma TG levels

Singh and Liu (2019) Kim et al. (2019) Zhu et al. (2018) Wang et al. (2014) Jia et al. (2008) Chang et al. (2012) Brusq et al. (2006) Briand et al. (2013) Xiao et al. (2012) Hu et al. (2012) Kong et al. (2004) He et al. (2016) Abidi et al. (2006) Wang et al. (2011)

Berberine dose 50 mg/kg/ day 100 mg/kg/ day 200 mg/kg/ day

Male C57BL/6 mice SpragueDawley male rats Female C57BL/ 6 J mice SpragueDawley male rats SpragueDawley male rats SpragueDawley male rats Male hamsters

200 mg/kg/ day

Daily gavage for 16 weeks

TC (28%) and HDL-C (41%)

100 mg/kg/ day

Inhibited both cholesterol and TG synthesis

Male hamsters

150 mg/kg/ day

Female C57BL/ 6 J mice SpragueDawley rat Female hamsters

10 and 30 mg/kg/ day 500 mg/kg

Orally twice a day for 10 days Daily injection for 2 weeks Daily gavage for 4 weeks

Hamsters

50 and 100 mg/kg/ day 46.7 mg/kg/ day

Daily gavage for 6 weeks

3 times a day gavage for 12 weeks Daily gavage for 10 days

LDL-C (35%), TG (34%)

TC (42%, 56%) and TG (37%, 47%) at 10 and 30 mg/kg/day, respectively T-C (9%) and TG (34.7%)

LDL-C (26%, 42%), at 50 and100 mg/kg/day, respectively

Daily gavage for 140 days

TC (19%) and HDL-C (11.34%)

Male hamsters

1.8 mg/kg/ day

Daily gavage for 24 days

TC (30%), TG (34%) and LDL-C (9.2%)

Male rats

7, 60, and 300 mg/kg/ day

Daily gavage for 12 weeks

Decreased TC and LDL-C, but increased HDL-C

(continued)

88

A. Fatahian et al.

Table 1 (continued) Animal model Rats

Berberine dose 200 mg/kg/ day

Male hamsters

50 and 100 mg/kg/ day

Route and time of administration Daily gavage for 16 weeks Daily gavage for 10 days

Effect on plasma lipid profile T-C (29%) and HDL-C (41%)

TC (44%, 70%,), TG (34%, 51%), and LDL-C (47%, 71%) at 50 and100 mg/kg/day, respectively

Ref. Chang et al. (2010) Li et al. (2008)

study that showed the oral administration of berberine (100 mg/kg/day) via gavage feeding or dietary supplementation can ameliorate severe hypercholesterolemia in hyperlipidemic rats (Kim et al. 2019). It is consistent with another study that demonstrated the oral gavage of berberine at both high dose (100 mg/kg/day) and low dose (50 mg/kg/day) could effectively and dose-dependently reduce elevated levels of plasma TC and LDL-C in ApoE/ mice fed with a high-fat diet, which the effect of high dose was stronger than that of the low dose (Wu et al. 2020). Further study revealed that berberine treatment (200 mg/kg/day) could decrease plasma levels of TC and LDL-C by 28% and 51% in wild-type mice fed a high cholesterol diet (Singh and Liu 2019). It was also found that intraperitoneal (i.p.) injection of berberine (1.5–5 mg/kg/day) could decrease plasma cholesterol to a similar amount as compared to oral administration; however, the therapeutic dose was decreased by 10 to 100-folds in i.p. route (Abidi et al. 2006; Hu and Davies 2010; Kim et al. 2009; Wu et al. 2010). Although i.p. injection of berberine shows similar cholesterollowering effects at a lower dose, it is not an acceptable route for administration in humans because of its inconvenience and invasiveness. The cholesterol-lowering potential of berberine is further supported by the study that showed oral administration of berberine (90 mg/kg/day) and simvastatin (6 mg/ kg/day) exerted similar reductions of LDL-C (27% and 28%, respectively) in rats fed with a high-cholesterol diet. Of note, a combination of berberine with simvastatin decreased plasma LDL-C by 46%, which was significantly higher than that of either berberine or simvastatin monotherapy, resulting in the reduction of simvastatin dose by 50% that still exerts similar lipid-lowering effect (Kong et al. 2008). It was also reported that combination therapy of berberine (orally, 30 mg/kg/day) with resveratrol (orally, 20 mg/kg/day) in hyperlipidemic mice decreased plasma levels of total cholesterol by 27% and LDL-C by 31.6%, which was significantly greater than that of berberine (10.5% and 9.8%) or the resveratrol (8.4% and 6.6%) monotherapy (Zhu et al. 2018). Therefore, berberine can be considered as an add-on therapy to improve lipid-lowering effects and decrease the therapeutic dosage. Besides, the cholesterol-lowering effect of berberine is also documented by other studies where it has been shown that administration of berberine markedly reduced the plasma level of LDL-C in animal models of type 2 diabetes mellitus fed with high-fat diet (Zhang et al. 2008, 2011, 2014). Interestingly, in nonobese diabetic (NOD) mice, berberine supplementation (50, 150, and 500 mg/kg/day) significantly

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

89

decreased the ratio of LDL-C/TC in a dose-dependent manner (Chueh and Lin 2011). Moreover, studies carried out in animal models of hyperlipidemia and non-alcohol fatty liver disease (NAFLD) have investigated lipid-modulating effects of berberine. In a good agreement with the previous findings, the results of these studies have also confirmed that berberine (alone or in combination with other natural compounds such as curcumin) can facilitate the treatment of hyperlipidemia and NAFLD through its LDL-lowering effects (Feng et al. 2018; Kou et al. 2016; Zhou et al. 2017). To sum up, berberine in combination with lipid-lowering drugs can efficiently ameliorate hypercholesterolemia, whereby it can reduce therapeutic doses of such drugs, like statins, and thereby might improve treatment of patients who are statin intolerance or statin resistance even at high-dose therapy.

2.2

Clinical Evidence

The beneficial cholesterol-lowering effect of berberine has been verified in numerous clinical trials conducted on patients with mild to moderate hypercholesterolemia in various populations (Table 2). Recently, a systematic review and meta-analysis of 16 randomized clinical trials with a total of 2,147 participants was directed to evaluate the efficacy and safety of berberine in patients with dyslipidemia. The results indicate that berberine can improve plasma lipid profile in dyslipidemia with satisfactory safety, in which significant reduction of TC and LDL-C as well as no significant incidence of adverse events is evident (Ju et al. 2018). A placebocontrolled clinical trial investigating the cholesterol-lowering efficacy of berberine in patients with mild hyperlipidemia showed that oral administration of berberine capsule (900 mg/day, for 3 months) effectively improved the plasma levels of TC and LDL-C (Wang et al. 2016b). Comparably, a clinical trial conducted on Chinese population indicated that oral administration of berberine (1 g/day, for 3 months) decreased the plasma levels of TC by 29% and LDL-C by 35% in patients with hypercholesterolemia suffering from TC levels more than 200 mg/dL (Doggrell 2005). Similarly, another clinical trial conducted on the Caucasian population has demonstrated that berberine treatment (1 g/day, for 3 months) decreased TC by 11% and LDL-C by 16%, without any adverse effects, in hypercholesterolemic patients with low cardiovascular risk (Derosa et al. 2013). Moreover, it was shown that berberine treatment could lower LDL-C (20–30%) (Johnston et al. 2017; Koppen et al. 2017) at a percentage range close to those achieved by statin therapy (30–50%) (Dong et al. 2013). Cholesterol-lowering effects of berberine have been also indicated by the other studies that investigated the effects of berberine in combination with lipid-lowering nutraceuticals and/or drugs in hypercholesterolemic patients. Of note, berberine has poor intestinal absorption and low oral bioavailability, which limits its therapeutic efficacy. Silymarin, which is a flavonolignan-rich complex extracted from the milk thistle Silybum marianum (L.), has been found to possess hepato- and cardio-

90

A. Fatahian et al.

Table 2 Clinical trials evaluating cholesterol-lowering effect of berberine in hypercholesterolemic patients

Subject Prediabetic patients

Patients with mild-tomoderate hypercholesterolemia Hyperlipidemic patients

Moderately hypercholesterolemic patients Caucasians with low cardiovascular risk

Treatment group Berberine-containing polyherbal dietary supplement, n ¼ 40 Berberine-containing food supplement, n ¼ 90; placebo, n ¼ 90 Berberine-containing nutraceuticalsa, n ¼ 30 (15/15, M/F); placebo, n ¼ 9 (3/6, M/F) AP-1b, n ¼ 51 (18/33, M/F); placebo, n ¼ 51 (14/37, M/F) Berberine, n ¼ 71 (35/36, M/F); placebo, n ¼ 70 (35/35, M/F)

Does, frequency, duration 500 mg/day, once daily, 12 weeks 500 mg/day, once daily, 4 weeks 0.2 g/d, once daily, 12 weeks 500 mg/d, once a day, 12 weeks 1 g/d, twice daily, 3 months

AP-1, n ¼ 29 (20/9, M/F); placebo, n ¼ 30 (18/12, M/F) AP-1, n ¼ 152 (62/90, M/F); compared to baseline

500 mg/d, once daily, 18 weeks 500 mg/d, once daily, 6 months

Menopausal women with moderate dyslipidemia

Berberine + isoflavones, n ¼ 60; compared to baseline

Dyslipidemic patients

AP-1, n ¼ 933 (416/518, M/F); placebo, n ¼ 818 (384/434, M/F)

Berberine and isoflavones combination, 12 weeks 500 mg/d, once daily, 16 weeks

Elderly (>75 years) hypercholesterolemic patients Hypercholesterolemic patients

AP-1, n ¼ 40 (21/19, M/F); placebo, n ¼ 40 (20/20, M/F) AP-1, n ¼ 25 (13/12, M/F); placebo, n ¼ 25 (13/12, M/F)

500 mg/d, once daily, 12 months 500 mg/d, once daily, 6 weeks

Patients with metabolic syndrome Hypercholesterolemic patients

Effect on plasma lipid profile Decreasing TC and LDL-C LDL-C (26%) Non-HDL-C (15%), LDL-C (19%) TC (5%), LDL-C (7.8%) T-C (11.6%), LDL-C (16.4%), HDL-C (+9.1%) T-C (15%), LDL-C (23%) TC (24%), LDL-C (32%), non-HDL-C (30%), TG (20%) T-C (14%), LDL-C (12%), TG (19%) T-C (10%), LDL-C (13%), TG (7%), HDL-C (+8%) T-C (20%), LDL-C (31%) T-C (17%), LDL-C (23%)

Ref. Feinberg et al. (2019) D’Addato et al. (2017) Spigoni et al. (2017) Sola et al. (2014) Derosa et al. (2013)

Affuso et al. (2012) Pisciotta et al. (2012)

Cianci et al. (2012)

Trimarco et al. (2011)

Marazzi et al. (2011) Affuso et al. (2010) (continued)

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

91

Table 2 (continued)

Subject Hypercholesterolemic patients

Treatment group Berberine, n ¼ 24; compared to baseline (no sex ratio provided)

Does, frequency, duration 1 g/d, twice daily, 2 months

Moderate dyslipidemic subjects

Berberine, n ¼ 20 (8/12, M/F); AP-1, n ¼ 20 (8/12, M/F)

500 mg/d, once daily, 4 weeks

Hypercholesterolemic patients

Berberine, n ¼ 63 (35/28, M/F); placebo, n ¼ 28 (17/11, M/F)

1 g/d, twice daily, 3 months

Effect on plasma lipid profile TC (21.8%), LDL-C (23.8%), TG (22.1%) T-C (16%), LDL-C (20%), TG (22%), HDL-C (+7%); berberine and AP-1 did not differ T-C (29%), LDL-C (25%), TG (35%)

Ref. Kong et al. (2008)

Cicero et al. (2007)

Kong et al. (2004)

a

Berberine-containing nutraceutical: berberine 200 mg, chitosan 10 mg, monacolin K 3 mg, and CoQ10 10 mg b AP-1: 1 tablet contains berberine 500 mg, red yeast rice extract 200 mg (equivalent to 3 mg monacolins), policosanol 10 mg, CoQ10 2 mg, folic acid 0.2 mg, and astaxanthin 0.5 mg

protective activities and optimize intestinal berberine absorption (Guarino et al. 2017). A meta-analysis of 19 controlled and cross-sectional trials showed that a combination of berberine with silymarin could significantly improve the cholesterollowering effect of berberine (Bertuccioli et al. 2019). Results from a randomized double-blind, placebo-controlled clinical trial on patients with moderate hypercholesterolemia (LDL-C up to 130–190 mg/dL) without cardiovascular disease demonstrated that, after 2 months treatment, berberine combined with a dry extract of artichoke, a lipid-lowering agent, could exert a significant reduction in plasma TC by 19%, LDL-C by 16%, and non-HDL-C by 19%, with no significant side effect (Cicero et al. 2019). Furthermore, it has been shown that the oral administration of berberine alone (500 mg/day) or with a combination of lipid-lowering nutraceuticals (consisting policosanols, red yeast extract, folic acid, coenzyme Q10, and astaxanthin) for 4 weeks effectively reduced the plasma levels of LDL-C (by 20% and 25%, respectively) and total cholesterol (by 16% and 20%, respectively) in 40 subjects with moderate dyslipidemia (Cicero et al. 2007). A multicenter, randomized, doubleblind, placebo-controlled trial on patients with mild-to-moderate hypercholesterolemia showed that 4 weeks treatment with a daily oral dose of berberine-continuing nutraceutical supplement (red yeast rice, coenzyme Q10, and hydroxytyrosol) had a favorable efficacy and safety profile, wherein LDL-C was decreased by 26%

92

A. Fatahian et al.

(D’Addato et al. 2017). Another double-blind placebo-controlled study showed that 6 weeks oral administration of the mentioned combination of berberine and nutraceuticals could decrease plasma levels of LDL-C by 23% and TC by 19% without adverse effects in 25 hypercholesterolemic patients (Affuso et al. 2010). Moreover, 12-month treatment with this combination was found to reduce the plasma levels of TC by 20% and LDL-C by 31% in elderly hypercholesterolemic patients (more than 75 years) who were statin intolerance (Marazzi et al. 2011). The combination containing berberine was also compared with ezetimibe drug in hypercholesterolemic subjects who were intolerant or refusing to take statin; the combination therapy was more effective in decreasing LDL-C (32% vs  25%) and TC (24% vs  19%) (Pisciotta et al. 2012). Several clinical trials have also investigated the effects of berberine combined with statins in hypercholesterolemic patients. A recent comprehensive meta-analysis indicates that berberine treatment could synergistically increase cholesterol-lowering potential of statins, while the incidence of statin-related adverse reactions, such as elevation of transaminase and muscle aches, was markedly decreased (Zhang et al. 2019). Cholesterol-lowering efficacy of berberine is further verified by the other clinical trials in dyslipidemic subjects with various disease conditions, such as prediabetic (Feinberg et al. 2019) and diabetic patients (Bertuccioli et al. 2019; Yin et al. 2008; Zhang et al. 2010), patients with metabolic syndrome (Affuso et al. 2012; PérezRubio et al. 2013), and post-menopause subjects (Affuso et al. 2012). These findings confirm that berberine improves plasma levels of cholesterol in dyslipidemic conditions, and since dyslipidemia, particularly high level of plasma LDL-C, is an independent risk factor for promotion and progression of atherosclerotic lesions, this natural compound possesses potential preventive and/or therapeutic effects for treating atherosclerosis. Hence, given that atherosclerosis is a chronic disorder, further rigorous clinical studies are needed to better determine the longterm efficacy of berberine and its effects on the progression of atherosclerotic lesions.

2.3

Mechanisms Underlying the Cholesterol-Lowering Effect of Berberine

Berberine is known to reduce plasma cholesterol through modulating cholesterol metabolism and hemostasis. Berberine has been shown to suppress cholesterol biosynthesis through inhibiting 3-hydroxy-3-methyl-glutaryl-coenzyme A (HMG-CoA) reductase via activating AMP-activated protein kinase (AMPK) in hepatocytes. The activated AMPK phosphorylates the rate-limiting enzyme HMG-CoA reductase and leads to the inactivation of cholesterol biosynthesis (Brusq et al. 2006).

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

93

Fig. 1 Schematic view of mechanisms underlying LDLR-dependent cholesterol lowering effects of berberine. (a) Berberine can enhance LDLR expression through elevating LDLR mRNA stability via inducing ERK signaling activity that suppresses binding of heterogeneous nuclear ribonucleoprotein 1 (hnRNP I) to AREs, leading to inhibition of RNA degradation machinery. (b) Berberine can also induce LDLR expression through inducing the degradation of two cellular trans-activators, hepatocytes nuclear factor 1α (HNF1α) and sterol regulatory element-binding protein 2 (SREBP2)

Moreover, intestinal absorption of dietary cholesterol has an important role in cholesterol hemostasis, and berberine has been shown to inhibit various processes involved in intestinal absorption, including intraluminal cholesterol micellization, cholesterol uptake by enterocytes, and cholesterol esterification in the enterocytes via reducing expression of acyl-coenzyme A cholesterol acyltransferase-2 (Wang et al. 2014). Plasma cholesterol is mainly removed from the blood circulation by the liver LDL receptor (LDLR). LDLR deficiency increases plasma cholesterol and accelerates atherosclerosis (Umans-Eckenhausen et al. 2002; Xu and Weng 2020). Recently, an in vivo study showed that berberine treatment failed to decrease the plasma levels of cholesterol in hypercholesterolemic LDLR-deficient mice (Ldlr/), providing direct evidence that supports cholesterol-lowering effects of berberine might be mediated, at least in part, by regulating the liver LDLR (Singh and Liu 2019). Further mechanistic studies (Abidi et al. 2005, 2006; Dong et al. 2015; Kong et al. 2006; Li et al. 2009a; Momtazi et al. 2017) indicate that berberine can increase stability of LDLR at both mRNA (Fig. 1a) and protein (Fig. 1b) levels and, thereby, upregulate the expression of hepatic LDLR. The 30 -untranslated region (30 -UTR) of human LDLR mRNA contains three AU-rich elements (AREs) responsible for rapid mRNA turnover, which mediates berberine-induced mRNA stabilization. Berberine was found to extend LDLR mRNA half-life entirely through 30 UTR in an extracellular receptor-activated kinase (ERK) cascade-dependent manner, in which inhibit interactions of cis-regulatory sequences of 30 UTR and mRNA binding proteins that are downstream effectors of this signaling pathway (Abidi et al. 2005, 2006; Kong et al. 2006). AREs are the well-known RNA cis-regulatory elements that play a

94

A. Fatahian et al.

central role in the mRNA stability (Zhang et al. 2002). Further studies highlighted that berberine can inhibit binding of heterogeneous nuclear ribonucleoprotein 1 (hnRNP I), a destabilizing ARE binding protein, to the LDLR mRNA 30 UTR, whereby enhances LDLR mRNA stability resulting in increased LDLR expression and improved plasma LDL-C clearance (Li et al. 2009b; Singh et al. 2014). Besides, berberine also can increase the stability of LDLR protein on the surface of hepatocytes through post-translational regulation via modulatory effects on PCSK9/LDLR pathway (Cameron et al. 2008; Momtazi et al. 2017; Xiao et al. 2012). PCSK9 is mainly produced by hepatocytes and secreted to the bloodstream where it binds the extracellular domain of hepatic LDLR and targets it to lysosomal degradation, leading to insufficient hepatic LDLR to trap plasma circulating LDL-C (Qian et al. 2007), causing hypercholesterolemia and atherosclerosis-related cardiovascular disease (Abifadel et al. 2003). Mechanistically, berberine can decrease expression of PCSK9 through accelerating the degradation of two cellular transactivators, hepatocytes nuclear factor 1α (HNF1α) and sterol regulatory elementbinding protein 2 (SREBP2) that are essential for PCSK9 expression (Dong et al. 2015; Li et al. 2009a; Momtazi et al. 2017). To sum up, berberine can increase the liver LDLR at both mRNA and protein levels through two distinct mechanisms including inhibition of hnRNPI activity and reduction of PCSK9 expression, respectively. Cholesterol is finally eliminated from the liver via the conversion into bile acids and/or excretion as free cholesterol into bile, resulting in more cholesterol uptake by the liver from the bloodstream. Berberine was found to increase cholesterol secretion from the liver into bile (Guo et al. 2016; Li et al. 2015) and also elevate hepatic expression of enzymes regulating the synthesis of bile acid from cholesterol, such as mitochondrial sterol 27-hydroxylase, which was associated with a significant reduction of plasma cholesterol (Wang et al. 2010). In conclusion, berberine can regulate plasma cholesterol through inhibiting cholesterol biosynthesis by the liver together with intestinal absorption of dietary cholesterol, elevating liver LDLR stability, and enhancing cholesterol excretion via increasing biosynthesis of bile acid and secretion of free cholesterol.

3 Berberine’s Effects on Atherosclerosis Lesion Progression: In Vitro Evidence 3.1

Berberine Improves Endothelial Cell Function

Vascular endothelial dysfunction is known to be a primary event promoting initiation and progression of atherosclerotic lesions, thus improving or repairing endothelium functionality that can exert beneficial effects on the atherosclerotic process (Mordi and Tzemos 2014). Endothelial dysfunction is characterized by elevated oxidative stress and decreased nitric oxide (NO) synthesis/availability, leading to

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

95

Fig. 2 Improved functions of berberine-treated endothelial cells

proinflammatory and proliferative processes that assist all steps of atherogenesis progression (Yuyun et al. 2018). Berberine was found to protect against endothelial dysfunction in both the cultured human umbilical vein endothelial cells (HUVECs) and blood vessels isolated from rat aorta. This protective effect of berberine is through activating AMPK signaling that upregulates the expression of endothelial nitric oxide synthase (eNOS) and downregulates the expression of NADPH oxidase (NOX4) resulting in decreased generation of reactive oxygen species (ROS) and increased production of NO (Wang et al. 2009; Zhang et al. 2013). The activated endothelial cells express adhesion molecules, such as E- and P-selectins, intercellular adhesion molecules (ICAM) like ICAM-1 and vascular cell adhesion molecules (VCAM) like VCAM-1, which act as a receptor for recruitment of the inflammatory monocytes (Collins et al. 2000; Dong et al. 1998; Shih et al. 1999). The improved functionality of injured endothelial cells by berberine treatment can be further supported by other studies showing berberine suppressed the production of ICAM-1 and VCAM-1, leading to a decrease adhesion of monocytes to endothelial cells (Chen et al. 2014; Ko et al. 2007; Wu et al. 2012). To sum up, berberine can improve endothelial cell function through balancing NO and ROS production via modulating AMPK/eNOS/NOX4 signaling pathway together with reducing the adhesion capacity to monocytes (Fig. 2).

3.2

Berberine Modulates Atherogenic Activities of Macrophages

Macrophages, derived from recruited monocytes, play a central role in atherosclerotic plaque formation and progression. After adhesion, monocytes subsequently infiltrate into the subendothelial layer and differentiate into macrophages. When these cells transmigrate across the endothelial monolayer into the intima, they uptake ox-LDL-C and form the so-called foam cells (Moore et al. 2013; Moore and Tabas 2011). Berberine was shown to decrease the expression of the infiltration marker CD68, revealing decreased macrophage infiltration into aortic plaques (Chen et al. 2014). Mechanistically, berberine can suppress transendothelial migration through reducing the toll-like receptor (TLR)-mediated macrophage migration ability via inhibiting

96

A. Fatahian et al.

the enzymatic activity of Src that is an inducible tyrosine kinase having a pivotal role in cell movement (Cheng et al. 2015). It is further approved by the other study that showed berberine treatment could significantly and highly reduce macrophages contents in mice’s atherosclerotic plaques (Yang et al. 2020). The impaired cholesterol efflux and the excessive internalization of LDL-C lead to cholesterol accumulation in the infiltrated macrophages, causing foam cell formation and atherosclerosis lesion progression (Khera et al. 2011). Reverse cholesterol transporters, such as class B scavenger receptor type I (SR-BI) and ATP-binding cassette transporter (ABC)A1 and G1 (ABCG1), play crucial roles in the efflux of intracellular cholesterol in foam cells (Ohashi et al. 2005). Upon macrophage cholesterol loading, autophagy enhances the lysosomal hydrolysis of stored cholesterol droplets and generates free cholesterol mainly for ABCA1mediated efflux, whereby it facilitates cholesterol efflux (Ouimet et al. 2011; Razani et al. 2012; Wang et al. 2016a). During atherogenic states, the autophagy process is imparted in macrophages, and berberine treatment was found to enhance autophagy in macrophages, prevent autophagy resistance in the foam cells, and consequently induce cholesterol efflux in both treated cells. Berberine could promote autophagydependent cholesterol efflux through suppression of the PI3K/AKT/mTOR pathway resulting in ABCA1-dependent cholesterol efflux in macrophage-derived foam cells (Kou et al. 2017). The positive effect of berberine on cholesterol efflux is further confirmed by other studies that show berberine can increase expression of SR-BI (Chi et al. 2014) and ABCA1/G1 via activating the liver X receptor α (LXRα) transcription factor (Lee et al. 2010; Yang et al. 2020). Meanwhile, autophagy dysfunction and consequent intracellular lipid accumulation induces macrophage-mediated inflammation and thus exacerbates atherosclerotic development (Razani et al. 2012). Berberine was shown to inhibit oxLDLinduced inflammatory factors, such as macrophage inflammatory protein 1 alpha (MIP-1α) and RANTES in macrophages, through inducing autophagy that was mediated by activation of the AMPK/mTOR signaling pathway (Fan et al. 2015). Berberine was also found to decrease macrophage inflammation through reducing the expression and secretion of tumor necrosis factor-alpha (TNF-α), monocyte chemoattractant protein-1 (MCP-1), and Interleukin 6 (IL-6), in vitro (Chen et al. 2008). Potential inhibitory effect of berberine on atherosclerotic plaque progression can be supported by other studies that show therapeutic approaches enhancing autophagy in macrophages potentially prevent or treat atherosclerosis progression (Maiuri et al. 2013; Schrijvers et al. 2011). Besides, berberine was also demonstrated to inhibit cholesterol accumulation and foam cell formation promoted by the influx of either native or oxLDL-C in macrophages. These cells can internalize native LDL-C through a receptor-independent route called macropinocytosis that has been demonstrated to promote the formation of foam cells and suggests a new therapeutic target to decrease cholesterol accumulation in atherosclerotic plaques (Kruth 2013). Of note, berberine was found to inhibit macropinocytosis and, thereby, decrease cholesterol accumulation and corresponding negative feedbacks promoted by foam cell formation (Zimetti et al. 2015). Macrophages also internalize oxLDL-C using scavenger receptors class A SR

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

97

(SR-A), CD36, and lectin-like oxidized LDL receptor-1 (LOX-1), and berberine treatment could reduce the elevated expression of the mentioned receptors and, thereby, decrease cholesterol influx and lipid content in oxLDL-exposed macrophages/foam cells (Chi et al. 2014; Guan et al. 2010; Yang et al. 2020). Importantly, a plaque with a large lipid core and covered by a thin fibrous cap is at a higher risk for rupture (Schaar et al. 2004). Since the atherosclerotic plaque rupture followed by thrombus formation causes myocardial infarction, stroke, and death (Carr et al. 1996), suppressing the rupture of unstable plaques are crucial to prevent emergency medical conditions. The ruptured fibrous cap is found to be rich in macrophage and foam cells generating matrix metalloproteinases (MMPs) and extracellular MMP inducer (EMMPRIN), which digest extracellular matrix proteins and weaken the fibrous cap, resulting in the vulnerability of atherosclerotic plaques (Gough et al. 2006; Ha et al. 2020; Perrucci et al. 2020; Schmidt et al. 2006). Berberine treatment was indicted to decrease elevated expression of MMPs and EMMPRIN in oxLDL-exposed macrophages (Chen et al. 2014; Huang et al. 2012; Huang et al. 2011), suggesting that berberine can stabilize vulnerable plaques. To sum up, berberine has potential to prevent or treat atherosclerotic plaque formation and progression, at least in part, through suppressing macrophage migration into the intima, inhibiting foam cell formation and inflammation via suppressing intracellular cholesterol accumulation through enhancing autophagy and cholesterol efflux and reducing cholesterol influx, as well as increasing plaque stability (Fig. 3).

3.3

Berberine Blocks Proliferation and Migration of Vascular Smooth Muscle Cells

Vascular smooth muscle cells (VSMCs) are a major cell type contributing to atherosclerotic lesion progression. Proliferation and migration of VSMCs from media to intima layers of vascular wall and formation of muscle-origin foam cells after lipid ingestion promote the intimal thickening and, thereby, exert an important role in atherosclerosis pathology (Basatemur et al. 2019; Bennett et al. 2016). Therefore, inhibiting VSMCs proliferation and migration can beneficially affect atherosclerotic plaque initiation and progression. Arterial injury is known to promote cell cycle re-entry in VSMCs, following a wave of immediate early gene expression. After the injury, activation of extracellular signal-regulated kinase (ERK) is among the earliest of biochemical changes and is strongly associated with subsequent expression of early response transcription factors and growth factors (Kim et al. 1998; Koyama et al. 1998; Roostalu and Wong 2018). ERK is a key transducer of extracellular signals that induce cell growth and movement, which are essential for the promotion and development of vascular lesions (Lu et al. 2020). ERK transmits mitogenic signals by translocation to the nucleus where activates many of its substrates, such as early growth response factor

Fig. 3 Alleviating effects of berberine on atherogenic function of macrophages. Berberine can inhibit intracellular cholesterol accumulation and foam cell formation through inducing autophagy via suppressing AMPK/PI3K/AKT/mTOR signaling pathway. The autophagy activation leads to the lysosomal

98 A. Fatahian et al.

hydrolysis of cholesterol droplets and the increased ATP-binding cassette transporter A1 (ABCA1) and ABCG1 activity, resulting in the enhanced cholesterol efflux activity. Activated autophagy can also suppress oxLDL-induced inflammatory factors, such as macrophage inflammatory Protein 1 alpha (MIP-1α) and RANTES in macrophages. Berberine can also increase expression of ABCA1 and class B scavenger receptor type I (SR-BI) through inducing the activity of LXR-α transcription factor. Besides, berberine can also reduce intracellular cholesterol accumulation through reducing the expression of cholesterol influx receptors, including class A scavenger receptor (SR-A), CD38, and LOX-1. Berberine mediates stabilization of atherosclerotic plaques through reducing the macrophage-produced matrix metalloproteinase-9 (MMP-9) via inhibiting activity of extracellular MMP inducer (EMMPRIN)

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . . 99

100

A. Fatahian et al.

1 (Egr-1) (Gille et al. 1995; Goetze et al. 1999). Egr-1 is a zinc finger transcription factor that is lowly expressed in the normal vessel wall but is rapidly and transiently expressed in VSMCs in response to arterial injury (Khachigian et al. 1996; Kim et al. 1995). The activated Egr-1 regulates the expression of several genes participating in cell growth and differentiation, such as platelet-derived growth factor (PDGF) (Havis and Duprez 2020; Khachigian et al. 1996; Khachigian et al. 1995; Silverman et al. 1997) and Cyclins (Guillemot et al. 2001; Havis and Duprez 2020; Yan et al. 1997), which are known to be implicated in the pathogenesis of atherosclerosis (Guillemot et al. 2001; Havis and Duprez 2020; Khachigian and Collins 1997; Yan et al. 1997). Berberine has been found to abolish injury-induced VSMC regrowth through the inactivation of ERK and inhibiting the expression of Egr-1 and downstream production of growth mediators Cyclin D1 and PDGF-A, thereby preventing early signaling related with cell cycle re-entry induced by injury. Of note, the inhibitory effect of berberine was identified to be by suppressing the activity of mitogen-activated protein kinase 1/2 (MEK1/2) that mediates activation of ERK/ Egr-1 signaling by injury (Liang et al. 2006). Moreover, PDGF is an important growth factor released after vascular injury and is related to VSMC proliferation and migration. Berberine was shown to inhibit PDGF-stimulated VSMC proliferation through activating AMPK/p53/p21Cip1 signaling and inactivation the Rac1/Cyclin D/Cyclin-dependent kinase (Cdk) leading to G1 arrest (Liang et al. 2008). AMPK is a serine/threonine protein kinase that is known to suppress PDGF-induced proliferation in human aortic VSMC (Igata et al. 2005). The activated AMPK can induce a cell cycle G1 arrest via AMPK-dependent phosphorylation of p53 in human VSMCs (Igata et al. 2005). Inhibition of PDGFinduced VSMC proliferation by berberine was indicated to be mediated in part through increasing the activity of AMPK, which led to the activation of p53 phosphorylation and up-regulation of the Cdk inhibitor p21Cip1 (Liang et al. 2008). Furthermore, Rac1 is a signaling GTPase that regulates a wide variety of cellular activities, including cell proliferation, migration, and apoptosis. Rac1 is known to mediate PDGF-induced VSMC proliferation through activating the production of cell cycle regulatory molecules Cyclins and Cdks that are responsible for G1/S cell cycle transition (Liang et al. 2008). Of note, berberine was shown to inhibit PDGF-induced Rac1 activation and upregulation of Cyclin D1, Cyclin D3, Cdk2, and Cdk4, whereby it induces G1-phase arrest in VSMCs (Liang et al. 2008). In addition to promoting cell proliferation, PDGF can also stimulate migration in VSMCs, as PDGF is the most potent reported chemoattractant for VSMCs (Zhang et al. 2018). Many reports show PDGF increases both Rac1 activity and cell migration (Al-Koussa et al. 2020). Herewith, berberine was also found to PDGFinduced VSMCs migration through inhibiting activation of Rac1 and another GTPase, Cdc42, that mediate the migratory function of PDGF through VSMCs injury (Liang et al. 2008). Regarding the mechanism of berberine on the inhibition of Rac1 and Cdc42, there has been evidence that AMPK activation could lead to suppression of HMG-CoA reductase, the rate-limiting enzyme of cholesterol synthesis (Liang et al. 2008). Suppression of HMG-CoA reductase decreases cholesterol synthesis as well as

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

101

some critical isoprenoids downstream of mevalonate such as farnesyl pyrophosphate and geranylgeranyl pyrophosphate, which are essential for membrane translocation and activation of Rac1 and Cdc42 (Liang et al. 2008). Therefore, AMPK activation by berberine can drive suppression of HMG-CoA reductase and reduce downstream isoprenoids that are required for Rac1 and Cdc42. Hence, berberine can indirectly inhibit activation of Rac1 and Cdc42 and prevent cell proliferation and migration induced by PDGF in VSMCs. In conclusion, berberine-elicited anti-proliferative and anti-migratory effects in the injured VSMCs are related to a multifaceted attack on multiple signaling targets that critically participated in growth suppression. Given that abnormal proliferation and migration of VSMCs is a hallmark event in atherosclerotic lesions development, such findings can support the preventive effect of berberine on initiation and progression of atherosclerosis lesions.

4 Berberine’s Effects on Atherosclerosis Lesion Progression: In Vivo Evidence There are several in vivo studies on mouse models of atherosclerosis that support the abovementioned in vitro studies showing potential anti-atherogenic effects of berberine. ApoE/ mouse fed with the atherogenic diet is the most common model of human atherosclerosis that is widely used to evaluate the preventive or therapeutic effects of drugs on atherosclerosis lesion progression. It was found that oral gavage of berberine at a low dose (50 mg/kg/day) and a high dose (100 mg/kg/day) after 13 weeks could significantly decrease atherosclerotic plaque area and lesion size in ApoE/ mice bearing atherosclerosis, in which the high dose was found to be more effective (Wu et al. 2020). Further study reveals that berberine treatment (8 weeks) could decrease vascular inflammation and oxidative stress via an AMPK-dependent mechanism, leading to a significant reduction of atherosclerotic plaque progression in mouse aorta (Wang et al. 2011). However, another in vivo study indicated that berberine treatment at a lower dose (orally 10 mg/kg/day, for 16 weeks) failed to exert a significant effect on atherosclerosis plaque progression in hypercholesterolemic ApoE/ mice, while the same dose of its derivatives, dihydroberberine and 8,8-dimethyldi-hydroberberine possessing higher bioavailability, were found to at the same dose effectively decrease atherosclerotic plaque size and improve plaque stability (Chen et al. 2014). This finding was suggested to be due to poor intestinal absorption of berberine (Chen et al. 2014). Thus, oral administrating of berberine derivatives with improved intestinal absorption or its administration via alternative routes can enhance therapeutic potential. It can be further supported by a recent in vivo study that showed berberine (10 mg/kg/ day, for 14 weeks) via intraperitoneal injection in hypercholesterolemic ApoE/ mice could decrease the size of atherosclerotic plaque in aortic sinus by 45% (Yang et al. 2020).

102

A. Fatahian et al.

Cellular and molecular examination of atherosclerosis lesions isolated from berberine-treated mice reveals that berberine can suppress atherosclerosis plaque progression through mechanisms consistent with those found by in vitro studies discussed in previous sections. Of note, berberine has been indicated to reduce macrophage content in mice’s atherosclerotic plaques through reducing macrophage adhesion and infiltration to the vascular endothelium via inhibiting the expression of two critical adhesion molecules ICAM-1 and VCAM-1 (Chen et al. 2014; Yang et al. 2020). Moreover, berberine treatment was shown to reduce lipids contents in aortic macrophages through increasing expression of ABCA1/G1-mediated cholesterol efflux and reducing expression of SR-A/CD36-mediated cholesterol influx in hypercholesterolemic ApoE/ mice (Yang et al. 2020). Likewise, immunohistochemical assays of mice’s aortic plaque showed that berberine treatment decreased protein expression of MMP-9 and EMMPRIN and whereby increased collagen content and, consequently, improved fibrous cap thickness and plaque stability (Chen et al. 2014). Overall, these results indicate that berberine inherently has the potential to protect and ameliorate the progression of atherosclerotic plaques. Since there is no clinically approved lipid-lowering agent that can target atherosclerotic plaques and locally inhibit the function of cellular and molecular mediators in lesion progression, these findings underscore the therapeutic importance of berberine for preventing and treating atherosclerosis. However, low bioavailability of berberine can influence such beneficial effects, and, therefore, further preclinical studies are required to determine its effectual derivatives or administration routes, such as intravenous injection, for providing an efficient anti-atherosclerotic therapeutic tool.

5 Conclusion Berberine is found to be a promising anti-atherogenic agent protecting against atherosclerosis progression, as it can modulate cholesterol metabolism and hemostasis and the other important atherogenic factors directly involved in atherosclerotic lesion formation. The majority of findings consistently demonstrate that the lipidlowering effect of berberine is mainly in term of reducing TC and LDL-C through several LDLR-dependent and LDLR -independent mechanisms. Berberine decreases plasma levels of LDL-C through upregulating the mRNA and protein expression of the hepatic LDLR via two distinct mechanisms, including evaluation of LDLR mRNA stability and inhibition of PCSK9, a main negative regulator of LDLR protein. Besides the LDLR-mediated mechanism, berberine also can impact cholesterol hemostasis through regulating the other important factors, including liver biosynthesis of cholesterol, intestinal absorption of dietary cholesterol, as well as bile acid synthesis and secretion, and secretion of free cholesterol. Berberine has been also found to have anti-inflammatory and anti-oxidant abilities, whereby it can modulate function and proliferation of inflammatory macrophages, VSMC, and endothelial cells that directly collaborate in atherosclerotic lesion formation.

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

103

Among the signaling pathways through which berberine regulates intracellular processes, AMPK has a central and critical role. Interestingly, activated AMPK contributes to various mechanisms involved in the atheroprotective effect of berberine, including suppressing cholesterol biosynthesis, improving endothelial dysfunction via upregulating the expression of eNOS and downregulating the expression of NOX4, as well as blocking injury-stimulated VSMCs regrowth and proliferation. Acknowledgments The authors appreciate the cooperation of Departments of Medical Immunology and Medical Biotechnology of Mashhad University of Medical Sciences. Conflicts of Interest The authors declare that there are no conflicts of interest and financial support for the present review article.

References Abidi P, Zhou Y, Jiang J-D, Liu J (2005) Extracellular signal-regulated kinase–dependent stabilization of hepatic low-density lipoprotein receptor mRNA by herbal medicine berberine. Arterioscler Thromb Vasc Biol 25(10):2170–2176 Abidi P, Chen W, Kraemer FB, Li H, Liu J (2006) The medicinal plant goldenseal is a natural LDL-lowering agent with multiple bioactive components and new action mechanisms. J Lipid Res 47(10):2134–2147 Abifadel M, Varret M, Rabès J-P, Allard D, Ouguerram K, Devillers M et al (2003) Mutations in PCSK9 cause autosomal dominant hypercholesterolemia. Nat Genet 34(2):154–156 Affuso F, Ruvolo A, Micillo F, Saccà L, Fazio S (2010) Effects of a nutraceutical combination (berberine, red yeast rice and policosanols) on lipid levels and endothelial function randomized, double-blind, placebo-controlled study. Nutr Metab Cardiovasc Dis 20(9):656–661 Affuso F, Mercurio V, Ruvolo A, Pirozzi C, Micillo F, Carlomagno G et al (2012) A nutraceutical combination improves insulin sensitivity in patients with metabolic syndrome. World J Cardiol 4(3):77 Al-Koussa H, El Atat O, Jaafar L, Tashjian H, El-Sibai M (2020) The role of Rho GTPases in motility and invasion of glioblastoma cells. Anal Cell Pathol 2020 Ayati SH, Fazeli B, Momtazi-Borojeni AA, Cicero AFG, Pirro M, Sahebkar A (2017) Regulatory effects of berberine on microRNome in cancer and other conditions. Crit Rev Oncol Hematol 116:147–158 Basatemur GL, Jørgensen HF, Clarke MCH, Bennett MR, Mallat Z (2019) Vascular smooth muscle cells in atherosclerosis. Nat Rev Cardiol 1:18 Bennett MR, Sinha S, Owens GK (2016) Vascular smooth muscle cells in atherosclerosis. Circ Res 118(4):692–702 Bertuccioli A, Moricoli S, Amatori S, Rocchi MBL, Vici G, Sisti D (2019) Berberine and dyslipidemia: different applications and biopharmaceutical formulations without statin-like molecules—a meta-analysis. J Med Food 23(2):101–113 Briand F, Thieblemont Q, Muzotte E, Sulpice T (2013) Upregulating reverse cholesterol transport with cholesteryl ester transfer protein inhibition requires combination with the LDL-lowering drug berberine in dyslipidemic hamsters. Arterioscler Thromb Vasc Biol 33(1):13–23 Brusq J-M, Ancellin N, Grondin P, Guillard R, Martin S, Saintillan Y, Issandou M (2006) Inhibition of lipid synthesis through activation of AMP kinase: an additional mechanism for the hypolipidemic effects of berberine. J Lipid Res 47(6):1281–1288 Cameron J, Ranheim T, Kulseth MA, Leren TP, Berge KE (2008) Berberine decreases PCSK9 expression in HepG2 cells. Atherosclerosis 201(2):266–273

104

A. Fatahian et al.

Carr S, Farb A, Pearce WH, Virmani R, Yao JST (1996) Atherosclerotic plaque rupture in symptomatic carotid artery stenosis. J Vasc Surg 23(5):755–766 Chang XX, Yan HM, Fei J, Jiang MH, Zhu HG, Lu DR, Gao X (2010) Berberine reduces methylation of the MTTP promoter and alleviates fatty liver induced by a high-fat diet in rats. J Lipid Res 51(9):2504–2515 Chang X-x, Yan H-m, Xu Q, Xia M-f, Bian H, Zhu T-f, Gao X (2012) The effects of berberine on hyperhomocysteinemia and hyperlipidemia in rats fed with a long-term high-fat diet. Lipids Health Dis 11(1):86 Chen FL, Yang ZH, Liu Y, Li LX, Liang WC, Wang XC et al (2008) Berberine inhibits the expression of TNFα, MCP-1, and IL-6 in AcLDL-stimulated macrophages through PPARγ pathway. Endocrine 33(3):331–337 Chen J, Cao J, Fang L, Liu B, Zhou Q, Sun Y et al (2014) Berberine derivatives reduce atherosclerotic plaque size and vulnerability in apoE/ mice. J Transl Med 12(1):326 Cheng WE, Chang MY, Wei JY, Chen YJ, Maa MC, Leu TH (2015) Berberine reduces toll-like receptor-mediated macrophage migration by suppression of Src enhancement. Eur J Pharmacol 757:1–10. https://doi.org/10.1016/j.ejphar.2015.03.013 Chi L, Peng L, Hu X, Pan N, Zhang Y (2014) Berberine combined with atorvastatin downregulates LOX1 expression through the ET1 receptor in monocyte/macrophages. Int J Mol Med 34 (1):283–290. https://doi.org/10.3892/ijmm.2014.1748 Chueh W-H, Lin J-Y (2011) Berberine, an isoquinoline alkaloid in herbal plants, protects pancreatic islets and serum lipids in nonobese diabetic mice. J Agric Food Chem 59(14):8021–8027 Cianci A, Cicero AFG, Colacurci N, Matarazzo MG, Leo D, Vincenzo (2012) Activity of isoflavones and berberine on vasomotor symptoms and lipid profile in menopausal women. Gynecol Endocrinol 28(9):699–702 Cicero AFG, Rovati LC, Setnikar I (2007) Eulipidemic effects of berberine administered alone or in combination with other natural cholesterol-lowering agents. Arzneimittelforschung 57 (01):26–30 Cicero AF, Fogacci F, Bove M, Giovannini M, Veronesi M, Borghi C (2019) Short-term effects of dry extracts of artichokeand berberis in hypercholesterolemic patients without cardiovascular disease. Am J Cardiol 123(4):588–591 Collins RG, Velji R, Guevara NV, Hicks MJ, Chan L, Beaudet AL (2000) P-selectin or intercellular adhesion molecule (ICAM)-1 deficiency substantially protects against atherosclerosis in apolipoprotein E–deficient mice. J Exp Med 191(1):189–194 D’Addato S, Scandiani L, Mombelli G, Focanti F, Pelacchi F, Salvatori E et al (2017) Effect of a food supplement containing berberine, monacolin K, hydroxytyrosol and coenzyme Q10 on lipid levels: a randomized, double-blind, placebo controlled study. Drug Des Devel Ther 11:1585 Derosa G, D’Angelo A, Bonaventura A, Bianchi L, Romano D, Maffioli P (2013) Effects of berberine on lipid profile in subjects with low cardiovascular risk. Expert Opin Biol Ther 13 (4):475–482 Doggrell SA (2005) Berberine--a novel approach to cholesterol lowering. Expert Opin Investig Drugs 14(5):683–685. https://doi.org/10.1517/13543784.14.5.683 Dong ZM, Chapman SM, Brown AA, Frenette PS, Hynes RO, Wagner DD (1998) The combined role of P-and E-selectins in atherosclerosis. J Clin Invest 102(1):145–152 Dong H, Zhao Y, Zhao L, Lu F (2013) The effects of berberine on blood lipids: a systemic review and meta-analysis of randomized controlled trials. Planta Med 79(06):437–446 Dong B, Li H, Singh AB, Cao A, Liu J (2015) Inhibition of PCSK9 transcription by berberine involves down-regulation of hepatic HNF1alpha protein expression through the ubiquitinproteasome degradation pathway. J Biol Chem 290(7):4047–4058. https://doi.org/10.1074/ jbc.M114.597229 Fan X, Wang J, Hou J, Lin C, Bensoussan A, Chang D et al (2015) Berberine alleviates ox-LDL induced inflammatory factors by up-regulation of autophagy via AMPK/mTOR signaling pathway. J Transl Med 13:92–92. https://doi.org/10.1186/s12967-015-0450-z

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

105

Feinberg T, Wieland LS, Miller LE, Munir K, Pollin TI, Shuldiner AR et al (2019) Polyherbal dietary supplementation for prediabetic adults: study protocol for a randomized controlled trial. Trials 20(1):1–13 Feng W-w, Kuang S-y, Tu C, Ma Z-j, Pang J-y, Wang Y-h et al (2018) Natural products berberine and curcumin exhibited better ameliorative effects on rats with non-alcohol fatty liver disease than lovastatin. Biomed Pharmacother 99:325–333 Ference BA, Ginsberg HN, Graham I, Ray KK, Packard CJ, Bruckert E et al (2017) Low-density lipoproteins cause atherosclerotic cardiovascular disease. 1. Evidence from genetic, epidemiologic, and clinical studies. A consensus statement from the European Atherosclerosis Society Consensus Panel. Eur Heart J 38(32):2459–2472 Gille H, Kortenjann M, Thomae O, Moomaw C, Slaughter C, Cobb MH, Shaw PE (1995) ERK phosphorylation potentiates Elk-1-mediated ternary complex formation and transactivation. EMBO J 14(5):951–962 Goetze S, Xi X-P, Graf K, Fleck E, Hsueh WA, Law RE (1999) Troglitazone inhibits angiotensin IIinduced extracellular signal-regulated kinase 1/2 nuclear translocation and activation in vascular smooth muscle cells. FEBS Lett 452(3):277–282 Gough PJ, Gomez IG, Wille PT, Raines EW (2006) Macrophage expression of active MMP-9 induces acute plaque disruption in apoE-deficient mice. J Clin Invest 116(1):59–69 Guan S, Wang B, Li W, Guan J, Fang X (2010) Effects of berberine on expression of LOX-1 and SR-BI in human macrophage-derived foam cells induced by ox-LDL. Am J Chin Med 38 (06):1161–1169 Guarino G, Strollo F, Carbone L, Della TC, Letizia M, Marino G, Gentile S (2017) Bioimpedance analysis, metabolic effects and safety of the association Berberis aristata/Bilybum marianum: a 52-week double-blind, placebo-controlled study in obese patients with type 2 diabetes. J Biol Regul Homeost Agents 31(2):495–502 Guillemot L, Levy A, Raymondjean M, Rothhut B (2001) Angiotensin II-induced transcriptional activation of the cyclin D1 gene is mediated by Egr-1 in CHO-AT1A cells. J Biol Chem 276 (42):39394–39403 Guo Y, Zhang Y, Huang W, Selwyn FP, Klaassen CD (2016) Dose-response effect of berberine on bile acid profile and gut microbiota in mice. BMC Complement Altern Med 16(1):394. https:// doi.org/10.1186/s12906-016-1367-7 Ha S-H, Choi H, Park J-Y, Abekura F, Lee Y-C, Kim J-R, Kim C-H (2020) Mycobacterium tuberculosis–secreted protein, ESAT-6, inhibits lipopolysaccharide-induced MMP-9 expression and inflammation through NF-κB and MAPK signaling in RAW 264.7 macrophage cells. Inflammation 43(1):54–65 Havis E, Duprez D (2020) Egr1 transcription factor is a multifaceted regulator of matrix production in tendons and other connective tissues. Int J Mol Sci 21(5):1664 He K, Kou S, Zou Z, Hu Y, Feng M, Han B et al (2016) Hypolipidemic effects of alkaloids from Rhizoma Coptidis in diet-induced hyperlipidemic hamsters. Planta Med 82(08):690–697 Hu Y, Davies GE (2010) Berberine inhibits adipogenesis in high-fat diet-induced obesity mice. Fitoterapia 81(5):358–366 Hu Y, Ehli EA, Kittelsrud J, Ronan PJ, Munger K, Downey T et al (2012) Lipid-lowering effect of berberine in human subjects and rats. Phytomedicine 19(10):861–867 Huang Z, Wang L, Meng S, Wang Y, Chen T, Wang C (2011) Berberine reduces both MMP-9 and EMMPRIN expression through prevention of p38 pathway activation in PMA-induced macrophages. Int J Cardiol 146(2):153–158 Huang Z, Meng S, Wang L, Wang Y, Chen T, Wang C (2012) Suppression of oxLDL-induced MMP-9 and EMMPRIN expression by berberine via inhibition of NF-kappaB activation in human THP-1 macrophages. Anat Rec 295(1):78–86. https://doi.org/10.1002/ar.21489 Igata M, Motoshima H, Tsuruzoe K, Kojima K, Matsumura T, Kondo T et al (2005) Adenosine monophosphate-activated protein kinase suppresses vascular smooth muscle cell proliferation through the inhibition of cell cycle progression. Circ Res 97(8):837–844

106

A. Fatahian et al.

Imanshahidi M, Hosseinzadeh H (2008) Pharmacological and therapeutic effects of Berberis vulgaris and its active constituent, berberine. Phytother Res 22(8):999–1012 Jia X, Chen Y, Zidichouski J, Zhang J, Sun C, Wang Y (2008) Co-administration of berberine and plant stanols synergistically reduces plasma cholesterol in rats. Atherosclerosis 201(1):101–107 Johnston TP, Korolenko TA, Pirro M, Sahebkar A (2017) Preventing cardiovascular heart disease: promising nutraceutical and non-nutraceutical treatments for cholesterol management. Pharmacol Res 120:219–225 Ju J, Li J, Lin Q, Xu H (2018) Efficacy and safety of berberine for dyslipidaemias: a systematic review and meta-analysis of randomized clinical trials. Phytomedicine 50:25–34 Khachigian LM, Williams AJ, Collins T (1995) Interplay of Sp1 and Egr-1 in the proximal plateletderived growth factor A-chain promoter in cultured vascular endothelial cells. J Biol Chem 270 (46):27679–27686 Khachigian LM, Lindner V, Williams AJ, Collins T (1996) Egr-1-induced endothelial gene expression: a common theme in vascular injury. Science 271(5254):1427–1431 Khachigian LM, Collins T (1997) Inducible expression of Egr-1–dependent genes: a paradigm of transcriptional activation in vascular endothelium. Circ Res 81(4):457–461 Khera AV, Cuchel M, de la Llera-Moya M, Rodrigues A, Burke MF, Jafri K et al (2011) Cholesterol efflux capacity, high-density lipoprotein function, and atherosclerosis. N Engl J Med 364(2):127–135 Kim S, Kawamura M, Wanibuchi H, Ohta K, Hamaguchi A, Omura T et al (1995) Angiotensin II type 1 receptor blockade inhibits the expression of immediate-early genes and fibronectin in rat injured artery. Circulation 92(1):88–95 Kim S, Izumi Y, Yano M, Hamaguchi A, Miura K, Yamanaka S et al (1998) Angiotensin blockade inhibits activation of mitogen-activated protein kinases in rat balloon-injured artery. Circulation 97(17):1731–1737 Kim WS, Lee YS, Cha SH, Jeong HW, Choe SS, Lee M-R et al (2009) Berberine improves lipid dysregulation in obesity by controlling central and peripheral AMPK activity. Am J Physiol Endocrinol Metab 296(4):E812–E819 Kim M, Kim T-W, Kim C-J, Shin M-S, Hong M, Park H-S, Park S-S (2019) Berberine ameliorates brain inflammation in poloxamer 407-induced hyperlipidemic rats. Int Neurourol J 23(Suppl 2): S102 Ko YJ, Lee J-S, Park BC, Shin HM, Kim J-A (2007) Inhibitory effects of Zoagumhwan water extract and berberine on angiotensin II-induced monocyte chemoattractant protein (MCP)-1 expression and monocyte adhesion to endothelial cells. Vasc Pharmacol 47(2–3):189–196 Kong W, Wei J, Abidi P, Lin M, Inaba S, Li C et al (2004) Berberine is a novel cholesterol-lowering drug working through a unique mechanism distinct from statins. Nat Med 10(12):1344–1351 Kong W, Wei J, Abidi P, Lin M, Inaba S, Li C et al (2006) Berberine is a promising novel cholesterol-lowering drug working through a unique mechanism distinct from statins. In: Paper presented at the Atherosclerosis Supplements Kong W-J, Wei J, Zuo Z-Y, Wang Y-M, Song D-Q, You X-F et al (2008) Combination of simvastatin with berberine improves the lipid-lowering efficacy. Metabolism 57(8):1029–1037 Koppen LM, Whitaker A, Rosene A, Beckett RD (2017) Efficacy of berberine alone and in combination for the treatment of hyperlipidemia: a systematic review. J Evid Based Complementary Altern Med 22(4):956–968 Kou S, Han B, Wang Y, Huang T, He K, Han Y et al (2016) Synergetic cholesterol-lowering effects of main alkaloids from Rhizoma Coptidis in HepG2 cells and hypercholesterolemia hamsters. Life Sci 151:50–60 Kou JY, Li Y, Zhong ZY, Jiang YQ, Li XS, Han XB et al (2017) Berberine-sonodynamic therapy induces autophagy and lipid unloading in macrophage. Cell Death Dis 8(1):e2558–e2558. https://doi.org/10.1038/cddis.2016.354 Koyama H, Olson NE, Dastvan FF, Reidy MA (1998) Cell replication in the arterial wall: activation of signaling pathway following in vivo injury. Circ Res 82(6):713–721

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

107

Kruth HS (2013) Fluid-phase pinocytosis of LDL by macrophages: a novel target to reduce macrophage cholesterol accumulation in atherosclerotic lesions. Curr Pharm Des 19 (33):5865–5872 Lee TS, Pan CC, Peng CC, Kou YR, Chen CY, Ching LC et al (2010) Anti-atherogenic effect of berberine on LXRalpha-ABCA1-dependent cholesterol efflux in macrophages. J Cell Biochem 111(1):104–110. https://doi.org/10.1002/jcb.22667 Li Y-H, Yang P, Kong W-J, Wang Y-X, Hu C-Q, Zuo Z-Y et al (2008) Berberine Analogues as a novel class of the low-density-lipoprotein receptor up-regulators: synthesis, structure activity relationships, and cholesterol-lowering efficacy. J Med Chem 52(2):492–501 Li H, Chen W, Zhou Y, Abidi P, Sharpe O, Robinson WH et al (2009a) Identification of mRNA binding proteins that regulate the stability of LDL receptor mRNA through AU-rich elements. J Lipid Res 50(5):820–831. https://doi.org/10.1194/jlr.M800375-JLR200 Li H, Dong B, Park SW, Lee HS, Chen W, Liu J (2009b) Hepatocyte nuclear factor 1alpha plays a critical role in PCSK9 gene transcription and regulation by the natural hypocholesterolemic compound berberine. J Biol Chem 284(42):28885–28895. https://doi.org/10.1074/jbc.M109. 052407 Li G-S, Liu X-H, Zhu H, Huang L, Liu Y-L, Ma C-M, Qin C (2011) Berberine-improved visceral white adipose tissue insulin resistance associated with altered sterol regulatory element-binding proteins, liver X receptors, and peroxisome proliferator-activated receptors transcriptional programs in diabetic hamsters. Biol Pharm Bull 34(5):644–654 Li Z, Geng Y-N, Jiang J-D, Kong W-J (2014) Antioxidant and anti-inflammatory activities of berberine in the treatment of diabetes mellitus. Evid Based Complement Alternat Med 2014:289264 Li X-Y, Zhao Z-X, Huang M, Feng R, He C-Y, Ma C et al (2015) Effect of Berberine on promoting the excretion of cholesterol in high-fat diet-induced hyperlipidemic hamsters. J Transl Med 13 (1):278 Liang K-W, Ting C-T, Yin S-C, Chen Y-T, Lin S-J, Liao JK, Hsu S-L (2006) Berberine suppresses MEK/ERK-dependent Egr-1 signaling pathway and inhibits vascular smooth muscle cell regrowth after in vitro mechanical injury. Biochem Pharmacol 71(6):806–817. https://doi.org/ 10.1016/j.bcp.2005.12.028 Liang K-W, Yin S-C, Ting C-T, Lin S-J, Hsueh C-M, Chen C-Y, Hsu S-L (2008) Berberine inhibits platelet-derived growth factor-induced growth and migration partly through an AMPKdependent pathway in vascular smooth muscle cells. Eur J Pharmacol 590(1):343–354. https://doi.org/10.1016/j.ejphar.2008.06.034 Lu Y, Sun X, Peng L, Jiang W, Li W, Yuan H, Cai J (2020) Angiotensin II-Induced vascular remodeling and hypertension involves cathepsin L/V-MEK/ERK mediated mechanism. Int J Cardiol 298:98–106 Maiuri MC, Grassia G, Platt AM, Carnuccio R, Ialenti A, Maffia P (2013) Macrophage autophagy in atherosclerosis. Mediat Inflamm 2013:584715 Marazzi G, Cacciotti L, Pelliccia F, Iaia L, Volterrani M, Caminiti G et al (2011) Long-term effects of nutraceuticals (berberine, red yeast rice, policosanol) in elderly hypercholesterolemic patients. Adv Ther 28(12):1105–1113 Momtazi AA, Banach M, Pirro M, Katsiki N, Sahebkar A (2017) Regulation of PCSK9 by nutraceuticals. Pharmacol Res 120:157–169 Moore KJ, Tabas I (2011) Macrophages in the pathogenesis of atherosclerosis. Cell 145 (3):341–355 Moore KJ, Sheedy FJ, Fisher EA (2013) Macrophages in atherosclerosis: a dynamic balance. Nat Rev Immunol 13(10):709 Mordi I, Tzemos N (2014) Is reversal of endothelial dysfunction still an attractive target in modern cardiology? World J Cardiol 6(8):824 Ohashi R, Mu H, Wang X, Yao Q, Chen C (2005) Reverse cholesterol transport and cholesterol efflux in atherosclerosis. QJM 98(12):845–856

108

A. Fatahian et al.

Ouimet M, Franklin V, Mak E, Liao X, Tabas I, Marcel YL (2011) Autophagy regulates cholesterol efflux from macrophage foam cells via lysosomal acid lipase. Cell Metab 13(6):655–667 Pérez-Rubio KG, González-Ortiz M, Martínez-Abundis E, Robles-Cervantes JA, EspinelBermúdez MC (2013) Effect of berberine administration on metabolic syndrome, insulin sensitivity, and insulin secretion. Metab Syndr Relat Disord 11(5):366–369 Perrucci GL, Rurali E, Corlianò M, Balzo M, Piccoli M, Moschetta D et al (2020) Cyclophilin A/EMMPRIN axis is involved in pro-fibrotic processes associated with thoracic aortic aneurysm of marfan syndrome patients. Cell 9(1):154 Pisciotta L, Bellocchio A, Bertolini S (2012) Nutraceutical pill containing berberine versus ezetimibe on plasma lipid pattern in hypercholesterolemic subjects and its additive effect in patients with familial hypercholesterolemia on stable cholesterol-lowering treatment. Lipids Health Dis 11(1):123 Qian Y-W, Schmidt RJ, Zhang Y, Chu S, Lin A, Wang H et al (2007) Secreted PCSK9 downregulates low density lipoprotein receptor through receptor-mediated endocytosis. J Lipid Res 48(7):1488–1498 Razani B, Feng C, Coleman T, Emanuel R, Wen H, Hwang S et al (2012) Autophagy links inflammasomes to atherosclerotic progression. Cell Metab 15(4):534–544 Ridker PM, Revkin J, Amarenco P, Brunell R, Curto M, Civeira F et al (2017) Cardiovascular efficacy and safety of bococizumab in high-risk patients. N Engl J Med 376(16):1527–1539 Roostalu U, Wong JKF (2018) Arterial smooth muscle dynamics in development and repair. Dev Biol 435(2):109–121 Sabatine MS, Giugliano RP, Keech AC, Honarpour N, Wiviott SD, Murphy SA et al (2017) FOURIER Steering Committee and Investigators. Evolocumab and clinical outcomes in patients with cardiovascular disease. N Engl J Med 376(18):1713–1722 Sahebkar A, Watts GF (2013) New therapies targeting apoB metabolism for high-risk patients with inherited dyslipidaemias: what can the clinician expect? Cardiovasc Drugs Ther 27(6):559–567. https://doi.org/10.1007/s10557-013-6479-4 Schaar JA, Muller JE, Falk E, Virmani R, Fuster V, Serruys PW et al (2004) Terminology for highrisk and vulnerable coronary artery plaques. Eur Heart J 25(12):1077–1082 Schmidt R, Redecke V, Breitfeld Y, Wantia N, Miethke T, Massberg S et al (2006) EMMPRIN (CD 147) is a central activator of extracellular matrix degradation by Chlamydia pneumoniaeinfected monocytes. Thromb Haemost 95(01):151–158 Schrijvers DM, Meyer D, Guido RY, Martinet W (2011) Autophagy in atherosclerosis: a potential drug target for plaque stabilization. Arterioscler Thromb Vasc Biol 31(12):2787–2791 Shih PT, Brennan M-L, Vora DK, Territo MC, Strahl D, Elices MJ et al (1999) Blocking very late antigen-4 integrin decreases leukocyte entry and fatty streak formation in mice fed an atherogenic diet. Circ Res 84(3):345–351 Silverman ES, Khachigian LM, Lindner V, Williams AJ, Collins T (1997) Inducible PDGF A-chain transcription in smooth muscle cells is mediated by Egr-1 displacement of Sp1 and Sp3. Am J Phys Heart Circ Phys 273(3):H1415–H1426 Singh AB, Liu J (2019) Berberine decreases plasma triglyceride levels and upregulates hepatic TRIB1 in LDLR wild type mice and in LDLR deficient mice. Sci Rep 9(1):1–14 Singh IP, Mahajan S (2013) Berberine and its derivatives: a patent review (2009–2012). Expert Opin Ther Pat 23(2):215–231 Singh A, Duggal S, Kaur N, Singh J (2010) Berberine: alkaloid with wide spectrum of pharmacological activities. J Nat Prod 3:64–75 Singh AB, Li H, Kan CF, Dong B, Nicolls MR, Liu J (2014) The critical role of mRNA destabilizing protein heterogeneous nuclear ribonucleoprotein d in 30 untranslated regionmediated decay of low-density lipoprotein receptor mRNA in liver tissue. Arterioscler Thromb Vasc Biol 34(1):8–16. https://doi.org/10.1161/atvbaha.112.301131 Sola R, Valls R-M, Puzo J, Calabuig J-R, Brea A, Pedret A et al (2014) Effects of poly-bioactive compounds on lipid profile and body weight in a moderately hypercholesterolemic population with low cardiovascular disease risk: a multicenter randomized trial. PLoS One 9(8):e101978

Promising Anti-atherosclerotic Effect of Berberine: Evidence from In Vitro, In. . .

109

Spigoni V, Aldigeri R, Antonini M, Micheli M, Fantuzzi F, Fratter A et al (2017) Effects of a new nutraceutical formulation (berberine, red yeast rice and chitosan) on non-HDL cholesterol levels in individuals with dyslipidemia: results from a randomized, double blind, placebo-controlled study. Int J Mol Sci 18(7):1498 Srivastava S, Srivastava M, Misra A, Pandey G, Rawat AKS (2015) A review on biological and chemical diversity in Berberis (Berberidaceae). EXCLI J 14:247 Trimarco B, Benvenuti C, Rozza F, Cimmino CS, Giudice R, Crispo S (2011) Clinical evidence of efficacy of red yeast rice and berberine in a large controlled study versus diet. Mediterr J Nutr Metab 4(2):133–139 Toth PP, Patti AM, Giglio RV, Nikolic D, Castellino G, Rizzo M, Banach M (2018) Management of statin intolerance in 2018: still more questions than answers. Am J Cardiovasc Drugs 18 (3):157–173 Umans-Eckenhausen MAW, Sijbrands EJG, Kastelein JJP, Defesche JC (2002) Low-density lipoprotein receptor gene mutations and cardiovascular risk in a large genetic cascade screening population. Circulation 106(24):3031–3036 Wang Y, Zidichouski JA (2018) Update on the benefits and mechanisms of action of the bioactive vegetal alkaloid berberine on lipid metabolism and homeostasis. Cholesterol 2018:7173920 Wang Y, Huang Y, Lam KSL, Li Y, Wong WT, Ye H et al (2009) Berberine prevents hyperglycemia-induced endothelial injury and enhances vasodilatation via adenosine monophosphate-activated protein kinase and endothelial nitric oxide synthase. Cardiovasc Res 82(3):484–492 Wang Y, Jia X, Ghanam K, Beaurepaire C, Zidichouski J, Miller L (2010) Berberine and plant stanols synergistically inhibit cholesterol absorption in hamsters. Atherosclerosis 209 (1):111–117. https://doi.org/10.1016/j.atherosclerosis.2009.08.050 Wang H, Shi L, Yin H, Zhou Q (2011) Study on effect of berberine on modulating lipid and CPT IA gene expression. Zhongguo Zhong Yao Za Zhi 36(19):2715–2718 Wang FL, Tang LQ, Yang F, Zhu LN, Cai M, Wei W (2013) Renoprotective effects of berberine and its possible molecular mechanisms in combination of high-fat diet and low-dose streptozotocin-induced diabetic rats. Mol Biol Rep 40(3):2405–2418 Wang Y, Yi X, Ghanam K, Zhang S, Zhao T, Zhu X (2014) Berberine decreases cholesterol levels in rats through multiple mechanisms, including inhibition of cholesterol absorption. Metabolism 63(9):1167–1177 Wang L, Jiang Y, Song X, Guo C, Zhu F, Wang X et al (2016a) Pdcd4 deficiency enhances macrophage lipoautophagy and attenuates foam cell formation and atherosclerosis in mice. Cell Death Dis 7(1):e2055–e2055 Wang L, Peng LY, Wei GH, Ge H (2016b) Therapeutic effects of berberine capsule on patients with mild hyperlipidemia. Zhongguo Zhong Xi Yi Jie He Za Zhi 36(6):681–684 Ward NC, Watts GF, Eckel RH (2019) Statin toxicity: mechanistic insights and clinical implications. Circ Res 124(2):328–350 Wu N, Sarna LK, Siow YL, Karmin O (2010) Regulation of hepatic cholesterol biosynthesis by berberine during hyperhomocysteinemia. Am J Phys Regul Integr Comp Phys 300(3):R635– R643 Wu YH, Chuang SY, Hong WC, Lai YJ, Chang GJ, Pang JHS (2012) Berberine reduces leukocyte adhesion to LPS-stimulated endothelial cells and VCAM-1 expression both in vivo and in vitro. Int J Immunopathol Pharmacol 25(3):741–750 Wu M, Yang S, Wang S, Cao Y, Zhao R, Li X et al (2020) Effect of berberine on atherosclerosis and gut microbiota modulation and their correlation in high-fat diet-fed ApoE/ mice. Front Pharmacol 11:223 Xiao H-B, Sun Z-L, Zhang H-B, Zhang D-S (2012) Berberine inhibits dyslipidemia in C57BL/6 mice with lipopolysaccharide induced inflammation. Pharmacol Rep 64(4):889–895 Xu S, Weng J (2020) Familial hypercholesterolemia and atherosclerosis: animal models and therapeutic advances. Trends Endocrinol Metab 31(5):331–333

110

A. Fatahian et al.

Yan Y-X, Nakagawa H, Lee M-H, Rustgi AK (1997) Transforming growth factor-α enhances cyclin D1 transcription through the binding of early growth response protein to a cis-regulatory element in the cyclin D1 promoter. J Biol Chem 272(52):33181–33190 Yang X-J, Liu F, Feng N, Ding X-S, Chen Y, Zhu S-X et al (2020) Berberine attenuates cholesterol accumulation in macrophage foam cells by suppressing AP-1 activity and activation of the Nrf2/ HO-1 pathway. J Cardiovasc Pharmacol 75(1):45–53 Yin J, Xing H, Ye J (2008) Efficacy of berberine in patients with type 2 diabetes mellitus. Metabolism 57(5):712–717 Yuyun MF, Ng LL, Ng GA (2018) Endothelial dysfunction, endothelial nitric oxide bioavailability, tetrahydrobiopterin, and 5-methyltetrahydrofolate in cardiovascular disease. Where are we with therapy? Microvasc Res 119:7–12. https://doi.org/10.1016/j.mvr.2018.03.012 Zhang T, Kruys V, Huez G, Gueydan C (2002) AU-rich element-mediated translational control: complexity and multiple activities of trans-activating factors. Portland Press, London Zhang W, Xu Y-c, Guo F-j, Meng Y, Li M-l (2008) Anti-diabetic effects of cinnamaldehyde and berberine and their impacts on retinol-binding protein 4 expression in rats with type 2 diabetes mellitus. Chin Med J 121(21):2124 Zhang H, Wei J, Xue R, Wu J-D, Zhao W, Wang Z-Z et al (2010) Berberine lowers blood glucose in type 2 diabetes mellitus patients through increasing insulin receptor expression. Metabolism 59 (2):285–292 Zhang Q, Xiao X, Feng K, Wang T, Li W, Yuan T et al (2011) Berberine moderates glucose and lipid metabolism through multipathway mechanism. Evid Based Complement Alternat Med 2011:924851 Zhang M, Wang CM, Li J, Meng ZJ, Wei SN, Li J et al (2013) Berberine protects against palmitateinduced endothelial dysfunction: involvements of upregulation of AMPK and eNOS and downregulation of NOX4. Mediat Inflamm 2013:260464. https://doi.org/10.1155/2013/260464 Zhang X-J, Deng Y-X, Shi Q-Z, He M-Y, Chen B, Qiu X-M (2014) Hypolipidemic effect of the Chinese polyherbal Huanglian Jiedu decoction in type 2 diabetic rats and its possible mechanism. Phytomedicine 21(5):615–623 Zhang L-S, Zhang J-H, Feng R, Jin X-Y, Yang F-W, Ji Z-C et al (2019) Efficacy and safety of berberine alone or combined with statins for the treatment of hyperlipidemia: a systematic review and meta-analysis of randomized controlled clinical trials. Am J Chin Med 47 (04):751–767 Zhang L, Shao J, Zhou Y, Chen H, Qi H, Wang Y et al (2018) Inhibition of PDGF-BB-induced proliferation and migration in VSMCs by proanthocyanidin A2: involvement of KDR and Jak2/STAT-3/cPLA2 signaling pathways. Biomed Pharmacother 98:847–855 Zhou X, Ren F, Wei H, Liu L, Shen T, Xu S et al (2017) Combination of berberine and evodiamine inhibits intestinal cholesterol absorption in high fat diet induced hyperlipidemic rats. Lipids Health Dis 16(1):239 Zhu X, Yang J, Zhu W, Yin X, Yang B, Wei Y, Guo X (2018) Combination of berberine with resveratrol improves the lipid-lowering efficacy. Int J Mol Sci 19(12):3903 Zimetti F, Adorni MP, Ronda N, Gatti R, Bernini F, Favari E (2015) The natural compound berberine positively affects macrophage functions involved in atherogenesis. Nutr Metab Cardiovasc Dis 25(2):195–201. https://doi.org/10.1016/j.numecd.2014.08.004