Stem Cell Research & Therapy

Thematic series - open for submissions, stromal cells and progenitor cells for osteoarticular regeneration.

stem cells research papers

Organoids and Tissue/Organ Chips

stem cells research papers

Sergey Nivens / stock.adobe.com

Stem cell therapy of COVID-19 and other respiratory diseases

diabetes retinopathy

koolsabuy / stock.adobe.com

Adult stem cells in retinal diseases: where do we go from here?

Featured articles.

feature image 3

Vascular organoids: unveiling advantages, applications, challenges, and disease modelling strategies

feature image 4

Biodistribution of mesenchymal stromal cell-derived extracellular vesicles administered during acute lung injury

  • Most accessed
  • Collections

Allogeneic bone marrow mesenchymal stem cell-derived exosomes alleviate human hypoxic AKI-on-a-Chip within a tight treatment window

Authors: Sefa Burak Çam, Eda Çiftci, Nazlıhan Gürbüz, Bülent Altun and Petek Korkusuz

An integrated toolkit for human microglia functional genomics

Authors: Imdadul Haq, Jason C. Ngo, Nainika Roy, Richard L. Pan, Nadiya Nawsheen, Rebecca Chiu, Ya Zhang, Masashi Fujita, Rajesh K. Soni, Xuebing Wu, David A. Bennett, Vilas Menon, Marta Olah and Falak Sher

Oral ulcer treatment using human tonsil-derived mesenchymal stem cells encapsulated in trimethyl chitosan hydrogel: an animal model study

Authors: Hyun Seok Ryu, Celine Abueva, Andrew Padalhin, So Young Park, Seung Hyeon Yoo, Hwee Hyon Seo, Phil-Sang Chung and Seung Hoon Woo

Utilizing bioprinting to engineer spatially organized tissues from the bottom-up

Authors: Yichen Zhan, Wenbin Jiang, Zhirong Liu, Zhenxing Wang, Ke Guo and Jiaming Sun

Small-molecule α-lipoic acid targets ELK1 to balance human neutrophil and erythrocyte differentiation

Authors: Yimeng Zhang, Ya Zhou, Xiaohong Li, Xu Pan, Ju Bai, Yijin Chen, Zhenyang Lai, Qiang Chen, Feng Ma and Yong Dong

Most recent articles RSS

View all articles

Stem cells: past, present, and future

Authors: Wojciech Zakrzewski, Maciej Dobrzyński, Maria Szymonowicz and Zbigniew Rybak

Platelet-rich plasma preparation for regenerative medicine: optimization and quantification of cytokines and growth factors

Authors: Paola Romina Amable, Rosana Bizon Vieira Carias, Marcus Vinicius Telles Teixeira, Ítalo da Cruz Pacheco, Ronaldo José Farias Corrêa do Amaral, José Mauro Granjeiro and Radovan Borojevic

CAR T cells in solid tumors: challenges and opportunities

Authors: Faroogh Marofi, Roza Motavalli, Vladimir A. Safonov, Lakshmi Thangavelu, Alexei Valerievich Yumashev, Markov Alexander, Navid Shomali, Max Stanley Chartrand, Yashwant Pathak, Mostafa Jarahian, Sepideh Izadi, Ali Hassanzadeh, Naghmeh Shirafkan, Safa Tahmasebi and Farhad Motavalli Khiavi

Stem cells: a potential treatment option for kidney diseases

Authors: Dongwei Liu, Fei Cheng, Shaokang Pan and Zhangsuo Liu

Most accessed articles RSS

Topical collection Stromal cells and progenitor cells for osteoarticular regeneration Edited by Christian Jorgensen Topical collection Organoids and Tissue/Organ Chips Edited by Albert J Banes and Rajashekhar Gangaraju

Thematic series Adult stem cells in retinal diseases: where do we go from here? Edited by Rajashekhar Gangaraju

Thematic series Stem cell therapy of COVID-19 and other respiratory diseases   Edited by Hong-Long James Ji, Michael A. Matthay and Yuanlin Song

Thematic series NK cells as the next option for cancer treatment   Edited by Michael O'Dwyer

Cross-journal collection Coronavirus research highlights

Thematic series Stem cells and gene editing Edited by Stephen H. Tsang

Cross-journal collection Pluripotent Stem Cells

Thematic series Regenerative neurology Edited by Simon Koblar and Anne Hamilton-Bruce

Thematic series Extracellular vesicles and regenerative medicine Edited by Jeffrey Karp, Kelvin Ng and Armand Keating

Cross-journal collection Mesenchymal stem/stromal cells – an update Edited by Richard Schäfer and Selim Kuci

Cross-journal collection Biology and clinical applications of stem cells for autoimmune and musculoskeletal disorders Edited by Christian Jorgensen and Anthony Hollander

Thematic series Functional imaging in regenerative medicine Edited by Timothy O'Brien and Rocky Tuan

Thematic series Emerging investigators Edited by Timothy O'Brien and Rocky Tuan

Thematic series Stem cells and genitourinary regeneration Edited by John D Jackson

Thematic series Cardiovascular regeneration Edited by Ronald Li

Thematic series Physical influences on stem cells Edited by Gordana Vunjak-Novakovic

Thematic series Stem cell research in the Asia-Pacific Edited by Oscar Lee, Songtao Shi, Yufang Shi and Ying Jin

Thematic series Clinical applications of stem cells Edited by Mahendra Rao

Review series Immunology and stem cells Edited by Christian Jorgensen

Review series Epigenetics and regulation

Review series Stem cell niche

Review series Induced pluripotent stem cells

altmetric

Trending articles in Stem Cell Research & Therapy

Click here to view which Articles have been shared the most in the last quarter!

Stem Cell Research & Therapy: 10th Anniversary

To mark the 10th year anniversary, we have reviewed the milestone achievements and highlighted some of the best content selected by our Editors-in-Chief and Associate Editors. Read more here .

Editors-in-Chief

Rocky S Tuan, The Chinese University of Hong Kong, Hong Kong SAR, China  Timothy O'Brien, University of Galway, Ireland

Journal submission information

Before you submit your manuscript to Stem Cell Research & Therapy , please ensure you have read and noted the below: 

  • Authors are requested to provide institutional email addresses. At the minimum, at least the corresponding author and submitting author should provide an institutional email address.
  • Image integrity and standards: Cropped gels and blots can be included in the main text if it improves the clarity and conciseness of the presentation. In such cases, the cropping of the blot must be clearly evident and must be mentioned in the figure legend. Corresponding uncropped full-length gels and blot must be included in the supplementary files. These uncropped images should indicate where they were cropped, be labelled as in the main text and placed in a single supplementary figure. The manuscript's figure legends should state that 'Full-length blots/gels are presented in Supplementary Figure X'. Further information can be found under 'Digital image integrity' which are detailed on our  Standards of Reporting  page.
  • Authors should clearly declare in the Acknowledgement Section if there is any work that was outsourced or data not collected by authors. 
  • Authors might be asked to provide original source data (raw data / original data / individual data points) if the reviewers or editors have any question.
  • Authors should provide these information in the ethics declaration if the study involve human participants, human material, human data, or animals: (1) Title of the approved project; (2) Name of the institutional approval committee or unit; (3) Approval number; (4) Date of approval. Authors might be asked to provide ethics approval documents if the reviewers or editors have any question. 

Aims and scope

Stem Cell Research & Therapy is the major forum for translational research into stem cell therapies. An international peer-reviewed journal, it publishes high-quality open access research articles with a special emphasis on basic, translational and clinical research into stem cell therapeutics and regenerative therapies, including animal models and clinical trials. The journal also provides reviews, viewpoints, commentaries, reports and methods.

Topics covered include, but are not limited to: 

  • Adult stem cells
  • Animal models
  • Biophysics and mechanobiology
  • Cancer stem cells
  • Cell culture and manufacture
  • Clinical studies
  • Disease modeling and drug screening
  • Embryonic stem cells
  • Ethical, legal and social aspects
  • Genetics and epigenetics
  • Genomics/proteomics/metabolomics
  • Induced pluripotent stem cells
  • Regenerative medicine
  • Stem cell differentiation, proliferation and migration
  • Stem cell therapy/transplantation
  • Technologies
  • Tissue engineering and biomaterials

Announcing the launch of In Review

Stem Cell Research & Therapy , in partnership with Research Square, is now offering  In Review.  Authors choosing this free optional service will be able to:

  • Share their work with fellow researchers to read, comment on, and cite even before publication
  • Showcase their work to funders and others with a citable DOI while it is still under review
  • Track their manuscript - including seeing when reviewers are invited, and when reports are received 

Editors' quotes

stem cells research papers

About the journal

STEM CELLS , a peer-reviewed journal published monthly, provides a forum for prompt publication of original investigative papers and concise reviews. STEM CELLS is read and written by clinical and basic scientists whose expertise encompasses the rapidly expanding fields of stem and progenitor cell biology.

STEM CELLS welcomes original articles and concise reviews describing basic laboratory investigations of stem cells and the translation of their clinical aspects of characterization and manipulation from the bench to patient care. The journal covers all aspects of stem cell research including embryonic stem cells/induced pluripotent stem cells; tissue-specific stem cells; stem cell technology: epigenetics, genomics, proteomics, and metabonomics; cancer stem cells; translational and clinical research; and regenerative medicine.

STEM CELLS covers: 

Cancer Stem Cells

  • Cellular origin of cancer stem cells
  • Genetic constraints, including epigenetic modifications, that influence growth process
  • The cancer stem cell niche, and cancer stem cells, as therapeutic targets
  • Mathematical models of cancer cell evolution and mechanisms of survival

Embryonic Stem Cells/Induced Pluripotent Stem (iPS) Cells

  • Signaling pathways/factors regulating and/or initiating ESC/iPS cells pluripotency and differentiation
  • Derivation and organization of primitive, early structures with or without tissue specificity

Regenerative Medicine

  • Potential applications for stem cell-based strategies in pathological conditions
  • Tissue engineering and characterization of engineered tissues
  • Medical device and artificial organ development
  • Stem cell transplantation and technologies that will maintain, improve or restore the function of diseased organs
  • Gene therapies for inherited diseases

Stem Cell Technology: Epigenetics, Genomics, Proteomics, and Metabonomics

  • Mutagenesis and epigenetic influences
  • Surveys of gene and/or protein expression with a link to structure or function
  • Quantitative measurements of the dynamic multiparametric metabolic response of living systems to pathophysiological stimuli or genetic modification
  • Prediction of the outcome (efficacy or toxicity) of a drug or xenobiotic intervention in an individual based on a mathematical model of preintervention metabolite signatures

Tissue-Specific Stem Cells

  • Regulation of and pathways that govern stem cell self-renewal and/or differentiation
  • Cellular interactions and signaling pathways necessary for tissue specificity
  • Molecules and signaling pathways that modulated homing or mobilization of tissue-specific stem/progenitor cells
  • Functional properties of tissue-specific and progenitor cells
  • Improving engraftment for efficient and less costly cellular therapy for adult stem cells
  • Understanding and avoiding teratoma formation

Translational and Clinical Research

  • Studies describing new model systems with potential for drug development
  • Translation into pharmaceutical, clinical, or practical findings

Abstracting and indexing information

  • Academic OneFile (GALE Cengage)
  • Biological Abstracts (Thomson Reuters)
  • Biological Sciences (ProQuest)
  • BIOSIS Previews (Thomson Reuters)
  • Biotechnology & Bioengineering Abstracts (ProQuest)
  • Biotechnology Research Abstracts (ProQuest)
  • Calcium & Calcified Tissue Abstracts (ProQuest)
  • CAS: Chemical Abstracts Service (ACS)
  • Chemoreception Abstracts (ProQuest)
  • CSA Engineering Research Database (ProQuest)
  • CSA Nucleic Acids Abstracts (ProQuest)
  • CSA Technology Research Database (ProQuest)
  • Current Contents: Life Sciences (Thomson Reuters)
  • Embase (Elsevier)
  • Genetics Abstracts (ProQuest)
  • Health Reference Center Academic (GALE Cengage)
  • HEED: Health Economic Evaluations Database (Wiley-Blackwell)
  • InfoTrac (GALE Cengage)
  • Materials Science & Engineering Database (ProQuest)
  • MEDLINE/PubMed (NLM)
  • Neurosciences Abstracts (ProQuest)
  • ProQuest Biological Science Collection (ProQuest)
  • ProQuest Central (ProQuest)
  • ProQuest Health & Medical Complete (ProQuest)
  • ProQuest Natural Science Collection (ProQuest)
  • ProQuest SciTech Collection (ProQuest)
  • ProQuest Technology Collection (ProQuest)
  • Science Citation Index (Thomson Reuters)
  • SCOPUS (Elsevier)
  • VINITI (All-Russian Institute of Science & Technological Information)
  • Web of Science (Thomson Reuters)
  • Recommend to Your Librarian
  • Advertising and Corporate Services

Affiliations

Stem Cells

  • Online ISSN 1549-4918
  • Copyright © 2024 Oxford University Press
  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Institutional account management
  • Rights and permissions
  • Get help with access
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • My Account Login
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Review Article
  • Open access
  • Published: 06 August 2022

Stem cell-based therapy for human diseases

  • Duc M. Hoang   ORCID: orcid.org/0000-0001-5444-561X 1 ,
  • Phuong T. Pham 2 ,
  • Trung Q. Bach 1 ,
  • Anh T. L. Ngo 2 ,
  • Quyen T. Nguyen 1 ,
  • Trang T. K. Phan 1 ,
  • Giang H. Nguyen 1 ,
  • Phuong T. T. Le 1 ,
  • Van T. Hoang 1 ,
  • Nicholas R. Forsyth 3 ,
  • Michael Heke 4 &
  • Liem Thanh Nguyen 1  

Signal Transduction and Targeted Therapy volume  7 , Article number:  272 ( 2022 ) Cite this article

65k Accesses

215 Citations

66 Altmetric

Metrics details

  • Mesenchymal stem cells
  • Stem-cell research

Recent advancements in stem cell technology open a new door for patients suffering from diseases and disorders that have yet to be treated. Stem cell-based therapy, including human pluripotent stem cells (hPSCs) and multipotent mesenchymal stem cells (MSCs), has recently emerged as a key player in regenerative medicine. hPSCs are defined as self-renewable cell types conferring the ability to differentiate into various cellular phenotypes of the human body, including three germ layers. MSCs are multipotent progenitor cells possessing self-renewal ability (limited in vitro) and differentiation potential into mesenchymal lineages, according to the International Society for Cell and Gene Therapy (ISCT). This review provides an update on recent clinical applications using either hPSCs or MSCs derived from bone marrow (BM), adipose tissue (AT), or the umbilical cord (UC) for the treatment of human diseases, including neurological disorders, pulmonary dysfunctions, metabolic/endocrine-related diseases, reproductive disorders, skin burns, and cardiovascular conditions. Moreover, we discuss our own clinical trial experiences on targeted therapies using MSCs in a clinical setting, and we propose and discuss the MSC tissue origin concept and how MSC origin may contribute to the role of MSCs in downstream applications, with the ultimate objective of facilitating translational research in regenerative medicine into clinical applications. The mechanisms discussed here support the proposed hypothesis that BM-MSCs are potentially good candidates for brain and spinal cord injury treatment, AT-MSCs are potentially good candidates for reproductive disorder treatment and skin regeneration, and UC-MSCs are potentially good candidates for pulmonary disease and acute respiratory distress syndrome treatment.

Similar content being viewed by others

stem cells research papers

Mesenchymal stem cell perspective: cell biology to clinical progress

Mark F. Pittenger, Dennis E. Discher, … Arnold I. Caplan

stem cells research papers

Mesenchymal stem cell-derived extracellular vesicles for immunomodulation and regeneration: a next generation therapeutic tool?

Meng Kou, Li Huang, … Qizhou Lian

stem cells research papers

Next-generation stem cells — ushering in a new era of cell-based therapies

Erin A. Kimbrel & Robert Lanza

Introduction

The successful approval of cancer immunotherapies in the US and mesenchymal stem cell (MSC)-based therapies in Europe have turned the wheel of regenerative medicine to become prominent treatment modalities. 1 , 2 , 3 Cell-based therapy, especially stem cells, provides new hope for patients suffering from incurable diseases where treatment approaches focus on management of the disease not treat it. Stem cell-based therapy is an important branch of regenerative medicine with the ultimate goal of enhancing the body repair machinery via stimulation, modulation, and regulation of the endogenous stem cell population and/or replenishing the cell pool toward tissue homeostasis and regeneration. 4 Since the stem cell definition was introduced with their unique properties of self-renewal and differentiation, they have been subjected to numerous basic research and clinical studies and are defined as potential therapeutic agents. As the main agenda of regenerative medicine is related to tissue regeneration and cellular replacement and to achieve these targets, different types of stem cells have been used, including human pluripotent stem cells (hPSCs), multipotent stem cells and progenitor cells. 5 However, the emergence of private and unproven clinics that claim the effectiveness of stem cell therapy as “magic cells” has raised highly publicized concerns about the safety of stem cell therapy. The most notable case involved the injection of a cell population derived from fractionated lipoaspirate into the eyes of three patients diagnosed with macular degeneration, resulting in the loss of vision for these patients. 6 Thus, as regenerative medicine continues to progress and evolve and to clear the myth of the “magic” cells, this review provides a brief overview of stem cell-based therapy for the treatment of human diseases.

Stem cell therapy is a novel therapeutic approach that utilizes the unique properties of stem cells, including self-renewal and differentiation, to regenerate damaged cells and tissues in the human body or replace these cells with new, healthy and fully functional cells by delivering exogenous cells into a patient. 7 Stem cells for cell-based therapy can be of (1) autologous, also known as self-to-self therapy, an approach using the patient’s own cells, and (2) allogeneic sources, which use cells from a healthy donor for the treatment. 8 The term “stem cell” were first used by the eminent German biologist Ernst Haeckel to describe the properties of fertilized egg to give rise to all cells of the organism in 1868. 9 The history of stem cell therapy started in 1888, when the definition of stem cell was first coined by two German zoologists Theodor Heinrich Boveri and Valentin Haecker, 9 who set out to identify the distinct cell population in the embryo capable of differentiating to more specialized cells (Fig. 1a ). In 1902, studies carried out by the histologist Franz Ernst Christian Neumann, who was working on bone marrow research, and Alexander Alexandrowitsch Maximov demonstrated the presence of common progenitor cells that give rise to mature blood cells, a process also known as haematopoiesis. 10 From this study, Maximov proposed the concept of polyblasts, which later were named stem cells based on their proliferation and differentiation by Ernst Haeckel. 11 Maximov described a hematopoietic population presented in the bone marrow. In 1939, the first case report described the transplantation of human bone marrow for a patient diagnosed with aplastic anemia. Twenty years later, in 1958, the first stem cell transplantation was performed by the French oncologist George Mathe to treat six nuclear researchers who were accidentally exposed to radioactive substances using bone marrow transplantation. 12 Another study by George Mathe in 1963 shed light on the scientific community, as he successfully conducted bone marrow transplantation in a patient with leukemia. The first allogeneic hematopoietic stem cell transplantation (HSCT) was pioneered by Dr. E. Donnall Thomas in 1957. 13 In this initial study, all six patients died, and only two patients showed evidence of transient engraftment due to the unknown quantities and potential hazards of bone marrow transplantation at that time. In 1969, Dr. E. Donnall Thomas conducted the first bone marrow transplantation in the US, although the success of the allogeneic treatment remained exclusive. In 1972, the year marked the discovery of cyclosporine (the immune suppressive drug), 14 the first successes of allogeneic transplantation for aplastic anemia and acute myeloid leukemia were reported in a 16-year-old girl. 15 From the 1960s to the 1970s, series of works conducted by Friendenstein and coworkers on bone marrow aspirates demonstrated the relationship between osteogenic differentiation and a minor subpopulation of cells derived from bone marrow. 16 These cells were later proven to be distinguishable from the hematopoietic population and to be able to proliferate rapidly as adherent cells in tissue culture vessels. Another important breakthrough from Friendenstein’s team was the discovery that these cells could form the colony-forming unit when bone marrow was seeded as suspension culture following by differentiation into osteoblasts, adipocytes, and chondrocytes, suggesting that these cells confer the ability to proliferate and differentiate into different cell types. 17 In 1991, combined with the discovery of human embryonic stem cells (hESCs), which will be discussed in the next section, the term “mesenchymal stem cells”, previously known as stromal stem cells or “osteogenic” stem cells, was first coined in Caplan and widely used to date. 18 Starting with bone marrow transplantation 60 years ago, the journey of stem cell therapy has developed throughout the years to become a novel therapeutic agent of regenerative medicine to treat numerous incurable diseases, which will be reviewed and discussed in this review, including neurological disorders, pulmonary dysfunctions, metabolic/endocrine-related diseases, reproductive disorders, skin burns, and cardiovascular conditions).

figure 1

Stem cell-based therapy: the history and cell source. a The timeline of major discoveries and advances in basic research and clinical applications of stem cell-based therapy. The term “stem cells” was first described in 1888, setting the first milestone in regenerative medicine. The hematopoietic progenitor cells were first discovered in 1902. In 1939, the first bone marrow transplantation was conducted in the treatment of aplasmic anemia. Since then, the translation of basic research to preclinical studies to clinical trials has driven the development of stem cell-based therapy by many discoveries and milestones. The isolations of “mesenchymal stem cells” in 1991 following by the discovery of human pluripotent stem cells have recently contributed to the progress of stem cell-based therapy in the treatment of human diseases. b Schematic of the different cell sources that can be used in stem cell-based therapy. (1) Human pluripotent stem cells, including embryonic stem cells (derived from inner cell mass of blastocyst) and induced pluripotent stem cells confer the ability to proliferate indefinitely in vitro and differentiate into numerous cell types of the human body, including three germ layers. (2) Mesenchymal stem cells are multipotent stem cells derived from mesoderm possessing self-renewal ability (limited in vitro) and differentiation potential into mesenchymal lineages. The differentiated/somatic cells can be reprogrammed back to the pluripotent stage using OSKM factors to generate induced pluripotent stem cells. It is important to note that stem cells show a relatively higher risk of tumor formation and lower risk of immune rejection (in the case of mesenchymal stem cells) when compared to that of somatic cells. The figure was created with BioRender.com

In this review, we described the different types of stem cell-based therapies (Fig. 1b ), including hPSCs and MSCs, and provided an overview of their definition, history, and outstanding clinical applications. In addition, we further created the first literature portfolio for the “targeted therapy” of MSCs based on their origin, delineating their different tissue origins and downstream applications with an in-depth discussion of their mechanism of action. Finally, we provide our perspective on why the tissue origin of MSCs could contribute greatly to their downstream applications as a proposed hypothesis that needs to be proven or disproven in the future to further enhance the safety and effectiveness of stem cell-based therapy.

Stem cell-based therapy: an overview of current clinical applications

Cardiovascular diseases.

The clinical applications of stem cell-based therapies for heart diseases have been recently discussed comprehensively in the reviews 19 , 20 and therefore will be elaborated in this study as the focus discussions related to hPSCs and MSCs in the following sections. In general, the safety profiles of stem cell-based therapies are supported by a large body of preclinical and clinical studies, especially adult stem cell therapy (such as MSC-based products). However, clinical trials have not yet yielded data supporting the efficacy of the treatment, as numerous studies have shown paradoxical results and no statistically significant differences in infarct size, cardiac function, or clinical outcomes, even in phase III trials. 21 The results of a meta-analysis showed that stem cells derived from different sources did not exhibit any therapeutic effects on the improvement of myocardial contractility, cardiovascular remodeling, or clinical outcomes. 22 The disappointing results obtained from the clinical trials thus far could be explained by the fact that the administered cells may exert their therapeutic effects via an immune modulation rather than regenerative function. Thus, well-designed, randomized and placebo-controlled phase III trials with appropriate cell-preparation methods, patient selection, follow-up schedules and suitable clinical measurements need to be conducted to determine the efficacy of the treatments. In addition, concerns related to optimum cell source and dose, delivery route and timing of administration, cell distribution post administration and the mechanism of action also need to be addressed. In the following section of this review, we present clinical trials related to MSC-based therapy in cardiovascular disease with the aim of discussing the contradictory results of these trials and analyzing the potential challenges underlying the current approaches.

Digestive system diseases

Gastrointestinal diseases are among the most diagnosed conditions in the developed world, altering the life of one-third of individuals in Western countries. The gastrointestinal tract is protected from adverse substances in the gut environment by a single layer of epithelial cells that are known to have great regenerative ability in response to injuries and normal cell turnover. 23 These epithelial cells have a rapid turnover rate of every 2–7 days under normal conditions and even more rapidly following tissue damage and inflammation. This rapid proliferation ability is possible owing to the presence of a specific stem cell population that is strictly compartmentalized in the intestinal crypts. 24 The gastrointestinal tract is highly vulnerable to damage, tissue inflammation and diseases once the degradation of the mucosal lining layer occurs. The exposure of intestinal stem cells to the surrounding environment of the gut might result in the direct destruction of the stem cell layer or disruption of intestinal functions and lead to overt clinical symptoms. 25 In addition, the accumulation of stem cell defects as well as the presence of chronic inflammation and stress also contributes to the reduction of intestinal stem cell quality.

In terms of digestive disorders, Crohn’s disease (CD) and ulcerative colitis are the two major forms of inflammatory bowel disease (IBD) and represent a significant burden on the healthcare system. The former is a chronic, uncontrolled inflammatory condition of the intestinal mucosa characterized by segmental transmural mucosal inflammation and granulomatous changes. 26 The latter is a chronic inflammatory bowel disease affecting the colon and rectum, characterized by mucosal inflammation initiating in the rectum and extending proximal to the colon in a continuous fashion. 27 Cellular therapy in the treatment of CD can be divided into haematopoietic stem cell-based therapy and MSC-based therapy. The indication and recommendation of using HSCs for the treatment of IBD were proposed in 1995 by an international committee with four important criteria: (1) refractory to immunosuppressive treatment; (2) persistence of the disease conditions indicated via endoscopy, colonoscopy or magnetic resonance enterography; (3) patients who underwent an imminent surgical procedure with a high risk of short bowel syndromes or refractory colonic disease; and (4) patients who refused to treat persistent perianal lesions using coloproctectomy with a definitive stroma implant. 28 In the standard operation procedure, patents’ HSCs were recruited using cyclophosphamide, which is associated with granulocyte colony-stimulating factor (G-CSF), at two different administered concentrations (4 g/m 2 and 2 g/m 2 ). Recently, it was reported that high doses of cyclophosphamide do not improve the number of recruited HSCs but increase the risk of cardiac and bladder toxicity. An interest in using HSCTs in CD originated from case reports that autologous HSCTs can induce sustained disease remission in some 29 , 30 but not all patients 31 , 32 , 33 with CD. The first phase I trial was conducted in Chicago and recruited 12 patients with active moderate to severe CD refractory to conventional therapies. Eleven of 12 patients demonstrated sustained remission after a median follow-up of 18.5 months, and one patient developed recurrence of active CD. 31 The ASTIC trial (the Autologous Stem Cell Transplantation International Crohn Disease) was the first randomized clinical trial with the largest cohort of patients undergoing HSCT for refractory CD in 2015. 34 The early report of the trial showed no statistically significant improvement in clinical outcomes of mobilization and autologous HSCT compared with mobilization followed by conventional therapy. In addition, the procedure was associated with significant toxicity, leading to the suggestion that HSCT for patients with refractory CD should not be widespread. Interestingly, by using conventional assessments for clinical trials for CD, a group reassessed the outcomes of patients enrolled in the ASTIC trial showing clinical and endoscopic benefits, although a high number of adverse events were also detected. 35 A recent systematic review evaluated 18 human studies including 360 patients diagnosed with CD and showed that eleven studies confirmed the improvement of Crohn’s disease activity index between HSCT groups compared to the control group. 36 Towards the cell sources, HSCs are the better sources as they afforded more stable outcomes when compared to that of MSC-based therapy. 37 Moreover, autologous stem cells were better than their allogeneic counterparts. 36 The safety of stem cell-based therapy in the treatment of CD has attracted our attention, as the risk of infection in patients with CD was relatively higher than that in those undergoing administration to treat cancer or other diseases. During the stem cell mobilization process, patient immunity is significantly compromised, leading to a high risk of infection, and requires carefully nursed and suitable antibiotic treatment to reduce the development of adverse events. Taken together, stem cell-based therapy for digestive disease reduced inflammation and improved the patient’s quality of life as well as bowel functions, although the high risk of adverse events needs to be carefully monitored to further improve patient safety and treatment outcomes.

Liver diseases

The liver is the largest vital organ in the human body and performs essential biological functions, including detoxification of the organism, metabolism, supporting digestion, vitamin storage, and other functions. 38 The disruption of liver homeostasis and function might lead to the development of pathological conditions such as liver failure, cirrhosis, cancer, alcoholic liver disease, nonalcoholic fatty liver disease (NAFLD), and autoimmune liver disease (ALD). Orthotropic liver transplantation is the only effective treatment for severe liver diseases, but the number of available and suitable donor organs is very limited. Currently, stem cell-based therapies in the treatment of liver disease are associated with HSCs, MSCs, hPSCs, and liver progenitor cells.

Liver failure is a critical condition characterized by severe liver dysfunctions or decompensation caused by numerous factors with a relatively high mortality rate. Stem cell-based therapy is a novel alternative approach in the treatment of liver failure, as it is believed to participate in the enhancement of liver regeneration and recovery. The results of a meta-analysis including four randomized controlled trials and six nonrandomized controlled trials in the treatment of acute-on-chronic liver failure (ACLF) demonstrated that clinical outcomes of stem cell therapy were achieved in the short term, requiring multiple doses of stem cells to prolong the therapeutic effects. 39 , 40 Interestingly, although MSC-based therapies improved liver functions, including the model of end-stage liver disease score, albumin level, total bilirubin, and coagulation, beneficial effects on survival rate and aminotransferase level were not observed. 41 A randomized controlled trial illustrated the improvement of liver functions and reduction of severe infections in patients with hepatitis B virus-related ACLF receiving allogeneic bone marrow-derived MSCs (BM-MSCs) via peripheral infusion. 42 HSCs from peripheral blood after the G-CSF mobilization process were used in a phase I clinical trial and exhibited an improvement in serum bilirubin and albumin in patients with chronic liver failure without any specific adverse events related to the administration. 43 Taken together, an overview of stem cell-based therapy in the treatment of liver failure indicates the potential therapeutic effects on liver functions with a strong safety profile, although larger randomized controlled trials are still needed to assure the conclusions.

Liver cirrhosis is one of the major causes of morbidity and mortality worldwide and is characterized by diffuse nodular regeneration with dense fibrotic septa and subsequent parenchymal extinction leading to the collapse of liver vascular structure. 44 In fact, liver cirrhosis is considered the end-stage of liver disease that eventually leads to death unless liver transplantation is performed. Stem cell-based therapy, especially MSCs, currently emerges as a potential treatment with encouraging results for treating liver cirrhosis. In a clinical trial using umbilical cord-derived MSCs (UC-MSCs), 45 chronic hepatitis B patients with decompensated liver cirrhosis were divided into two groups: the MSC group ( n  = 30) and the control group ( n  = 15). 45 The results showed a significant reduction in ascites volume in the MSC group compared with the control. Liver function was also significantly improved in the MSC groups, as indicated by the increase in serum albumin concentration, reduction in total serum bilirubin levels, and decrease in the sodium model for end-stage liver disease score. 45 Similar results were also reported from a phase II trial using BM-MSCs in 25 patients with HCV-induced liver cirrhosis. 46 Consistent with these studies, three other clinical trials targeting liver cirrhosis caused by hepatitis B and alcoholic cirrhosis were conducted and confirmed that MSC administration enhanced and recovered liver functions. 47 , 48 , 49 With the large cohort study as the clinical trial conducted by Fang, the safety and potential therapeutic effects of MSC-based therapies could be further strengthened and confirmed the feasibility of the treatment in virus-related liver cirrhosis. 49 In terms of delivery route, a randomized controlled phase 2 trial suggested that systemic delivery of BM-MSCs does not show therapeutic effects on patients with liver cirrhosis. 50 MSCs are not the only cell source for liver cirrhosis. Recently, an open-label clinical trial conducted in 19 children with liver cirrhosis due to biliary atresia after the Kasai operation illustrated the safety and feasibility of the approach by showing the improvement of liver function after bone marrow mononuclear cell (BMNC) administration assessed by biochemical tests and pediatric end-stage liver disease (PELD) scores. 51 Another study using BMNCs in 32 decompensated liver cirrhosis patients illustrated the safety and effectiveness of BMNC administration in comparison with the control group. 52 Recently, a long-term analysis of patients receiving peripheral blood-derived stem cells indicated a significant improvement in the long-term survival rate when compared to the control group, and the risk of hepatocellular carcinoma formation did not increase. 53 CD133 + HSC infusion was performed in a multicentre, open, randomized controlled phase 2 trial in patients with liver cirrhosis; the results did not support the improvement of liver conditions, and cirrhosis persisted. 54 Notably, these results are in line with a previous randomized controlled study, which also reported that G-CSF and bone marrow-derived stem cells delivered via the hepatic artery did not introduce therapeutic potential as expected. 55 Thus, stem cell-based therapy for liver cirrhosis is still in its immature stage and requires larger trials with well-designed experiments to confirm the efficacy of the treatment.

Nonalcoholic fatty liver disease (NAFLD) is the most common medical condition caused by genetic and lifestyle factors and results in a severe liver condition and increased cardiovascular risk. 56 NAFLD is the hidden enemy, as most patients are asymptomatic for a long time, and their routine life is unaffected. Thus, the detection, identification, and management of NAFLD conditions are challenging tasks, as patients diagnosed with NAFLD often develop nonalcoholic steatohepatitis, cirrhosis, and hepatocellular carcinoma. 57 Although preclinical studies have shown that stem cell administration could enhance liver function in NAFLD models, a limited number of clinical trials were performed in human subjects. Recently, a multi-institutional clinical trial using freshly isolated autologous adipose tissue-derived regenerative cells was performed in Japan to treat seven NAFLD patients. 58 The results illustrated the improvement in the serum albumin level of six patients and prothrombin activity of five patients, and no treatment-related adverse events or severe adverse events were observed. This study illustrates the therapeutic potential of stem cell-based therapy in the treatment of NAFLD.

Autoimmune liver disease (ALD) is a severe liver condition affecting children and adults worldwide, with a female predominance. 59 The condition occurs in genetically predisposed patients when a stimulator, such as virus infection, leads to a T-cell-mediated autoimmune response directed against liver autoantigens. As a result, patients with ALD might develop liver cirrhosis, hepatocellular carcinoma, and, in severe cases, death. To date, HSCT and bone marrow transplantation are the two common stem cell-based therapies exhibiting therapeutic potential for ALD in clinical trials. An interesting report illustrated that haploidentical HSCTs could cure ALD in patients with sickle cells. 60 This report is particularly important, as it illustrates the potential therapeutic approach of using haploidentical HSCTs to treat patients with both sickle cells and ALD. Another case report described a 19-year-old man with a 4-year history of ALD who developed acute lymphoblastic leukemia and required allogeneic bone marrow transplantation from this wholesome brother. 61 The clinical data showed that immunosuppressive therapy for transplantation generated ALD remission in the patient. 62 However, the data also provided valid information related to the sustained remission and the normalization of ASGPR-specific suppressor-inducer T-cell activity following bone marrow transplantation, suggesting that these suppressor functions originated from donor T cells. 61 Thus, it was suggested that if standard immunosuppressive treatment fails, alternative cellular immunotherapy would be a viable option for patients with ALD. Primary biliary cholangitis (PBC), usually known as primary biliary cirrhosis, is a type of ALD characterized by a slow, progressive destruction of small bile ducts of the liver leading to the formation of cirrhosis and accumulation of bile and other toxins in the liver. A pilot, single-arm trial from China recruited seven patents with PBC who had a suboptimal response to ursodeoxycholic acid (UDCA) treatment. 63 These patients received UDCA treatment in combination with three rounds of allogeneic UC-MSCs at 4-week intervals with a dose of 0.5 × 10 6 cells/kg of patient body weight via the peripheral vein. No treatment-related adverse events or severe adverse events were observed throughout the course of the study. The clinical data indicated significant improvement in liver function, including reduction of serum ALP and gamma-glutamyltransferase (GGT) at 48 weeks post administration. The common symptoms of PBC, including fatigue, pruritus, and hypogastric ascites volume, were also reduced, supporting the feasibility of MSC-based therapy in the treatment of PBC, although major limitations of the study were nonrandomized, no control group and small sample size. Another study was conducted in China with ten PBC patients who underwent incompetent UDCA treatment for more than 1 year. These patients received a range of 3–5 allogeneic BM-MSCs/kg body weight by intravenous infusion. 64 Although these early studies have several limitations, such as small sample size, nonrandomization, and no control group, their preliminary data related to safety and efficacy herald the prospects and support the feasibility of stem cell-based therapy in the treatment of ALD.

In summary, the current number of trials for liver disease using stem cell-based therapy has provided fundamental data supporting the safety and potential therapeutic effects in various liver diseases. Unfortunately, due to the small number of trials, several obstacles need to be overcome to prove the effectiveness of the treatments, including (1) stem cell source and dose, (2) administration route, (3) time of intervention, and (4) clinical assessments during the follow-up period. Only by addressing these challenges we will be able to prove, facilitate and promote stem cell-based therapy as a mainstream treatment for liver diseases.

Arthritis is a general term describing cartilage conditions that cause pain and inflammation of the joints. Osteoarthritis (OA) is the most common form of arthritis caused by persistent degeneration and poor recovery of articular cartilage. 65 OA affects one or several diarthrodial joints, such as small joints at the hand and large joints at the knee and hips, leading to severe pain and subsequent reduction in the mobility of patients. There are two types of OA: (1) primary OA or idiopathic OA and secondary OA caused by causative factors such as trauma, surgery, and abnormal joint development at birth. 66 As conventional treatments for OA are not consistent in their effectiveness and might cause unbearable pain as well as long-term rehabilitation (in the case of joint replacement), there is a need for a more reliable, less painful, and curative therapy targeting the root of OA. 67 Thus, stem cell therapy has recently emerged as an alternative approach for OA and has drawn great attention in the regenerative field.

The administration of HSCs has been proven to reduce bone lesions, enhance bone regeneration and stimulate the vascularization process in degenerative cartilage. Attempts were made to evaluate the efficacy of peripheral blood stem cells in ten OA patients by three intraarticular injections. Post-administration analysis indicated a reduction in the WOMAC index with a significant reduction in all parameters. All patients completed 6-min walk tests with an increase of more than 54 meters. MRI analysis indicated an improvement in cartilage thickness, suggesting that cartilage degeneration was reduced post administration. To further enhance the therapeutic potential of HSCT, CD34 + stem cells were proposed to be used in combination with the rehabilitation algorithm, which included three stages: preoperative, hospitalization and outpatient periods. 68 Currently, a large wave of studies has been directed to MSC-based therapy for the treatment of OA due to their immunoregulatory functions and anti-inflammatory characteristics. MSCs have been used as the main cell source in several multiple and small-scale trials, proving their safety profile and potential effectiveness in alleviating pain, reducing cartilage degeneration, and enhancing the regeneration of cartilage structure and morphology in some cases. However, the best source of MSCs, whether from bone marrow, adipose tissue, or umbilical cord, for the management of OA is still a great question to be answered. A systematic review investigating over sixty-one of 3172 articles with approximately 2390 OA patients supported the positive effects of MSC-based therapy on OA patients, although the study also pointed out the fact that these therapeutic potentials were based on limited high-quality evidence and long-term follow-up. 69 Moreover, the study found no obvious evidence supporting the most effective source of MSCs for treating OA. Another systematic review covering 36 clinical trials, of which 14 studies were randomized trials, provides an interesting view in terms of the efficacy of autologous MSC-based therapy in the treatment of OA. 70 In terms of BM-MSCs, 14 clinical trials reported the clinical outcomes at the 1-year follow-up, in which 57% of trials reported clinical outcomes that were significantly better in comparison with the control group. However, strength analysis of the data set showed that outcomes from six trials were low, whereas the outcomes of the remaining eight trials were extremely low. Moreover, the positive evidence obtained from MRI analysis was low to very low strength of evidence after 1-year post administration. 70 Similar results were also found in the outcome analysis of autologous adipose tissue-derived MSCs (AT-MSCs). Thus, the review indicated low quality of evidence for the therapeutic potential of MSC therapy on clinical outcomes and MRI analysis. The low quality of clinical outcomes could be explained by the differences in interventions (including cell sources, cell doses, and administration routes), combination treatments (with hyaluronic acid, 71 peripheral blood plasma, 72 etc.), control treatments and clinical outcome measurements between randomized clinical trials. 73 In addition, the data of the systematic analysis could not prove the better source of MSCs for OA treatment. Taken together, although stem cell-based therapy has been shown to be safe and feasible in the management of OA, the authors support the notion that stem cell-based therapy could be considered an alternative treatment for OA when first-line treatments, such as education, exercise, and body weight management, have failed.

Cancer treatment

Stem cell therapy in the treatment of cancer is a sensitive term and needs to be used and discussed with caution. Clinicians and researchers should protect patients with cancer from expensive and potentially dangerous or ineffective stem cell-based therapy and patients without a cancer diagnosis from the risk of malignancy development. In general, unproven stem cell clinics employed three cell-based therapies for cancer management, including autologous HSCTs, stromal vascular fraction (SVF), and multipotent stem cells, such as MSCs. Allogeneic HSCTs confer the ability to generate donor lymphocytes that contribute to the suppression and regression of hematological malignancies and select solid tumors, a specific condition known as “graft-versus-tumor effects”. 74 However, stem cell clinics provide allogeneic cell-based therapy for the treatment of solid malignancies despite limited scientific evidence supporting the safety and efficacy of the treatment. High-quality evidence from the Cochrane library shows that marrow transplantation via autologous HSCTs in combination with high-dose chemotherapy does not improve the overall survival of women with metastatic breast cancer. In addition, a study including more than 41,000 breast cancer patients demonstrated no significant difference in survival benefits between patients who received HSCTs following high-dose chemotherapy and patients who underwent conventional treatment. 75 Thus, the use of autologous T-cell transplants as monotherapy and advertising stem cell-based therapies as if they are medically approved or preferred treatment of solid tumors is considered untrue statements and needs to be alerted to cancer patients. 76

Over the past decades, many preclinical studies have demonstrated the potential of MSC-based therapy in cancer treatment due to their unique properties. They confer the ability to migrate toward damaged sites via inherent tropism controlled by growth factors, chemokines, and cytokines. MSCs express specific C–X–C chemokine receptor type 4 (CXCR4) and other chemokine receptors (including CCR1, CCR2, CCR4, CCR7, etc.) that are essential to respond to the surrounding signals. 77 In addition, specific adherent proteins, including CD49d, CD44, CD54, CD102, and CD106, are also expressed on the MSC surface, allowing them to attach, rotate, migrate, and penetrate the blood vessel lumen to infiltrate the damaged tissue. 78 Similar to damaged tissues, tumors secrete a wide range of chemoattractant that also attract MSC migration via the CXCL12/CXCR4 axis. Previous studies also found that MSC migration toward the cancer site is tightly controlled by diffusible cytokines such as interleukin 8 (IL-8) and growth factors including transforming growth factor-beta 1 (TGF-β1), 79 platelet-derived growth factor (PDGF), 80 fibroblast growth factor 2 (FGF-2), 81 vascular endothelial growth factor (VEGF), 81 and extracellular matrix molecules such as matrix metalloproteinase-2 (MMP-2). 82 Once MSCs migrate successfully to cancerous tissue, accumulating evidence demonstrates the interaction between MSCs and cancer cells to exhibit their protumour and antitumour effects, which are the major concerns of MSC-based therapy. MSCs are well-known for their regenerative effects that regulate tissue repair and recovery. This unique ability is also attributed to the protumour functions of these cells. A previous study reported that breast cancer cells induce MSC secretion of chemokine (C–C motif) ligand 5 (CCL-5), which regulates the tumor invasion process. 83 , 84 Other studies also found that MSCs secrete a wide range of growth factors (VEGF, basic FGF, HGF, PDGF, etc.) that inhibits apoptosis of cancer cells. 85 Moreover, MSCs also respond to signals released from cancer cells, such as TGF-β, 86 to transform into cancer-associated fibroblasts, a specific cell type residing within the tumor microenvironment capable of promoting tumorigenesis. 87 Although MSCs have been proven to be involved in protumour activities, they also have potent tumor suppression abilities that have been used to develop cancer treatments. It has been suggested that MSCs exhibit their tumor inhibitory effects by inhibiting the Wnt and AKT signaling pathways, 88 reducing the angiogenesis process, 89 stimulating inflammatory cell infiltration, 90 and inducing tumor cell cycle arrest and apoptosis. 91 To date, the exact functions of MSCs in both protumour and antitumor activities are still a controversial issue across the stem cell field. Other approaches exploit gene editing and tissue engineering to convert MSCs into “a Trojan horse” that could exhibit antitumor functions. In addition, MSCs can also be modified to express specific anticancer miRNAs exhibiting tumor-suppressive behaviors. 92 However, genetically modified MSCs are still underdeveloped and require intensive investigation in the clinical setting.

To date, ~25 clinical trials have been registered on ClinicalTrials.gov aimed at using MSCs as a therapeutic treatment for cancer. 93 These trials are mostly phase 1 and 2 studies focusing on evaluating the safety and efficacy of the treatment. Studies exploiting MSC-based therapy have combined MSCs with an oncolytic virus approach. Oncolytic viruses are specific types of viruses that can be genetically engineered or naturally present, conferring the ability to selectively infect cancer cells and kill them without damaging the surrounding healthy cells. 94 A completed phase I/II study using BM-MSCs infected with the oncolytic adenovirus ICOVIR5 in the treatment of metastatic and refractory solid tumors in children and adult patients demonstrated the safety of the treatment and provided preliminary data supporting their therapeutic potential. 95 The same group also reported a complete disappearance of all signs of cancer in response to MSC-based therapy in one pediatric case three years post administration. 96 A reported study in 2019 claimed that adipose-derived MSCs infected with vaccinia virus have the potential to eradicate resistant tumor cells via the combination of potent virus amplification and senitization of the tumor cells to virus infection. 97 However, in a recently published review, a valid question was posed regarding the 2019 study that “do these reported data merit inclusion in the publication record when they were collected by such groups using a dubious therapeutic that was eventually confiscated by US Marshals?” 76

Taken together, cancer research and therapy have entered an innovative and fascinating era with advancements in traditional therapies such as chemotherapy, radiotherapy, and surgery on one hand and stem cell-based therapy on the other hand. Although stem cell-based therapy has been considered a novel and attractive therapeutic approach for cancer treatment, it has been hampered by contradictory results describing the protumour and antitumour effects in preclinical studies. Despite this contradictory reality, the use of stem cell-based therapy, especially MSCs, offers new hope to cancer patients by providing a new and more effective tool in personalized medicine. The authors support the use of MSC-based therapy as a Trojan horse to deliver specific anticancer functions toward cancer cells to suppress their proliferation, eradicate cancer cells, or limit the vascularization process of cancerous tissue to improve the clinical safety and efficacy of the treatment.

Human pluripotent stem cell-based therapy: a growing giant

The discovery of hPSCs, including human embryonic stem cells (hESCs) and human induced pluripotent stem cells (hiPSCs), has revolutionized stem cell research and cell-based therapy. 98 hESCs were first isolated from blastocyst-stage embryos in 1998, 99 followed by breakthrough reprogramming research that converted somatic cells into hiPSCs using just four genetic factors. 100 , 101 Methods have been developed to maintain these cells long-term in vitro and initiate their differentiation into a wide variety of cell types, opening a new era in regenerative medicine, particularly cell therapy to replace lost or damaged tissues.

History of hPSCs

hPSCs are defined as self-renewable cell types that confer the ability to differentiate into various cellular phenotypes of the human body, including three germ layers. 102 Historically, the first pluripotent cell lines to be generated were embryonic carcinoma (EC) cell lines established from human germ cell tumors 103 and murine undifferentiated compartments. 104 Although EC cells are a powerful tool in vitro, these cells are not suitable for clinical applications due to their cancer-derived origin and aneuploidy genotype. 105 The first murine ESCs were established in 1981 based on the culture techniques obtained from EC research. 106 Murine ESCs are derived from the inner cell mass (ICM) of the pre-implantation blastocyst, a unique biological structure that contains outer trophoblast layers that give rise to the placenta and ICM. 107 In vivo ESCs only exist for a short period during the embryo’s development, and they can be isolated and maintained indefinitely in vitro in an undifferentiated state. The discovery of murine ESCs has dramatically changed the field of biomedical research and regenerative medicine over the last 40 years. Since then, enormous investigations have been made to isolate and culture ESCs from other species, including hESCs, in 1998. 99 The success of Thomson et al. in 1998 triggered the great controversy in media and ethical research boards across the globe, with particularly strong objections being raised to the use of human embryos for research purposes. 108 Several studies using hESCs have been conducted demonstrating their therapeutic potential in the clinical setting. However, the use of hESCs is limited due to (1) the ethical barrier related to the destruction of human embryos and (2) the potential risk of immunological rejection, as hESCs are isolated from pre-implantation blastocysts, which are not autologous in origin. To overcome these two great obstacles, several research groups have been trying to develop technology to generate hESCs, including nuclear transfer technology, the well-known strategy that creates Dolly sheep, although the generation of human nuclear transfer ESCs remains technically challenging. 109 Taking a different approach, in 2006, Yamanaka and Takahashi generated artificial PSCs from adult and embryonic mouse somatic cells using four transcription factors ( Oct-3/4 , Sox2 , Klf4 , and c-Myc , called OSKM) reduced from 24 factors. 100 Thereafter, in 2007, Takahashi and colleagues successfully generated the first hiPSCs exhibiting molecular and biological features similar to those of hESCs using the same OSKM factors. 101 Since then, hiPSCs have been widely studied to expand our knowledge of the pathogenesis of numerous diseases and aid in developing new cell-based therapies as well as personalized medicine.

Clinical applications of hPSCs

Since its beginning 24 years ago, hPSC research has evolved momentously toward applications in regenerative medicine, disease modeling, drug screening and discovery, and stem cell-based therapy. In clinical trial settings, the uses of hESCs are restricted by ethical concerns and tight regulation, and the limited preclinical data support their therapeutic potential. However, it is important to acknowledge several successful outcomes of hESC-based therapies in treating human diseases. In 2012, Steven Schwartz and his team reported the first clinical evidence of using hESC-derived retinal pigment epithelium (RPE) in the treatment of Stargardt’s macular dystrophy, the most common pediatric macular degeneration, and an individual with dry age-related macular degeneration. 110 , 111 With a differentiation efficiency of RPE greater than 99%, 5 × 10 4 RPEs were injected into the subretinal space of one eye in each patient. As the hESC source of RPE differentiation was exposed to mouse embryonic stem cells, it was considered a xenotransplantation product and required a lower dose of immunosuppression treatment. This study showed that hESCs improved the vision of patients by differentiating into functional RPE without any severe adverse events. The trial was then expanded into two open-label, phase I/II studies with the published results in 2015 supporting the primary findings. 112 In these trials, patients were divided into three groups receiving three different doses of hESC-derived RPE, including 10 × 10 4 , 15 × 10 4 and 50 × 10 4 RPE cells per eye. After 22 months of follow-up, 19 patients showed improvement in eyesight, seven patients exhibited no improvement, and one patient experienced a further loss of eyesight. The technical challenge of hESC-derived RPE engraftment was an unbalanced proliferation of RPE post administration, which was observed in 72% of treated patients. A similar approach was also conducted in two South Korean patients diagnosed with age-induced macular degeneration and two patients with Stargardt macular dystrophy. 113 The results supported the safety of hESC-derived RPE cells and illustrated an improvement in visual acuity in three patients. Recently, clinical-graded hESC-derived RPE cells were also developed by Chinese researchers under xeno-free culture conditions to treat patients with wet age-related degeneration. 114 As hESC development is still associated with ethical concerns and immunological complications related to allogeneic administration, hiPSC-derived RPE cells have emerged as a potential cell source for macular degeneration. Although RPE differentiation protocols have been developed and optimized to improve the efficacy of hiPSC-derived RPE cells, they are still insufficient, time-consuming and labor intensive. 115 , 116 For clinical application, an efficient differentiation of “primed” to “naïve” state hiPSCs toward the RPE was developed using feeder-free culture conditions utilizing the transient inhibition of the FGF/MAPK signaling pathway. 117 Overexpression of specific transcription factors in hiPSCs throughout the differentiation process is also an interesting approach to generate a large number of RPE cells for clinical use. In a recent study, overexpression of three eye-field transcription factors, including OTX2, PAX6, and MITF , stimulated RPE differentiation in hiPSCs and generated functional RPE cells suitable for transplantation. 118 To date, although reported data from phase I/II clinical trials have been produced enough to support the safety of hESC-derived RPE cells, the treatment is still in its immature stage. Thus, future studies should focus on the development of the cellular manufacturing process of RPE and the subretinal administration route to further improve the outcomes of RPE fabrication and engraftment into the patient’s retina (recommended review 119 ).

Numerous studies have demonstrated that hESC-derived cardiomyocytes exhibit cardiac transcription factors and display a cardiomyocyte phenotype and immature electrical phenotype. In addition, using hPSC-derived cardiomyocytes could provide a large number of cells required for true remuscularization and transplantation. Thus, these cells can be a promising novel therapeutic approach for the treatment of human cardiovascular diseases. In a case report, hESC-derived cardiomyocytes showed potential therapeutic effects in patients with severe heart failure without any subsequent complications. 120 This study was a phase I trial (ESCORT [Transplantation of Human Embryonic Stem Cell-derived Progenitors in Severe Heart Failure] trial) to evaluate the safety of cardiomyocyte progenitor cells derived from hESCs seeded in fibrin gel scaffolds for 10 patients with severe heart failure (NCT02057900). The encouraging results from this study demonstrated the feasibility of producing hESC-derived cardiomyocyte progenitor cells toward clinical-grade standards and combining them with a tissue-engineered scaffold to treat severe heart disease (the first patient of this trial has already reached the 7-year follow-up in October 2021). 121 Currently, the two ongoing clinical trials using hPSC-derived cardiomyocytes have drawn great attention, as their results would pave the way to lift the bar for approving therapies for commercial use. The first trial was conducted by a team led by cardiac surgeon Yoshiki Sawa at Osaka University using hiPSC-derived cardiomyocytes embedded in a cell sheet for engraftment (jRCT2052190081). The trials started first with three patients followed by ten patients to assess the safety of the approach. Once safety is met, the treatment can be sold commercially under Japan’s fast-track system for regenerative medicine. 122 Another trial used a collagen-based construct called BioVAT-HF to contain hiPSC-derived cardiomyocytes. The trial was divided into two parts to evaluate the cell dose: (Part A) recruiting 18 patients and (Part B) recruiting 35 patients to test a broad range of engineered human myocardium (EHM) doses. The expected results from this study will provide the “proof-of-concept” for the use of EHM in the stimulation of heart remuscularization in humans. To date, no adverse events or severe adverse events have been reported from these trials, supporting the safety of the procedure. However, as the number of treated patients was relatively small, limitations in drawing conclusions regarding efficacy are not yet possible. 21 , 123

One of the first clinical trials using hPSC-based therapy was conducted by Geron Corporation in 2010 using hESC-derived oligodendrocyte progenitor cells (OPC1) to treat spinal cord injury (SCI). The results confirmed the safety one year post administration in five participants, and magnetic resonance imaging demonstrated improvement of spinal cord deterioration in four participants. 124 Asterias Biotherapeutic (AST) continued the Geron study by conducting the SCiStar Phase I/IIa study to evaluate the therapeutic effects of AST-OPC1 (NCT02302157). The trial’s results published in clinicaltrials.gov demonstrated significant improvement in running speed, forelimb stride length, forelimb longitudinal deviations, and rear stride frequency. Interestingly, the recently published data of a phase 1, multicentre, nonrandomized, single-group assignment, interventional trial illustrated no evidence of neurological decline, enlarging masses, further spinal cord damage, or syrinx formation in patients 10 years post administration of the OPC1 product. 125 This data set provides solid evidence supporting the safety of OPC1 with an event-free period of up to 10 years, which strengthens the safety profile of the SCiStar trial.

Analysis of the global trends in clinical trials using hPSC-based therapy showed that 77.1% of studies were observational (no cells were administered into patient), and only 22.9% of studies used hPSC-derived cells as interventional treatment. 126 The number of studies using hiPSCs was relatively higher than that using hESCs, which was 74.8% compared to 25.2%, respectively. The majority of observational studies were performed in developed countries, including the USA (41.6%) and France (16.8%), whereas interventional studies were conducted in Asian countries, including China (36.7%), Japan (13.3%), and South Korea (10%). The trends in therapeutic studies were also clear in terms of targeted diseases. The three most studied diseases were ophthalmological conditions, circulatory disorders, and nervous systems. 127 However, it is surprising that the clinical applications of hPSCs have achieved little progress since the first hESCs were discovered worldwide. The relatively low number of clinical trials focusing on using iPSCs as therapeutic agents to administer into patients could be ascribed to the unstable genome of hiPSCs, 128 immunological rejection, 129 and the potential for tumor formation. 130

Mesenchymal stem/stromal cell-based therapy: is it time to consider their origin toward targeted therapy?

Approximately 55 years ago, fibroblast-like, plastic-adherent cells, later named mesenchymal stem cells (MSCs) by Arnold L. Caplan, 18 were discovered for the first time in mouse bone marrow (BM) and were later demonstrated to be able to form colony-like structures, proliferate, and differentiate into bone/reticular tissue, cartilage, and fat. 131 Protocols were subsequently established to directly culture this subpopulation of stromal cells from BM in vitro and to stimulate their differentiation into adipocytes, chondroblasts, and osteoblasts. 132 Since then, MSCs have been found in and derived from different human tissue sources, including adipose tissue (AT), the umbilical cord (UC), UC blood, the placenta, dental pulp, amniotic fluid, etc. 133 To standardize and define MSCs, the International Society for Cell and Gene Therapy (ISCT) set minimal identification criteria for MSCs derived from multiple tissue sources. 134 Among them, MSCs derived from AT, BM, and UC are the most commonly studied MSCs in human clinical trials, 135 and they constitute the three major tissue sources of MSCs that will be discussed in this review.

The discovery of MSCs opened an era during which preclinical studies and clinical trials have been performed to assess the safety and efficacy of MSCs in the treatment of various diseases. The major conclusion of these studies and trials is that MSC-based therapy is safe, although the outcomes have usually been either neutral or at best marginally positive in terms of the clinically relevant endpoints regardless of MSC tissue origin, route of infusion, dose, administration duration, and preconditioning. 136 It is important to note that a solid background of knowledge has been generated from all these studies that has fueled the recent translational research in MSC-based therapy. As MSCs have been intensively studied over the last 55 years and have become the subject of multiple reviews, systematic reviews, and meta-analyses, the objective of this paper is not to duplicate these publications. Rather, we will discuss the questions that both clinicians and researchers are currently exploring with regard to MSC-based therapy, diligently seeking answers to the following:

“With a solid body of data supporting their safety profiles derived from both preclinical and clinical studies, does the tissue origin of MSCs also play a role in their downstream clinical applications in the treatment of different human diseases?”

“Do MSCs derived from AT, BM, and UC exhibit similar efficacy in the treatment of neurological diseases, metabolic/endocrine-related disorders, reproductive dysfunction, skin burns, lung fibrosis, pulmonary disease, and cardiovascular conditions?”

To answer these questions, we will first focus on the most recently published clinical data regarding these targeted conditions, including neurological disorders, pulmonary dysfunctions, metabolic/endocrine-related diseases, reproductive disorders, skin burns, and heart-related diseases, to analyze the potential efficacy of MSCs derived from AT, BM, and UC. Based on the level of clinical improvement observed in each trial, we analyzed the potential efficacy of MSCs derived from each source to visualize the correlation between patient improvement and MSC sources. We will then address recent trends in the exclusive use of MSC-based products, focusing on the efficacy of treatment with MSCs from each of the abovementioned sources, and we will analyze the relationship between the respective efficacies of MSCs from these sources in relation to the targeted disease conditions. Finally, we propose a hypothesis and mechanism to achieve the currently still unmet objective of evaluating the use of MSCs from AT, BM, and UC in regenerative medicine.

An overview of MSC tissue origins and therapeutic potential

In general, MSCs are reported to be isolated from numerous tissue types, but all of these types can be organized into two major sources: adult 137 and perinatal sources 138 (Fig. 2 ). Adult sources of MSCs are defined as tissues that can be harvested or obtained from an individual, such as dental pulp, 139 BM, peripheral blood, 140 AT, 141 lungs, 142 hair, 143 or the heart. 144 Adult MSCs usually reside in specialized structures called stem cell niches, which provide the microenvironment, growth factors, cell-to-cell contacts and external signals necessary for maintaining stemness and differentiation ability. 145 BM was the first adult source of MSCs discovered by Friedenstein 131 and has become one of the most documented and largely used MSC sources to date, followed by AT. BM-MSCs are isolated and cultured in vitro from BM aspirates using a Ficoll gradient-centrifugation method 146 or a red blood cell lysate buffer to collect BM mononuclear cell populations, whereas AT-MSCs are obtained from stromal vascular fractions of enzymatically digested AT obtained through liposuction, 141 lipoplasty, or lipectomy procedures. 147 These tissue collection procedures are invasive and painful for the patient and are accompanied by a risk of infection, although BM aspiration and adipose liposuction are considered safe procedures for BM and AT biopsies. The number of MSCs that can be isolated from these adult tissues varies significantly in a tissue-dependent manner. The percentage of MSCs in BM mononuclear cells ranges from 0.001 to 0.01% following gradient centrifugation. 132 The number of MSCs in AT is at least 500 times higher than that in BM, with approximately 5,000 MSCs per 1 g of AT. Perinatal sources of MSCs consist of UC-derived components, such as UC, Wharton’s jelly, and UC blood, and placental structures, such as the placental membrane, amnion, chorion membrane, and amniotic fluid. 138 The collection of perinatal MSCs, such as UC-MSCs, is noninvasive, as the placenta, UC, UC blood, and amnion are considered waste products that are usually discarded after birth (with no ethical barriers). 148 Although MSCs represent only 10 −7 % the cells found in UC, their higher proliferation rate and rapid population doubling time allow these cells to rapidly replicate and increase in number during in vitro culture. 149 Under standardized xeno-free and serum-free culture platforms, AT-MSCs show a faster proliferation rate and a higher number of colony-forming units than BM-MSCs. 149 UC-MSCs have the fastest population doubling time compared to AT-MSCs and BM-MSCs in both conventional culture conditions and xeno- and serum-free environments. 149 MSCs extracted from AT, BM and UC exhibit all minimal criteria listed by the ISCT, including morphology (plastic adherence and spindle shape), MSC surface markers (95% positive for CD73, CD90 and CD105; less than 2% negative for CD11, CD13, CD19, CD34, CD45, and HLR-DR) and differentiation ability into chondrocytes, osteocytes, and adipocytes. 150

figure 2

The two major sources of MSCs: adult and perinatal sources. The adult sources of MSCs are specific tissue in human body where MSCs could be isolated, including bone marrow, adipose tissue, dental pulp, peripheral blood, menstrual blood, muscle, etc. The perinatal sources of MSCs consist of umbilical cord-derived components, such as umbilical cord, Wharton’s jelly, umbilical cord blood, and placental structures, such as placental membrane, amnion, chorion membrane, amniotic fluid, etc. The figure was created with BioRender.com

In fact, although MSCs derived from either adult or perinatal sources exhibit similar morphology and the basic characteristics of MSCs, studies have demonstrated that these cells also differ from each other. Regarding immunophenotyping, AT-MSCs express high levels of CD49d and low levels of Stro-1. An analysis of the expression of CD49d and CD106 showed that the former is strongly expressed in AT-MSCs, in contrast to BM-MSCs, whereas CD106 is expressed in BM-MSCs but not in AT-MSCs. 151 Increased expression of CD133, which is associated with stem cell regeneration, differentiation, and metabolic functions, 152 was observed in BM-MSCs compared to MSCs from other sources. 153 A recent study showed that CD146 expression in UC-MSCs was higher than that in AT- and BM-MSCs, 153 supporting the observation that UC-MSCs have a stronger attachment and a higher proliferation rate than MSCs from other sources, as CD146 is a key cell adhesion protein in vascular and endothelial cell types. 154 In terms of differentiation ability, donor-matched BM-MSCs exhibit a higher ability to differentiate into chondrogenic and osteogenic cell types than AT-MSCs, whereas AT-MSCs show a stronger capacity toward the adipogenic lineage. 150 The findings from an in vitro differentiation study indicated that BM-MSCs are prone to osteogenic differentiation, whereas AT-MSCs possess stronger adipogenic differentiation ability, which can be explained by the fact that the epigenetic memory obtained from either BM or AT drives the favored MSC differentiation along an osteoblastic or adipocytic lineage. 155 Interestingly, although UC-MSCs have the ability to differentiate into adipocytes, osteocytes, or chondrocytes, their osteogenic differentiation ability has been proven to be stronger than that of BM-MSCs. 156 The most interesting characteristic of MSCs is their immunoregulatory functions, which are speculated to be related to either cell-to-cell contact or growth factor and cytokine secretion in response to environmental/microenvironmental stimuli. MSCs from different sources almost completely inhibit the proliferation of peripheral blood mononuclear cells (PBMCs) at PBMC:MSC ratios of 1:1 and 10:1. 149 At a higher ratio, BM-MSCs showed a significantly higher inhibitory effect than AT- or UC-MSCs. 153 Direct analysis of the immunosuppressive effects of BM- and UC-MSCs has revealed that these cells exert similar inhibitory effects in vitro with different mechanisms involved. 157 With these conflicting data, the mechanism of action related to the immune response of MSCs from different sources is still poorly understood, and long-term investigations both in preclinical studies and in clinical trial settings are needed to shed light on this complex immunomodulation function.

The great concern in MSC-based therapy is the fate of these cells post administration, especially through different delivery routes, including systemic administration via an intravenous (IV) route or tissue-specific administration, such as dorsal pancreatic administration. It is important to understand the distribution of these cells after injection to expand our understanding of the underlying mechanisms of action of treatments; in addition, this knowledge is required by authorized bodies (the Food and Drug Administration (FDA) in the United States or the regulation of advanced-therapy medicinal products in Europe, No. 1394/2007) prior to using these cells in clinical trials. The preclinical data using various labeling techniques provide important information demonstrating that MSCs do not have unwanted homing that could lead to the incorrect differentiation of the cells or inappropriate tumor formation. In a mouse model, human BM-MSCs and AT-MSCs delivered via an IV route are rapidly trapped in the lungs and then recirculate through the body after the first embolization process, with a small number of infused cells found mainly in the liver after the second embolization. 158 Using the technetium-99 m labeling method, intravenously infused human cells showed long-term persistence up to 13 months in the bone, BM compartment, spleen, muscle, and cartilage. 159 A similar result was reported in baboons, confirming the long-term homing of human MSCs in various tissues post administration. 160 Although the retainment of MSCs in the lungs might potentially reduce their systemic therapeutic effects, 161 it provides a strong advantage when these cells are used in the treatment of respiratory diseases. Local injection of MSCs also revealed their tissue-specific homing, as an injection of MSCs via the renal artery route resulted in the majority of the injected cells being found in the renal cortex. 162 Numerous studies have been conducted to track the migration of administered MSCs in human subjects. Henriksson and his team used MSCs labeled with iron sucrose in the treatment of intervertebral disc degeneration. 163 Their study showed that chondrocytes differentiated from infused MSCs could be detected at the injured intervertebral discs at 8 months but not at 28 months. A study conducted in a patient with hemophilia A using In-oxine-labeled MSCs showed that the majority of the cells were trapped in the lungs and liver 1 h post administration, followed by a reduction in the lungs and an increase in the number of cells in the liver after 6 days. 164 Interestingly, a small proportion of infused MSCs were found in the hemarthrosis site at the right ankle after 24 h, suggesting that MSCs are attracted and migrate to the injured site. The distribution of MSCs was also reported in the treatment of 21 patients diagnosed with type 2 diabetes using 18-FDG-tagged MSCs and visualized using positron emission tomography (PET). 165 The results illustrated that local delivery of MSCs via an intraarterial route is more effective than delivery via an IV route, as MSCs home to the pancreatic head (pancreaticoduodenal artery) or body (splenic artery). Therefore, although the available data related to the biodistribution of infused MSCs are still limited, the results obtained from both preclinical and clinical studies illustrate a comparable set of data supporting results on homing, migration to the injured site, and the major organs where infused MSCs are located. The following comprehensive and interesting reviews are highly recommended. 166 , 167 , 168

To date, 1426 registered clinical trials spanning different trial phases have used MSCs for therapeutic purposes, which is four times the number reported in 2013. 169 , 170 As supported by a large body of preclinical studies and advancements in conducting clinical trials, MSCs have been proven to be effective in the treatment of numerous diseases, including nervous system and brain disorders, pulmonary diseases, 171 cardiovascular conditions, 172 wound healing, etc. The outcomes of MSC-based therapy have been the subject of many intensive reviews and systematic analyses with the solid conclusion that these cells exhibit strong safety profiles and positive outcomes in most tested conditions. 173 , 174 , 175 In addition, the available data have revealed several potential mechanisms that could explain the beneficial effects of MSCs, including their homing efficiency, differentiation potential, production of trophic factors (including cytokines, chemokines, and growth factors), and immunomodulatory abilities. However, it is still not known which MSC types should be used for which diseases, as it seems to be that MSCs exhibit beneficial effects regardless of their sources. 169

Acquired brain and spinal cord injury treatment: BM-MSCs have emerged as key players

The theory that brain cells can never regenerate has been challenged by the discovery of newly formed neurons in the human adult hippocampus or the migration of stem cells in the brain in animal models. 176 These observations have triggered hope for regeneration in the context of neuronal diseases by using exogenous stem cell sources to replenish or boost the stem cell population in the brain. Moreover, the limited regenerative capacity of the brain and spinal cord is an obstacle for traditional treatments of neurodegenerative diseases, such as autism, cerebral palsy, stroke, and spinal cord injury (SCI). As current treatments cannot halt the progression of these diseases, studies throughout the world have sought to exploit cell-based therapies to treat neurodegenerative diseases on the basis of advances in the understanding and development of stem cell technology, including the use of MSCs. Successful stem cell therapy for treating brain disease requires therapeutic cells to reach the injured sites, where they can repair, replace, or at least prevent the deteriorative effects of neuronal damage. 177 Hence, the gold standard of cell-based therapy is to deliver the cells to the target site, stimulate the tissue repair machinery, and regulate immunological responses via either cell-to-cell contact or paracrine effects. 178 Among 315 registered clinical trials using stem cells for the treatment of brain diseases, MSCs and hematopoietic stem cells (HSCs; CD34+ cells isolated from either BM aspirate or UC blood) are the two main cell types investigated, whereas approximately 102 clinical trials used MSCs and 62 trials used HSCs for the treatment of brain disease (main search data from clinicaltrial.gov). MSCs are widely used in almost all clinical trials targeting different neuronal diseases, including multiple sclerosis, 179 stroke, 180 SCI, 181 cerebral palsy, 182 hypoxic-ischemic encephalopathy, 183 autism, 184 Parkinson’s disease, 185 Alzheimer’s disease 185 and ataxia. Among these trials in which MSCs were the major cells used, nearly two-thirds were for stroke, SCI, or multiple sclerosis. MSCs have been widely used in 29 registered clinical trials for stroke, with BM-MSCs being used in 16 of these trials. With 26 registered clinical trials, SCI is the second most common indication for using MSCs, with 16 of these trials using mainly expanded BM-MSCs. For multiple sclerosis, 15 trials employed BM-MSCs among a total of 23 trials conducted for the treatment of this disease. Hence, it is important to note that in neuronal diseases and disorders, BM-MSCs have emerged as the most commonly used therapeutic cells among other MSCs, such as AT-MSCs and UC-MSCs.

The outcomes of the use of BM-MSCs in the treatment of neuronal diseases have been widely reported in various clinical trial types. A review by Zheng et al. indicated that although the treatments appeared to be safe in patients diagnosed with stroke, there is a need for well-designed phase II multicentre studies to confirm the outcomes. 173 One of the earliest trials using autologous BM-MSCs was conducted by Bang et al. in five patients diagnosed with stroke in 2005. The results supported the safety and showed an improved Barthel index (BI) in MSC-treated patients. 186 In a 2-year follow-up clinical trial, 16 patients with stroke received BM-MSC infusions, and the results showed that the treatment was safe and improved clinical outcomes, such as motor impairment scale scores. 187 A study conducted in 12 patients with ischemic stroke showed that autologous BM-MSCs expanded in vitro using autologous serum improved the patient’s modified Rankin Scale (mRS), with a mean lesion volume reduced by 20% at 1 week post cell infusion. 188 In 2011, a modest increase in the Fugl Meyer and modified BI scores was observed after autologous administration of BM-MSCs in patients with chronic stroke. 189 More recently, a prospective, open-label, randomized controlled trial with blinded outcome evaluation was conducted, with 39 patients and 15 patients in the BM-MSC administration and control groups, respectively. The results of this study indicated that autologous BM-MSCs with autologous serum administration were safe, but the treatment led to no improvements at 3 months in modified Rankin Scale (mRS) scores, although leg motor improvement was observed. 180 Researchers explored whether the efficacy of BM-MSC administration was maintained over time in a 5-year follow-up clinical trial. Patients (85) were randomly assigned to either the MSC group or the control group, and follow-ups on safety and efficacy were performed for 5 years, with 52 patients being examined at the end of the study. The MSC group exhibited a significant improvement in terms of decreased mRS scores, whereas the number of patients with an mRS score increase of 0–3 was statistically significant. 187 Although autologous BM-MSCs did not improve the Basel index, mRS, or National Institutes of Health Stroke Scale (NIHSS) score 2 years post infusion, patients who received BM-MSC therapy showed improvement in their motor function score. 190 In addition, a prospective, open-label, randomized controlled trial by Lee et al. showed that autologous BM-MSCs primed with autologous “ischemic” serum significantly improved motor functions in the MSC-treated group. Neuroimaging analysis also illustrated a significant increase in interhemispheric connectivity and ipsilesional connectivity in the MSC group. 191 Recently, a single intravenous infection of allogeneic BM-MSCs has been proven to be safe and feasible in patients with chronic stroke with a significant improvement in BI score and NIHSS score. 192

In two systematic reviews using MSCs for the treatment of SCI, BM-MSCs ( n  = 16) and UC-MSCs ( n  = 5) were reported to be safe and well-tolerated. 193 , 194 The results indicated significant improvements in the stem cell administration groups compared with the control groups in terms of a composite of the American Spinal Injury Association (ASIA) impairment scale (AIS) grade, AIS grade A, and ASIA sensory scores and bladder function (Table 1 ). However, larger experimental groups with a randomized and multicentre design are needed for further confirmation of these findings. For multiple sclerosis, several early-phase (phase I/II) registered clinical studies have used BM-MSCs. A study compared the potential efficacy of BM-MSC and BM mononuclear cell (BMMNC) transplantation in 105 patients with spastic cerebral palsy. 195 The results showed that the GMFM (gross motor function measure) and the FMFM (fine motor function measure) scores of the BM-MSC transplant group were higher than those of the BMNNC transplant group at 3, 6, and 12 months of assessment. In terms of autism spectrum disorder, a review of 254 children after BMMNC transplantation found that over 90% of patients’ ISAA (Indian Scale for Assessment of Autism) and CARS (Childhood Autism Rating Scale) scores improved. Young patients and those in whom autism spectrum disorder was detected early generally showed better improvement. 196

One of the biggest limitations when using BM-MSCs is the bone marrow aspiration process, as it is an invasive procedure that can introduce a risk of complications, especially in pediatric and elderly patients. 197 Therefore, UC-MSCs have been suggested as an alternative to BM-MSCs and are being studied in clinical trials for the treatment of neurological diseases in approximately 1550 patients throughout the world; however, only three studies have been completed, with data published recently. 198 A recent study showed that UC-MSC administration improved both gross motor function and cognitive skills, assessed using the Activities of Daily Living (ADL), Comprehensive Function Assessment (CFA), and GMFM, in patients diagnosed with cerebral palsy. The improvements peaked 6 months post administration and lasted for 12 months after the first transplantation. 199 In a single-targeted phase I/II clinical trial using UC-MSCs for the treatment of autism, Riordan et al. reported decreases in Autism Treatment Evaluation Checklist (ATEC) and CARS scores for eight patients, but the paper has been retracted due to a violation of the journal’s guidelines. 200 In an open-label, phase I study, UC-MSCs were used as the main cells to treat 12 patients with autism spectrum disorder via IV infusions. It is important to note that five participants developed new class I anti-human leukocyte antigen in response to the specific lot of manufactured UC-MSCs, although these responses did not exhibit any immunological response or clinical manifestations. Only 50% of participants showed improvements in at least two autism-specific measurements. 201 Although not as widely used as BM-MSCs, these trials have demonstrated the efficacy of using UC-MSCs in the treatment of SCIs. In a pilot clinical study, Yang et al. showed that the use of UC-MSCs has the potential to improve disease status through an increase in total ASIA and SCI Functional Rating Scale of the International Association of Neurorestoratology (IANR-SCIFRS) scores, as well as an improvement in pinprick, light touch, motor and sphincter scores. 202 A study of 22 patients with SCIs showed a potential therapeutic effect in 13 patients post UC-MSC infusion. 203 AT-MSCs were also used to treat SCI, with a single case report indicating an improvement in neurological and motor functions in a domestic ferret patient. 204 However, a result obtained from another phase I trial using AT-MSCs showed mild improvements in neurological function in a small number of patients. 205 A phase II, randomized, double-blind, placebo-controlled, single-center, pilot clinical trial using AT-MSCs in the treatment of acute ischemic stroke published a data set that supports the safety of the therapy, although patients who received AT-MSCs showed a nonsignificant improvement after 24 months of follow-up. 206 In all of the above studies, the safety of using either AT-MSCs or UC-MSCs was evaluated, and no significant reactions were reported after infusion.

Therefore, based on the number of recovered patients post-transplantation and the number of recruited patients in large-scale trials using BM-MSCs, it seems that BM-MSCs are the prominent cells in regard to treating neurodegenerative disease with potentially good outcomes (Table 1 ). It is important to note that we do not negate the fact that AT- and UC-MSCs also show positive outcomes in the treatment of neuronal diseases, with numerous ongoing large-scale, multicentre, randomized, and placebo-control trials, 207 , 208 but we suggest alternative and thoughtful decisions regarding which sources of MSCs are best for the treatment of neuronal diseases and degenerative disorders.

Respiratory disease and lung fibrosis: clinical data support UC as a good source of MSCs

In the last decade, significant increases in respiratory disease incidence due to air pollution, smoking behavior, population aging, and recently, respiratory virus infections such as coronavirus disease 2019 (COVID-19) 209 have been observed, leading to substantial burdens on public health and healthcare systems worldwide. Respiratory inflammatory diseases, including bronchopulmonary dysplasia (BPD), chronic obstructive pulmonary disease (COPD), and acute respiratory distress syndrome (ARDS), have recently emerged as three prevalent pulmonary diseases in children and adults. These conditions are usually associated with inflammatory cell infiltration, a disruption of alveolar structural integrity, a reduction in alveolar fluid clearance ability, cytokine release and associated cytokine storms, airway remodeling, and the development of pulmonary fibrosis. Traditional treatments are focused on relieving symptoms and preventing disease progression using surfactants, artificial respiratory support, mechanical ventilation, and antibiotic/anti-inflammatory drugs, with limited effects on the damaged airway, alveolar fluid clearance, and other detrimental effects caused by the inflammatory response. MSCs are known for their immunomodulatory abilities, showing potential in injury reduction and aiding lung recovery after injury. According to ClinicalTrials.gov, from 2017 to date, there have been 159 studies testing the application of MSCs in the treatment of pulmonary diseases, including but not limited to BPD, COPD, and ARDS, suggesting a trend in the use of MSCs as an alternative approach for the treatment of respiratory diseases, especially MSCs from UC as an “off-the-shelf” and allogeneic source.

Extremely premature infants are born with arrested lung development at the canalicular-saccular phases prior to alveolarization and before pulmonary maturation occurs, which results in the development of BPD. 210 These infants require intensive care during the first three months of life using postnatal interventions, including positive pressure mechanical ventilation, external oxygen support, and surfactant infusions, and the newborns have recurrent infections that further compromise normal lung development. 211 To date, 13 clinical trials have been proposed to use UC-MSCs in the treatment of BPD, recruiting ~566 premature infants throughout the world, including Vietnam, Korea, the United States, Spain, Australia, and China. The majority of these trials use UC-derived stem cells for phases I and II, focusing on evaluating the safety and efficacy of stem cell-based therapy. 212 Human UC tissue and its derivative components are considered the most attractive cell sources for MSCs in the treatment of BPD due to the ease of obtaining them, being readily available, with no ethical concerns, low antigenicity, a high cell proliferation rate, and superior regenerative potential. Chang et al. used MSCs derived from UC blood in a phase I dose-escalation clinical trial to treat 9 preterm infants via intratracheal administration to prevent the development of BPD. 213 All 9 preterm infants survived, and only three developed BPD; these infants had significantly decreased BPD severity compared with the historically matched control group. A follow-up study of the same patients after 24 months indicated that only one infant had an E. cloacae infection after discharge at 4 months, with subsequent disseminated intravascular coagulation, which was later proven to be unrelated to the intervention. The remaining eight patients survived with normal pulmonary development and function, suggesting that the therapy was safe. MSCs from UC blood were also used for the treatment of 12 extremely low birthweight preterm patients using the same administration route, which further confirmed the safety of the therapy in the treatment of BPD, although ten of 12 infants still developed severe BPD at 36 weeks. 214 Our group also reported the safety and potential efficacy of using UC-MSCs in the treatment of four preterm infants, and the results supported the safety of UC-MSCs and demonstrated that patients could be weaned from oxygen supply and develop normal lung structure and function. 215 A phase II clinical trial of 66 infants born at 23–28 weeks with a birthweight of 500–1250 g who were recruited and randomized into an MSC-administration group and a control group was conducted. Although the results supported the safety of MSC administration in preterm infants, the efficacy of the treatment was not supported by statistical analysis, potentially due to the small sample size. Subgroup analysis showed that patients with severe BPD born at 23–24 weeks showed a significant improvement in BPD severity, but those born at 25–28 weeks did not. 216 Hence, it is important to conduct controlled phase II clinical trials with larger cohort sizes to further substantiate the efficacy of UC blood-derived MSCs in the treatment of infants with BPD.

With more than 65 million patients worldwide, COPD was the third-leading cause of death in 2020, according to World Health Organization records. COPD is classified as a chronic inflammatory and destructive pulmonary disease characterized by a progressive reduction in lung function. Averyanov et al. performed a randomized, placebo-controlled phase I/IIa study in 20 patients with mild to moderate idiopathic pulmonary fibrosis (IPF). Treatment group patients received two IV doses of allogeneic MSCs (2 × 10 8  cells) every 3 months, and the second group received a placebo. 217 Evaluation tests were performed at weeks 13, 26, 39, and 52. The 6-min walking test distance (6MWTD) results showed that patient fitness improved from week 13 onwards and was maintained until up to the 52nd week. Pulmonary function indicators improved markedly before and after treatment in the treated group but did not change significantly in the placebo group. The goal of MSC therapy in the treatment of COPD is to promote the regeneration of parenchymal cells and alveolar structure and the restoration of lung function. Based on the results of a phase I trial of a commercial BM-MSC product, Prochymal TM , which led to improvements in pulmonary function in treated patients, a multicentre, double-blind, placebo-controlled phase II trial was conducted in 62 patients diagnosed with COPD to determine the safety and potential efficacy of the product. Although the results supported the safety of BM-MSCs, their effectiveness in the treatment of COPD was not assured. No statistically significant differences in FEV 1 or FEV 1% , total lung capacity, or carbon monoxide diffusing capacity were detected after 2 years of follow-up between the two treatment groups. To date, there have been five clinical trials using BM-MSCs as the main stem cells for the treatment of COPD, but the overall clinical outcomes did not demonstrate the potential therapeutic effects of the treatment. 218 , 219 , 220 , 221 , 222 In clinical trial NCT001110252, the results showed that there was an overall reduction in the process of COPD pathological development 3 years after the administration of BM-MSCs, although the trial had a phase I design, with no control group, and evaluated only a small cohort (four patients). 219 To alleviate local inflammatory progression in COPD, Oliveira et al. studied the combination treatment of one-way endobronchial valve (EBV) and BM-MSC intubation. 223 Ten GOLD (Global Initiative for Obstructive Lung Disease) stage C or D patients were equally divided into 2 groups: one group received a dose of 10 8 cells before valve insertion, and the other group received a normal saline infusion. The follow-up time was 90 days. Inflammation was significantly improved as assessed by the CRP (C-reactive protein) index at 30 and 90 days after infusion. In addition, improvements in St. George’s Respiratory Questionnaire (SGRQ) scores indicated improved patient quality of life. Furthermore, an investigation into the homing ability of MSCs in vivo was performed on 9 GOLD patients, from stage A to stage D. Each patient received two 2 × 10 6 BM-MSC/kg IV infusions 1-week apart. 224 The marking of MSCs with indium-111 showed that MSCs were retained in the pulmonary vasculature longer in patients with mild COPD and that the levels of inflammatory mediators improved after 7 days of treatment. The results of the evaluation survey conducted after 1 year showed that the number of COPD exacerbations decreased to six times/year compared to 11 times/year before treatment. In addition, AT-MSCs present in the stromal vascular fraction were used to treat patients with COPD, and no adverse events were observed after 12 months of follow-up, but the clinical improvements post administration were not clear. 225 The results from a phase I clinical trial using AT-MSCs in eight patients with COPD also reported no significant change in pulmonary function test parameters. 226 A study evaluating the use of AT-MSCs as adjunctive therapy for COPD in 12 patients was performed. 227 AT was obtained using standard liposuction, MSCs were isolated, and 150–300 million cells were intravenously infused. The patients showed improvements in quality of life, with improved SGRQ scores after 3 and 6 months of treatment. Recently, UC-MSCs have emerged as potential allogeneic stem cell candidates for the treatment of COPD. 228 In a pilot clinical study, it was demonstrated that allogeneic administration of UC-MSCs in the treatment of COPD was safe and potentially effective. 229 In one study, 20 patients, including 9 at stage C and 11 at stage D per the GOLD classification, with histories of smoking were recruited and received cell-based therapy. The patients who received UC-MSC treatment showed significant reductions in Modified Medical Research Council scores, COPD assessment test scores, and the number of pulmonary exacerbations 6 months post administration. The results of the second trial using UC-MSCs showed that the mean FEV 1 /FVC ratios were increased along with improvements in SGRQ scores and 6MWTDs at three months post administration. 230 Although thorough assessments of the effectiveness of UC-MSCs are still in the early stages, the number of trials using UC-MSCs for the treatment of COPD is increasing steadily, with larger sample sizes and stronger designs (randomized or matched case–control studies), providing a data set strongly supporting the future applications of UC-MSCs. 231

The ongoing pandemic of the 21st century, the COVID-19 pandemic, emerged as a major pulmonary health problem worldwide, with a relatively high mortality rate. Numerous studies, reviews, and systematic analyses have been conducted to discuss and expand our knowledge of the virus and propose different mechanisms by which the virus could alter the immune system. 232 One of the most critical mechanisms is the generation of cytokine storms, which result from the initiation of hyperreactions of the adaptive immune response to viral infection. 233 These cytokine storms are formed by the establishment of waves of hypercytokinaemia generated from overreactive immune cells, which enhance their expression of TNF-α, IL-6, and IL-10, preventing T-lymphocyte recruitment and proliferation and culminating in T-lymphocyte apoptosis and T-cell exhaustion. In COVID-19, once a cytokine storm is formed, it spreads from an initial focal area through the body via circulation, which has been discussed in a comprehensive review by Jamilloux et al. 234 At the time of writing this review, there were 74 clinical trials using MSCs from UC (29 trials; including WJ-derived MSCs (WJ-MSCs) and placenta-derived MSCs (PL-MSCs)), AT (15 trials), and BM (11 trials) (comprehensive review 171 , 235 ). Hence, UC-MSCs have emerged as the most common MSCs for the treatment of COVID-19, with a total of 1047 patients participating in these trials. Among these trials, 15 completed trials using UC-MSCs (including WJ- and PL-MSCs) have been reported, with clinical data from approximately 600 recruited patients. 232 Eight of these 15 studies used allogenic UC-MSC transplantation to treat critically ill patients. 236 A list of case reports using UC-MSCs showed that the treatments were safe and well-tolerated in 14 patients with COVID-19, with the primary outcomes including increased percentages and numbers of T cells, 237 , 238 improved respiratory and renal functions, 239 reductions in inflammatory biomarker levels, 240 and positive outcomes in the PaO 2 /FiO 2 ratio. 240 In a pilot study conducted in ten patients with severe COVID-19, a single dose of UC-MSCs was safe and improved clinical outcomes, although the study did not investigate whether multiple doses of UC-MSCs could further improve the outcomes. 241 Two trials without a control group were conducted in 47 patients, and the results indicated that UC-MSCs were safe and feasible for the treatment of patients with COVID-19. 235 , 242 A single-center, open-label, individually randomized, standard treatment-controlled trial was performed in 41 patients (12 patients assigned to the UC-MSC group), and the results showed that significant improvements in C-reactive protein levels, IL-6 levels, oxygen indices, and lymphocyte numbers were found in the MSC groups. Chest computed tomography (CT) illustrated significant reductions in lung inflammatory responses as reflected by CT findings, the number of lobes involved, and pulmonary consolidation. 238 In a phase I trial conducted in 18 hospitalized patients with COVID-19, UC-MSCs were administered via an IV route in nine patients (five patients with moderate COVID-19 and 4 patients with severe COVID-19) at days 0, 3, and 6, with no treatment-related adverse events or severe adverse events. 243 Only one patient in the UC-MSC group required mechanical ventilation, compared to four patients in the control group. However, the clinical outcomes, such as COVID-19 symptoms, laboratory test results, CT findings of lung damage, and pulmonary function test parameters, were improved in both groups. Interestingly, a 1-year follow-up of the same sample revealed that the patients who received UC-MSC administration improved in terms of whole-lung lesion volume compared to the control group. 244 Moreover, chest CT at 12 months showed significant regeneration of lung tissue in the MSC-administered groups, whereas lung fibrosis was found in all patients in the control group. This finding is of interest because it indicates that a long time is needed to detect the regenerative functions of MSC-based therapy, as the biological process to enhance lung tissue regeneration occurs relatively slowly and requires multiple steps. The effects of UC-MSCs in the attenuation and prevention of the development of cytokine storms were illustrated in an interventional, prospective, three-parallel arm study with two control arms conducted in 30 patients in moderate and critical clinical conditions. 245 The results indicated a significant decrease in proinflammatory cytokines (IFNγ, IL-6, IL-17A, IL-2, and IL-12) and an increase in anti-inflammatory cytokines (IL-10, IL-13, and IL-1ra), suggesting that UC-MSCs might participate in the prevention of cytokine storm development. Lanzoni et al. performed a double-blind, randomized, controlled trial and found that UC-MSC infusions significantly decreased cytokine levels at day 6 and improved survival in patients with COVID-19 with ARDS. In this trial, 24 patients were randomized and assigned 1:1 to receive either MSCs or placebo. 246 MSC treatment was associated with a significant improvement in the survival rate without serious adverse events. To date, other trials conducted using UC-MSCs as the main MSCs provide a solid data set on their safety and efficacy in preventing the development of cytokine storms, reducing the inflammatory response, improving pulmonary function, reducing intensive care unit (ICU) stay duration, enhancing lung tissue regeneration, and reducing lung fibrosis progression. 240 , 247 , 248 , 249 In two large cohort studies (phase I with 210 patients and phase II with 100 patients), the volume of lung lesions and solid component injuries of patients’ lungs were reduced significantly after the administration of UC-MSCs, 250 and clinical symptoms and inflammatory levels were improved. 251 Of the 26 reported clinical trials for the treatment of COVID-19 with MSCs, 1 study used AT-MSCs as the main MSCs. 236 Thirteen COVID-19 adult patients under invasive mechanical ventilation who had received previous antiviral and/or anti-inflammatory treatments (including steroids, lopinavir/ritonavir, hydroxychloroquine, and/or tocilizumab, among others) were treated with allogeneic AT-MSCs. With a mean follow-up time of 16 days after infusion, 9/13 patients’ clinical symptoms improved, and 7/13 patients were intubated. A decrease in inflammatory cytokines and an increase in immunoregulatory cells were also observed in patients, especially in the group of patients with overall clinical improvement. Although there is a lack of clinical efficacy data supporting the use of AT-MSCs in the treatment of patients with COVID-19, AT-MSCs are still potential candidates for inhibiting COVID-19 due to their high secretory activity, strong immune-modulatory effects, and homing ability. 252 , 253 , 254

For ARDS, in a phase IIa trial, 60 patients with moderate to severe disease were randomized into 2 groups. A group of 40 patients received a single infusion of BM-MSCs at a dose of 1 × 10 6 cells/kg body weight, and another 20 patients received a placebo. 255 After 6 and 24 h of infusion, the decrease in plasma inflammatory cytokine levels in the MSC group was significantly greater than that in the placebo group. For severe pulmonary hypertension (PH) associated with BPD (BPD-PH), in a small trial, two preterm infants born at 26–27 weeks of age were intravenously administered heterologous BM-MSCs at a dose of 5 × 10 6 cells per kg of body weight; the treatment reduced oxygen requirements and supported respiration in the infants. 256 The administration of allogeneic AT-MSCs in the treatment of ARDS appeared to be safe and well-tolerated in 12 adult patients, but clinical outcomes were not observed. 257 The results of two patients who received BM-MSCs showed that both patients had improved respiratory function and hemodynamic function and a reduction in multiorgan failure. 258 Although the safety of BM-MSCs was confirmed in a multicentre, open-label, dose-escalation, phase I clinical trial (The Stem cells for ARDS treatment—START trial), 259 no significant improvements were found in a phase II trial, including in respiratory function and ARDS conditions. 260 The safety profile of UC-MSCs is also supported by the findings of a previous phase I clinical trial conducted in 9 patients, which showed that a single IV administration of UC-MSCs was safe and led to positive outcomes in terms of respiratory function and a reduction in the inflammatory response. 261 The findings of this study were also supported by those of the REALIST (Repair of Acute Respiratory Distress with Stromal Cell Administration) trial, which further confirmed the maximum tolerated dose of allogeneic UC-MSCs in patients with moderate to severe ARDS. 262

Although AT- and BM-MSCs have demonstrated therapeutic potential with similar mechanisms of action, UC-MSCs have emerged as potential candidates in the treatment of pulmonary diseases due to their ease of production as “off-the-shelf” products, rapid proliferation, noninvasive isolation methods, and supreme immunological regulation as well as anti-inflammatory effects. 263 However, it is important to note that there is a need to conduct phase III clinical trials with larger cohorts and trials with at least two sources of MSCs in the treatment of pulmonary conditions to further confirm this speculation. 264 Table 2 summarizes several clinical trials with published results discussed in this review.

Endocrine disorders, infertility/reproductive function recovery, and skin burns: should we consider AT-MSCs as the main MSCs based on their origin?

Endocrine disorders.

The human body maintains function and homeostatic regulation via a complex network of endocrine glands that synthesize and release a wide range of hormones. The endocrine system regulates body functions, including heartbeat, bone regeneration, sexual function, and metabolic activity. Endocrine system dysregulation plays a vital role in the development of diabetes, thyroid disease, growth disorder, sexual dysfunction, reproductive malfunction, and other metabolic disorders. The central dogma of regenerative medicine is the use of adult stem cells as a footprint for tissue regeneration and organ renewal. The functions of these stem cells are tightly regulated by microenvironmental stimuli from the nervous system (rapid response) and endocrine signals via hormones, growth factors, and cytokines. This harmonized and orchestrated system creates a symphony of signals that directly regulate tissue homeostasis and repair after injury. The disruption of these complex networks results in an imbalance of tissue homeostasis and regeneration that can lead to the development of endocrine disorders in humans, such as diabetes, sexual hormone deficiency, premature ovarian failure (POF), and Asherman syndrome.

In recent years, obesity and diabetes (type 1 diabetes mellitus (T1DM) and type 2 diabetes mellitus (T2DM)) have been the two biggest challenges in endocrinology research, and the application of MSCs has emerged as a novel approach for therapeutic consideration. T1DM is characterized by the autoimmune destruction of pancreatic β-cells, whereas T2DM is defined as a combination of insulin resistance and pancreatic insulin-producing cell dysfunction. Regenerative medicine seeks to provide an exogenous cell source for replacing damaged or lost β-cells to achieve the goal of stabilizing patients’ blood glucose levels. To date, there are 28 clinical trials using MSCs in the treatment of T1DM ( http://www.clinicaltrials.gov , searched in October 2021), among which three trials were completed using autologous BM-MSCs (NCT01068951), allogeneic BM-MSCs (NCT00690066), and allogeneic AT-MSCs (NCT03920397). Interestingly, UC-MSCs were the most favored MSCs for the remaining trials. All published studies confirmed the safety of MSC therapy in the treatment of T1DM with no adverse events. The first study using autologous BM-MSCs showed that patients who were randomized into the MSC-administration group showed an increase in C-peptide levels in response to a mixed-meal tolerance test (MMTT) in comparison to the control group. 265 Unfortunately, there was no significant improvement in C-peptide levels, HbA1 C or insulin requirements. The use of autologous AT-MSCs in combination with vitamin D was safe and improved HbA1 C levels 6 months post administration. 266 WJ-MSCs were used as the main MSCs for the treatment of new-onset T1DM, which showed a significant improvement in both HbA1 C and C-peptide levels when compared to those of the control group at three and six months post administration. 267 , 268 The combination of allogeneic WJ-MSCs with autologous BM-derived mononuclear cells improved insulin secretion and reduced insulin requirements in patients with T1DM. 269 In terms of T2DM, 23 studies were registered on clinicaltrials.gov (searched in October 2021), with six completed studies (three studies used BM-MSCs and three studies used allogeneic UC-MSCs). Although the number of studies using MSCs for the treatment of T2DM is small, their findings support the safety of MSCs, with no severe adverse events observed during the course of these studies. 270 It was confirmed that MSC therapy potentially reduced fasting blood glucose and HbA1 C levels and increased C-peptide levels. However, these effects were short-term, and multiple doses were required to maintain the MSC effects. Interestingly, the autologous MSC approach in the treatment of patients with diabetes in general is hampered, as both BM-MSCs and AT-MSCs isolated from patients with diabetes showed reduced stemness and functional characteristics. 271 , 272 In addition, the durations of diabetes and obesity are strongly associated with autologous BM-MSC metabolic function, especially mitochondrial respiration, and the accumulation of mitochondrial DNA, which directly interfere with the functions of BM-MSCs and reduce the effectiveness of the therapy. 271 Therefore, the allogeneic approach using MSCs from healthy donors provides an alternative approach for stem cell therapy in the treatment of patients with diabetes.

Infertility and reproductive function recovery

Modern society is increasingly facing the problem of infertility, which is defined as the inability to become pregnant after more than 1 year of unprotected intercourse. 273 This problem has emerged as an important worldwide health issue and social burden. Assisted reproductive techniques and in vitro fertilization technology have recently become the most effective methods for the treatment of infertility in humans, but the use of these approaches is limited, as they cannot be applied in patients with no sperm or those who are unable to support implantation during pregnancy, they are associated with complications, they are time-consuming and expensive, and they are associated with ethical issues in certain territories. 274 Numerous conditions are related to infertility, including POF, nonobstructive azoospermia, endometrial dysfunction, and Asherman syndrome. Recent progress has been illustrated in preclinical studies for the potential applications of stem cell-based therapy for reproductive function recovery, especially recent studies in the field of MSCs, which provide new hope for patients with infertility and reproductive disorders. 275

POF is characterized by a loss of ovarian activity during middle age (before 40 years old) and affects 1–2% of women of reproductive age. 276 Patients diagnosed with POF exhibit oligo-/amenorrhea for at least 4 months, with increased levels of follicle-stimulating hormone (FSH) (>25 IU/L) on two occasions more than 1 month apart. 277 Diverse factors, such as genetic backgrounds, autoimmune disorders, environmental conditions, and iatrogenic and idiopathic situations, have been reported to be the cause of POF. 278 POF can be treated with limited effectiveness via psychosocial support, hormone replacement intervention, and fertility management. 279 MSCs from AT, BM, and UC have been used in the treatment of POF, with improvements in ovarian function in preclinical studies using chemotherapy-induced POF animal models. The early published POF study using BM-MSCs as the main cell source is a single case report in which a perimenopausal woman showed an improvement in follicular regeneration, and increased AMH levels resulted in a successful pregnancy followed by delivery of a healthy infant. 280 A report using autologous BM-MSCs in two women with POF illustrated an increase in baseline estrogen levels and the volume of the treated ovaries along with amelioration of menopausal symptoms. 281 The clinical procedures used in this early trial were invasive, as patients underwent two operations: (1) BM aspiration and (2) laparoscopy. A similar approach was used in two trials conducted in 10 women with POF (age range from 26–33 years old) and 30 patients (age from 18 to 40 years old). 282 A later study investigated two different routes of cell delivery, including laparoscopy and the ovarian artery, but the results have not been reported at this time. 282 Based on the positive outcomes of the mouse model, an autologous stem cell ovarian transplantation (ASCOT) trial was deployed using BM-derived stem cells with encouraging observations of improved ovarian function, as determined by elevated levels of AMH and AFC in 81.3% of participants, six pregnancies, and the successful delivery of three healthy babies. 283 A randomized trial (NCT03535480) was conducted in 20 patients with POF aged less than 39 years to further elaborate on the results of the ASCOT trial. 284 To date, there are no completed trials using AT-MSCs or UC-MSCs in the treatment of patients with POF, limiting the evaluation of these MSCs in the treatment of POF. The speculated reason is that POF is a rare disease, affecting 1% of women younger than 40 years, and with improvements in assisted productive technology, patients have several alternative options to enhance the recovery of reproductive function. 285

Wound healing and skin burns

Burns are the fourth most common injury worldwide, affecting ~11 million people, and are a major cause of death (180,000 patients annually). The severity of burns is defined based on the percentage of surface area burned, burn depth, burn location and patient age, and burns are usually classified into first-, second-, third-, and fourth-degree burns on the basis of their severity. 286 Postburn recovery depends on the severity of the burn and the effectiveness of treatment. Rapid healing may occur over weeks, while alternatively, healing can take months, with the ultimate result being scar formation and disability in patients with severe burns. Different from mechanical injury, burn injury is an invasive progression of damage to tissue at the burn site, including both mechanical damage to the skin surface and biological damage caused by natural apoptosis that prolongs excessive inflammation, oxidative stress, and impaired tissue perfusion. 287 To date, completely reversing the devastating damage of severe burns remains unachievable in medicine, and stem cell therapy provides an alternative option for patients with burn injury. The first case report of the use of BM-MSCs to treat a 45-year-old patient with burns on 40% of their body demonstrated the safety of the therapy and showed partial improvements in vascularization at the wound site and reduced coarse cicatrices. 288 , 289 Later, patients with second- and third-degree burns as well as deep burns were treated using either autologous BM-MSCs or allogeneic BM-MSCs by spraying the MSCs onto the burn sites or adding MSCs over a dermal matrix sheet to cover the wound. The results in these case reports revealed the potential efficacy of MSC-based therapy, which not only enhanced the speed of wound recovery but also reduced pain and improved blood supply without introducing infection. 288 , 290 , 291 In 2017, a study conducted in 60 patients with 10–25% of their total body surface areas burned treated with either autologous BM-MSCs or UC-MSCs showed that both MSC types improved the rate of healing and reduced the hospitalization period. 292 The drawback of BM-MSCs in the treatment of burns is the invasive harvesting method, which causes pain and possible complications in patients. Hence, treatment with allogeneic MSCs obtained from healthy donors is the method of choice, and AT- and UC-MSCs are two suitable candidates for this option. To date, a limited number of clinical trials have been conducted using MSC therapy. These trials have several limitations in trial design, such as a lack of a negative control group and blinding, small sample sizes, and the use of standardized measurement tools for burn injury and wound healing. Currently, AT-MSCs are being used in seven ongoing phase I and II trials in the treatment of burns. Hence, it is important to note that among the most widely studied MSCs, AT-MSCs have advantages over BM-MSCs when obtained from an allogeneic source, while their abilities in burn treatment remain to be determined. The main MSCs that should be used in the regeneration of burn tissue remain undefined (Table 3 ), and we observed the trend that AT-MSCs are more suitable candidates due to their biological nature, which contributes to the generation of keratinocytes and secretion profiles that strongly enhance the skin regeneration process. 293 , 294 , 295 , 296

MSC applications in cardiovascular disease: a promising but still controversial field

In the last two decades, great advancements have been achieved in the development of novel regenerative medicine and cardiovascular research, especially stem cell technology. 297 The discovery of human embryonic stem cells and human induced pluripotent stem cells (hiPSCs) opened a new door for basic research and therapeutic investigation of the use of these cells to treat different diseases. 298 However, the clinical path of hiPSCs and hiPSC-derived cardiomyocytes in the treatment of cardiovascular diseases is limited due to the potential for teratoma formation with hiPSCs and the immaturity of hiPSC-derived cardiomyocytes, which might pose a risk of cancer formation, 299 arrhythmia, and cardiac arrest to patients. 300 A recently emerged stem cell type is adult stem cells/progenitor cells, including MSCs, which can stimulate myocardial repair post administration due to their paracrine effects. Promising results of MSC-based therapy obtained from preclinical studies of cardiac diseases enhance the knowledge and strengthen the clinical research to investigate the safety and efficacy in a clinical trial setting. There are papers that discuss the importance of MSC therapy in the treatment of cardiovascular diseases, with the following references being highly recommended. 301 , 302 , 303 , 304 , 305 , 306 To date, 36 trials have evaluated the therapeutic potential of MSCs in different pathological conditions, with the most prevalent types being BM-MSCs (25 trials), followed by UC-MSCs (7 trials) and AT-MSCs (4 trials). 303 However, the reported results are contradictory and create controversy about the efficacy of the treatments.

One of the first trials using MSCs in the treatment of chronic heart failure was the Cardiopoietic Stem Cell Therapy in Heart Failure (C-CURE) trial, a multicentre, randomized clinical trial that recruited 47 patients. The trial findings supported the safety of BM-MSC therapy and provided a data set that demonstrated improvements in cardiovascular scores along with New York Heart Association functional class, quality of life, and general physical health. 307 Despite these encouraging results in the phase I trial, the treatment failed to achieve the primary outcomes in the phase II/III trial (CHART-1 trial), including no significant improvements in cardiac structure or function or patient quality of life. 308 A positive outcome was also found in a phase I/II, randomized pilot study called the POSEIDON trial, which was the first trial to demonstrate the superior effectiveness of the administration of allogeneic BM-MSCs compared to allogeneic MSCs from other sources. 309 , 310 Published results from the MSC-HF study, with 4 years of follow-up results, 311 , 312 and the TRIDENT study 313 illustrated the positive outcomes of BM-MSCs in the treatment of heart failure. However, a contradictory result from the recently published CONCERT-HF trial demonstrated that the administration of autologous BM-MSCs to patients diagnosed with chronic ischemic heart failure did not improve left ventricular function or reduce scar size at 12 months post administration, but the patient’s quality of life was improved. 314 This observation is similar to that of the TAC-HFT trial 315 but completely different from the reported results of the MSC-HF trial. A comprehensive investigation is still needed to determine the reasons behind these contradictory results. The largest clinical trial to date using BM-MSCs is the DREAM-HF study, which was a randomized, double-blind, placebo-controlled, phase III trial that was conducted at 55 sites across North America and recruited a total of 565 patients with ischemic and nonischaemic heart failure. 172 Although recent reports from the sponsor confirmed that the trial missed its primary endpoint (a reduction in recurrent heart failure-related hospitalization), other prespecified endpoints were met, such as a reduction in overall major adverse cardiac events (including death, myocardial infarction, and stroke). 306 Thus, a complete report from the DREAM-HF trial will provide pivotal data supporting the therapeutic potential of BM-MSCs in the treatment of heart failure and open a new path for the FDA to approve cell-based therapy for cardiovascular diseases.

The early trial using AT-derived cells was the PRECISE trial, which was a phase I, randomized, placebo-controlled, double-blind study that examined the safety and efficacy of adipose-derived regenerative cells (ADRCs) in the treatment of chronic ischemic cardiomyopathy. 316 ADRCs are a homogenous population of cells obtained from the vascular stromal fraction of AT, which contains a small proportion of AT-MSCs. 317 Although the study supported the safety of ADRC administration and illustrated a preserved functional capacity (peak VO 2 ) in the treated group and improvements in heart wall motion, neither poor left ventricle (LV) volume nor poor left ventricular ejection fraction (LVEF) was ameliorated. The follow-up trial of the PRECISE trial, called the ATHENA trial, was conducted in 31 patients, although the study was terminated prematurely because two cerebrovascular events occurred, which were not related to the cell product itself. 318 The results of the study illustrated increases in functional capacity, hospitalization rate, and MLHFQ scores, but the LV volume and LVEF were not significantly different between the two groups. Kastrup and colleagues conducted the first in vitro expanded AT-MSC trial in ten patients with ischemic heart disease and ischemic heart failure in 2017. The results confirmed that ready-to-use AT-MSCs were well-tolerated and potentially effective in the treatment of ischemic heart disease and heart failure. 319 Comparable results of AT-MSCs were also reported from the MyStromalCell Trial, which was a randomized placebo-controlled study. In this trial, 61 patients were randomized at a 2:1 ratio into two groups, with the results showing no significant difference in the primary endpoint, which was a change in the maximal bicycle exercise tolerance test (ETT) score from baseline to 6 months post administration. 320 A 3-year follow-up report from the MyStromalCell Trial confirmed that patients who received AT-MSC administration maintained their preserved exercise capacity and their cardiac symptoms improved, whereas the control group experienced a significant reduction in exercise performance and a worsened cardiovascular condition. 321

UC-MSCs are potential allogeneic cells for the treatment of cardiovascular disease, as they are “ready to use” and easy to isolate, they rapidly proliferate, and they secrete hepatocyte growth factors, 322 which are involved in cardioprotection and cardiovascular regeneration. 323 The pilot study using UC-MSCs in 30 patients with heart failure, called the RIMECARD trial, was the first reported trial for which the results supported the effectiveness of UC-MSCs, as seen in the improved ejection fraction, left ventricular function, functional status, and quality of life in patients administered UC-MSCs. 324 Encouraging results reported from a phase I/II HUC-HEART trial 325 showed improvements in LVEF and reductions in the size of the injured area of the myocardium. However, the opposite observations were also reported from a recently published phase I randomized trial using a combination of UC-MSCs and a collagen scaffold in patients with ischemic heart conditions, in which the size of fibrotic scar tissue was not significantly reduced. 326

Although MSCs from AT, BM, and UC have proven to be safe and feasible in the treatment of cardiovascular diseases, the correlation between the MSC types and their therapeutic potentials is still uncertain because different results have been reported from different clinical trials (Table 4 ). The mechanisms by which MSCs participate in recovery and enhance myocardial regeneration have been discussed comprehensively in a recently published review; 305 , 327 therefore, they will not be discussed in this review. In fact, the challenges of MSC-based therapy in cardiovascular diseases have been clearly described previously, 328 including (1) the lack of an in vitro evaluation of the transdifferentiation potential of MSCs to functional cardiac and endothelial cells, 329 (2) the uncontrollable differentiation of MSCs to undesirable cell types post administration, 330 and (3) the undistinguishable nature of MSCs derived from different sources with various levels of differentiation potential. 331 Therefore, the applications of MSC-based therapy in cardiovascular disease are still in their immature stage, with potential benefits to patients. Thus, there is a need to conduct large-scale, well-designed randomized clinical trials not only to confirm the therapeutic potential of MSCs from various sources but also to enhance our knowledge of cardiovascular regeneration post administration.

Proposed mechanism of BM-MSCs in the treatment of acquired brain and spinal injury

Bones are complex structures constituting a part of the vertebrate skeleton, and they play a vital role in the production of blood cells from HSCs. Similar to the functions of most vertebrate organs, bone function is tightly regulated by its constituents and by long-range signaling from AT and the adrenal glands, parathyroid glands, and nervous system. 332 The central nervous system (CNS) orchestrates the voluntary and involuntary input transmitted by a network of peripheral nerves, which act as the bridge between the nervous system and target organs. The CNS controls involuntary responses via the autonomic nervous system (ANS), consisting of the sympathetic nervous system and the parasympathetic nervous system, and voluntary responses via the somatic nervous system. The ANS penetrates deep into the BM cavity, reaching the regions of hematopoietic activity to deliver neurotransmitters that tightly regulate BM stem cell niches. 333 The BM microenvironment consists of various cell types that participate in the maintenance of HSC niches, which are composed of specialized cells, including BM-MSCs (Fig. 3a ). The release of a specific neurotransmitter, circadian norepinephrine, from the sympathetic nervous system at nerve terminals leads to a reduction in the circadian expression of C–X-C chemokine ligand 12 (CXCL12, which is also known as stromal cell-derived factor-1 (SDF-1)) by Nestin + /NG2 2+ BM-MSCs, resulting in the secretion of HSCs into the peripheral bloodstream. 334 , 335 In fact, BM-MSCs play a significant role in the regulation of HSC quiescence and are closely associated with arterioles and sympathetic nervous system nerve fibers. Nestin-expressing BM-MSCs have been shown to express high levels of SDF-1, stem cell factor (SCF), angiopoietin-1 (Ang-1), interleukin-7, vascular cell adhesion molecule 1 (VCAM-1), and osteopontin (OPN), which are directly involved in the regulation and maintenance of HSC quiescence. 336 The depletion of BM-MSCs in BM leads to the mobilization of HSCs into the peripheral bloodstream and spleen. The findings from a previous study demonstrated that reduced SDF-1 expression in norepinephrine-treated BM-MSCs resulted in the mobilization of CXCR4 + HSCs into circulation. 337 The ability of BM-MSCs to produce SDF-1 is tightly related to their neuronal protective functions. 338 SDF-1 is a member of a chemokine subfamily that orchestrates an enormous diversity of pathways and functions in the CNS, such as neuronal survival and proliferation. The chemokine has two receptors, CXCR4 and CXCR7, that are involved in the pathogenic development of neurodegenerative and neuroinflammatory diseases. 339 In the damaged brain, SDF-1 functions as a stem cell homing signal, and in acquired immune deficiency syndrome (AIDS), SDF-1 has been reported to be involved in the protection of damaged neurons by preventing apoptosis. In a traumatic brain injury model, SDF-1 was found to function as an inhibitor of the caspase-3 pathway by upregulating the Bcl-2/Bax ratio, which in turn protects neurons from apoptosis. 340 Moreover, the release of SDF-1 also facilitates cell recruitment, cell migration, and the homing of neuronal precursor cells in the adult CNS by activating the CXCR4 receptor. 341 , 342 Existing data support that SDF-1 acts as the guiding signal for the regeneration of axon growth in damaged neurons and enhances spinal nerve regeneration. 343 , 344 Hence, the ability of BM-MSCs to express SDF-1 in response to the neuronal environment provides a unique neuronal protective effect that could explain the potential therapeutic efficacy of BM-MSCs in the treatment of neurodegenerative diseases (Fig. 3b ).

figure 3

The nature of the “stem niche” of bone marrow-derived mesenchymal stem cells (BM-MSCs) supports their therapeutic potential in neuron-related diseases. a Bone marrow is a complex stem cell niche regulated directly by the central nervous system to maintain bone marrow homeostasis and haematopoietic stem cell (HSC) functions. MSCs in bone marrow respond to the environmental changes through the release of norepinephrine (NE) from the sympathetic nerves that regulate the synthesis of SDF-1 and the migration of HSCs through the sinusoids. The secretion of stem cell factors (SCFs), VCAM-1 and angiotensin-1 from MSCs also plays a significant role in the maintenance of HSCs. b BM-MSCs have the ability to produce and release SDF-1, which directly contributes to neuroprotective functions at the damaged site through interaction with its receptors CXCR4/7, located on the neuronal membrane. c Neuronal protection and the functional remyelination induced by BM-MSCs are also modulated by the release of a wide range of growth factors, including VEGF, BDNF, and NGF, by the BM-MSCs. d BM-MSCs also have the ability to regulate neuronal immune responses by direct interaction or paracrine communication with microglia. Figure was created with BioRender.com

The migration of exogenous MSCs after systemic administration to the brain is limited by the physical blood–brain barrier (BBB), which is a selective barrier formed by CNS endothelial cells to restrict the passage of molecules and cells. The mechanism of molecular movement across the BBB is well established, but how stem cells can bypass the BBB and home to the brain remains unclear. Recent studies have reported that MSCs are able to migrate through endothelial cell sheets by paracellular or transcellular transport followed by migration to the injured or inflammatory site of the brain. 345 , 346 During certain injuries or ischemic events, such as brain injury, stroke, or cerebral palsy, the integrity and efficiency of BBB protection is compromised, which allows MSC migration across the BBB via paracellular transport through the transient formation of interendothelial gaps. 347 CD24 expression has been detected in human BM-MSCs, which are regulated by TGF-β3, 348 allowing them to interact with activated endothelial cells via P-selectin and initiate the tethering and rolling steps of MSCs. 349 Additionally, BM-MSCs express high levels of CXCR4 or CXCR7, 350 , 351 which bind to integrin receptors, such as VLA-4, to activate the integrin-binding process and allow the cells to anchor to endothelial cells, followed by the migration of MSCs through the endothelial cell layer and basement membrane in a process called transmigration. 352 This process is facilitated by the secretion of matrix metalloproteinases (MMPs), which degrade the endothelial basement membrane, allowing BM-MSCs to enter the brain environment. 353 , 354 BM-MSCs can also regulate the integrity of the BBB via the secretion of tissue inhibitor of matrix metalloproteinase-3 (TIMP3), which has been shown to ameliorate the effects of a compromised BBB in traumatic brain injury. 355 The secretion of TIMP3 from MSCs directly blocked vascular endothelial growth factor a (VEGF-a)-induced breakdown of endothelial cell adherent junctions, demonstrating the potential mechanism of BM-MSCs in the regulation of BBB integrity.

The therapeutic applications of BM-MSCs in neurodegenerative conditions have been significantly increased by the demonstration of BM-MSC involvement in axonal and functional remyelination processes. Remyelination is a spontaneous regenerative process occurring in the human CNS to protect oligodendrocytes, neurons, and myelin sheaths from neuronal degenerative diseases. 356 Remyelination is considered a neuroprotective process that limits axonal degeneration by demyelination and neuronal damage. The first mechanism of action of BM-MSCs related to remyelination is the activation of the JAK/STAT3 pathway to regulate dorsal root ganglia development. 357 It was reported that BM-MSCs secrete vascular endothelial growth factor-A (VEGF-A), 358 brain-derived neurotrophic factor (BDNF), interleukin-6, and leukemia inhibitor factor (LIF), which directly function in neurogenesis and neurite growth. 357 VEGF-A is a key regulator of hemangiogenesis during development and bone homeostasis. Postnatally, osteoblast- and MSC-derived VEGF plays a critical role in maintaining and regulating bone homeostasis by stimulating MSC differentiation into osteoblasts and suppressing their adipogenic differentiation. 359 , 360 , 361 To balance osteoblast and adipogenic differentiation, VEGF forms a functional link with the nuclear envelope protein laminin A, which in turn directly regulates the osteoblast and adipocyte transcription factors Runx2 and PPARγ, respectively. 361 , 362 In the brain, VEGF is a potent growth factor mediating angiogenesis, neural migration, and neuroprotection. VEGF-A, secreted from BM-MSCs under in vitro xeno- and serum-free culture conditions, is the most studied member of the VEGF family and is suggested to play a protective role against cognitive impairment, such as in the context of Alzheimer’s disease pathology or stroke. 363 , 364 , 365 Recently, it was reported that the neurotrophic and neuroprotective function of VEGF is mediated through VEGFR2/Flk-1 receptors, which are expressed in the neuroproliferative zones and extend to astroglia and endothelial cells. 366 In animal models of intracerebral hemorrhage and cerebral ischemia, the transfusion of Flk-1-positive BM-MSCs promotes behavioral recovery and anti-inflammatory and angiogenic effects. 367 , 368 Moreover, supplementation with VEGF-A in neuronal disorders enhances intraneural angiogenesis, improves nerve regeneration, and promotes neurotrophic capacities, which in turn increase myelin thickness via the activation of the prosurvival transcription factor nuclear factor-kappa B (NF-kB). This activation, together with the downregulation of Mdm2 and increased expression of the pro-apoptotic transcription factor p53, is considered to be the neuroprotective process associated with an increased VEGF-A level. 369 , 370 , 371 An analysis of microRNA (miRNA) in extracellular vesicles (EVs) secreted from BM-MSCs revealed that BM-MSCs release substantial amounts of miRNA133b, which suppresses the expression of connective tissue growth factor (CTGF) and protects hippocampal neurons from apoptosis and inflammatory injury 372 , 373 , 374 (Fig. 3c ).

In terms of immunoregulatory functions, the administration of human BM-MSCs into immunocompetent mice subjected to SCI or brain ischemia showed that BM-MSCs exhibited a short-term neuronal protective function against neurological damage (Fig. 3d ). Further investigation demonstrated the ability of BM-MSCs to directly communicate with host microglia/macrophages and convert them from phenotypic polarization into alternative activated microglia/macrophages (AAMs), which are key players in axonal extension and the reconstruction of neuronal networks. 375 Other studies have also illustrated that the administration of AAMs directly to the injured spinal cord induced axonal regrowth and functional improvement. 376 The mechanism by which BM-MSCs activate the conversion of microglia/macrophages occurs through two representative macrophage-related chemokine axes, CCL2/CCR2 and CCL-5/CCR5, both of which exhibit acute or chronic elevation following brain injury or SCI. 377 The CCL2/CCR2 axis contributed to the enhancement of inflammatory function, and BM-MSC-mediated induction of CCL2 did not alter the total granulocyte number (Fig. 3d ). Although the chemokine-mediated mechanism of BM-MSCs in the activation of AAMs and enhanced axonal regeneration at the damage sites is evident, the direct mechanism by which the communication between BM-MSCs and the target cells results in these phenomena remains unclear, and further investigation is needed.

BM-MSCs also confer the ability to regulate the inflammatory regulation of the immune cells present in the brain by (1) promoting the polarization of macrophages toward the M2 type, (2) suppressing T-lymphocyte activities, (3) stimulating the proliferation and differentiation of regulatory T cells (Tregs), and (4) inhibiting the activation of natural killer (NK) cells. BM-MSCs secrete glial cell line-derived neurotrophic factor (GDNF), a specific growth factor that contributes directly to the transition of the microglial destructive M1 phenotype into the regenerative M2 phenotype during the neuroinflammatory process. 378 A similar result was also found in AT- 379 and UC-MSCs 380 under neuroinflammation-associated conditions, suggesting that AT-, BM-, and UC-MSCs share the same mechanism in promoting macrophage polarization. In terms of T-lymphocyte suppression, compared to MSCs from AT and BM, UC-MSCs show the strongest potential to inhibit the proliferation of T-lymphocytes by promoting cell cycle arrest (G0/G1 phase) and apoptosis. 381 In addition, UC-MSCs have been proven to be more effective in promoting the proliferation of Tregs 382 and inhibiting NK activation. 383 Although MSCs are well-known for their inflammatory regulatory ability, the mechanism is not exclusive to BM-MSCs, especially in neurological disorders. 384

Proposed mechanism of UC-MSCs in the treatment of pulmonary diseases and lung fibrosis

In contrast to AT-MSCs and BM-MSCs, UC-MSCs have lower expression of major histocompatibility complex I (MHC I) and no expression of MHC II, which prevents the complications of immune rejection. 385 Moreover, as UC is considered a waste product after birth, with the option of noninvasive collection, UC-MSCs are easier to obtain and culture than AD- and BM-MSCs. 386 These advantages of UC-MSCs have contributed to their use in the treatment of pulmonary diseases, especially during the rampant COVID-19 pandemic, as “off-the-shelf” products. Numerous pulmonary diseases have been the subject of applications of UC-MSCs, including BPD, COPD, ARDS, and COVID-19-induced ARDS. In BPD, premature infants are born before the alveolarization process, resulting in arrested lung development and alveolar maturation. Upon administration via an IV route, the majority of exogenous UC-MSCs reach the immature lung and directly interact with immune cells to exert their immunomodulatory properties via cell-to-cell interaction mechanisms (Fig. 4a ). UC-MSCs interact with T cells via the PD-L1 ligand, which binds to the PD-1 inhibitory molecule on T cells, resulting in the suppression of CD3+ T-cell proliferation and effector T-cell responses. 387 In addition, UC-MSCs also express CD54 (ICAM-1), which plays a crucial role in the immunomodulatory functions of T cells. 388 Direct contact between UC-MSCs and macrophages via CD54 expression on UC-MSCs promotes the immune regulation of UC-MSCs via the regulation of phagocytosis by monocytes. 389 Moreover, the contact of UC-MSCs with macrophages during proinflammatory responses increases the secretion of TSG-6 by UC-MSCs, which in turn promotes the inhibitory regulation of CD3+ T cells, macrophages, and monocytes by MSCs. 390 Recently, upregulation of SDF-1 was described in neonatal lung injury, especially in layers of the respiratory epithelium. 391 SDF-1 has been shown to participate in the migration and initiation of the homing process of MSCs via the CXCR4 receptors on their surface. 392 It was reported that UC-MSCs express low levels of CXCR4, allowing them to induce SDF-1-associated migration processes via the Akt, ERK, and p38 signal transduction pathways. 393 Hence, in BPD, the upregulation of SDF-1 together with the homing ability of UC-MSCs strongly supports the therapeutic effects of UC-MSCs in the treatment of BPD. Furthermore, UC-MSCs have the ability to communicate with immune cells via cell-to-cell contact to reduce proinflammatory responses and the production of proinflammatory cytokines (such as TGF-β, INF-γ, macrophage MIF, and TNF-α). The modulation of the human innate immune system by UC-MSCs is mediated by cell–cell interactions via CD54-LFA-1 that switch macrophage polarization processes, promoting the proliferation of M2 macrophages, which in turn reduce inflammatory responses in the immature lung. 394 Moreover, UC-MSCs also have the ability to produce VEGF and hepatocyte growth factors (HGFs), promoting angiogenesis and enhancing lung maturation. 395

figure 4

Adipose tissue-derived mesenchymal stem cells (AT-MSCs) and the nature of their tissue of origin support their use in therapeutic applications. a Adipose tissue is considered an endocrine organ, supporting and regulating various functions, including appetite regulation, immune regulation, sex hormone and glucocorticoid metabolism, energy production, the orchestration of reproduction, the control of vascularization, and blood flow, the regulation of coagulation, and angiogenesis and skin regeneration. b In terms of metabolic disorders, such as type 2 diabetes mellitus (T2DM), as adipose tissue is directly involved in the metabolism of glucose and lipids and the regulation of appetite, the detrimental effects of T2DM also alter the functions of AT-MSCs, which in turn, hampers their therapeutic effects. Hence, the use of autologous AT-MSCs is not recommended for the treatment of metabolic disorders, including T2DM, suggesting that allogeneic AT-MSCs from healthy donors could be a better alternative approach. c AT-MSCs are suitable for the treatment of reproductive disorders due to their unique ability to mobilize and home to the thecal layer of the injured ovary, enhance the regeneration and maturation of thecal cells, increase the structure and function of damaged ovaries via exosome-activated SMAD, decrease oxidative stress and autophagy, and increase the proliferation of granulosa cells via PI3K/AKT pathways. These functions are regulated specifically by growth hormones produced by AT-MSCs in response to the surrounding environment, including HGF, TGF-β, IGF-1, and EGF. d AT-MSCs are also good candidates for skin healing and regeneration as their growth factors strongly support neovascularization and angiogenesis by reducing PLL4, increase anti-apoptosis via the activation of PI3K/AKT pathways, regulate inflammation by downregulating NADPH oxidase isoform 1, and increase immunoregulation through the inhibition of NF-κB activation. The figure was created with BioRender.com

COPD is characterized by an increase in hyperinflammatory reactions in the lung, compromising lung function and increasing the development of lung fibrosis. The mechanism by which UC-MSCs contribute to the response to COPD is inflammatory regulation (Fig. 4b ). The administration of UC-MSCs prevented the infiltration of inflammatory cells in peribronchiolar, perivascular, and alveolar septa and switched macrophage polarization to M2. 396 A significant reduction in proinflammatory cytokines, including IL-1β, TNF-α, and IL-8, was also observed following UC-MSC administration. 224 MSCs, including UC-MSCs, have been reported to trigger the production of secretory leukocyte protease inhibitors in epithelial cells through the secretion of HGF and epidermal growth factor (EGF), which is believed to have beneficial effects on COPD. 397 , 398 In addition to their inflammatory regulation ability, UC-MSCs exhibit antimicrobial effects through the inhibition of bacterial growth and the alleviation of antibiotic resistance during Pseudomonas aeruginosa infection. 399 The combination of the regulation of the host immune response and the antimicrobial effects of UC-MSCs may be relevant for the prevention and treatment of COPD exacerbations, as inflammation and bacterial infections are important risk factors that significantly contribute to the morbidity and mortality of patients with COPD. In terms of regenerative functions, UC-MSCs were reported to be able to differentiate into type 2 alveolar epithelial cells in vitro and alleviate the development of pulmonary fibrosis via β-catenin-regulated cell apoptosis. 400 Furthermore, UC-MSCs enhanced alveolar epithelial cell migration and proliferation by increasing matrix metalloproteinase-2 levels and reduced their endogenous inhibitors, tissue inhibitors of matrix metalloproteinases, providing a potential mechanism underlying their anti-pulmonary-fibrosis effects. 401 , 402

In ARDS, especially that associated with COVID-19, the proinflammatory state is initiated by increases in plasma concentrations of proinflammatory cytokines, such as IL-1 beta, IL-7, IL-8, IL-9, IL-10, bFGF, granulocyte colony-stimulating factor (G-CSF), GM-CSF, IFN-γ, and TNF-α. The significant increases in the concentrations of these cytokines in patient plasma suggest the development of a cytokine storm, which is a leading cause of COVID-induced mortality. In addition to the immunomodulatory functions regulated via cell-to-cell interactions between UC-MSCs and immune cells, such as macrophages, monocytes, and T cells, UC-MSCs exert their functions via paracrine effects through the secretion of growth factors, cytokines, and exosomes (Fig. 4c ). The most relevant immunomodulatory function of UC-MSCs is considered to be their inhibition of effector T cells via the induction of T-cell apoptosis and cell cycle arrest by the production of indoleamine 2,3- dioxygenase (IDO), prostaglandin E2 (PGE-2), and TGF-β. Elevated levels of PGE-2 in patients with COVID-19 are reported to be a crucial factor in the initiation of inflammatory regulation by UC-MSCs post administration and prevent the development of cytokine storms by direct inhibition of T- and B lymphocytes. 403 UC-MSCs exert these inhibitory activities through a PGE-2-dependent mechanism. 404 It was reported that UC-MSCs confer the ability to secrete tolerogenic mediators, including TGF-β1, PGE-2, nitric oxide (NO), and TNF-α, which are directly involved in their immunoregulatory mechanism. The secretion of NO from UC-MSCs is reported to be associated with the desensitization of T cells via the IFN-inducible nitric oxide synthase (iNOS) pathways and to stimulate the migration of T cells in close proximity to MSCs that subsequently suppress T-cell sensitivities via NO. 405 Lung infection with viruses usually leads to impairments in alveolar fluid clearance and protein permeability. The administration of UC-MSCs enhances alveolar protection and restores fluid clearance in patients with COVID-19. UC-MSCs secrete growth factors associated with angiogenesis and the regeneration of pulmonary blood vessels and micronetworks, including angiotensin-1, VEGF, and HGF, which also reduce oxidative stress and prevent fibrosis formation in the lungs. These trophic factors have been identified as key players in the modulation of the microenvironment and promote pulmonary repair. Additionally, UC-MSCs are more effective than BM-MSCs in the restoration of impaired alveolar fluid clearance and the permeability of airways in vitro, supporting the use of UC-MSCs in the treatment of patients with pulmonary pneumonia. 406 In the context of pulmonary regeneration, UC-MSCs were shown to inhibit apoptosis and fibrosis in pulmonary tissue by activating the PI3K/AKT/mTOR pathways via the secretion of HGF, which also acts as an inhibitory stimulus that blocks alveolar epithelial-to-mesenchymal transition. 407 , 408 Moreover, UC-MSCs can reverse the process of fibrosis via enhanced expression of macrophage matrix-metallopeptidase-9 for collagen degradation and facilitate alveolar regeneration via Toll-like receptor-4 signaling pathways. 409 UC-MSCs were shown to communicate with CD4+ T cells through HGF induction not only to inhibit their differentiation into Th17 cells, reducing the secretion of IL-17 and IL-22 but also to switch their differentiation into regulatory T cells. 410 , 411 In addition, UC-MSCs conferred the ability to facilitate the number of M2 macrophages and reduce M1 cells via the control of the macrophage polarization process. 412

There are several potential mechanisms of UC-MSCs in the treatment of patients with pulmonary diseases and pneumonia, including the regulation of immune cell function, immunomodulation, the enhancement of alveolar fluid clearance and protein permeability, the modulation of endoplasmic reticulum stress, and the attenuation of pulmonary fibrosis. Hence, based on these discussions, UC-MSCs are recommended as suitable candidates for the treatment of pulmonary disease both in pediatric and adult patients.

Proposed mechanism of AT-MSCs in the treatment of endocrinological diseases, reproductive disorders, and skin burns

Human AT was first viewed as a passive reservoir for energy storage and later as a major site for sex hormone metabolism, the production of endocrine factors (such as adipsin and leptin), and a secretion source of bioactive peptides known as adipokines. 413 It is now clear that AT functions as a complex and highly active metabolic and endocrine organ, orchestrating numerous different biological features 414 (Fig. 5a ). In addition to adipocytes, AT contains hematopoietic-derived progenitor cells, connective tissue, nerve tissue, stromal cells, endothelial cells, MSCs, and pericytes. AT-MSCs and pericytes mobilize from their perivascular locations to aid in healing and tissue regeneration throughout the body. As AT is involved directly in energy storage and metabolism, AT-MSCs are also mediated and regulated by growth factors related to these pathways. In particular, interleukin-6 (IL-6), IL-33, and leptin regulate the maintenance of metabolic activities by increasing insulin sensitivity and preserving homeostasis related to AT. Nevertheless, in the development of obesity and diabetes, omental and subcutaneous AT maintains a low-grade state of inflammation, resulting in the impairment of glucose metabolism and potentially contributing to the development of insulin resistance. 415 In normal AT, direct regulation of Pre-B-cell leukemia homeobox (Pbx)-regulating protein-1 (PREP1) by leptin and thyroid growth factor-beta 1 (TGF-β1) in AT-MSCs and mature adipocytes is involved in the protective function and maintenance of AT homeostasis. However, under diabetic conditions, the balance between the expression of leptin and the secretion of TGF-β1 is compromised, resulting in the malfunction of AT-MSC metabolic activity and the proliferation, differentiation, and maturation of adipocytes. Therefore, the use of autologous AT-MSCs in the treatment of diabetic conditions is not a suitable option, as the functions of AT-MSCs are directly altered by diabetic conditions, which reduces their effectiveness in cell-based therapy (Fig. 5b ).

figure 5

Umbilical cord-derived mesenchymal stem cells (UC-MSCs) are good candidates for the treatment of pulmonary diseases. a Lung immaturity and fibrosis are the major problems of patients with bronchopulmonary dysplasia and lead to increased levels of SDF-1, the development of fibrosis, the induction of the inflammatory response, and the impairment of alveolarization. UC-MSCs are attracted to the damaged lung via the chemoattractant SDF-1, which is constantly released from the immature lung via SDF-1 and CXCR4 communication. Moreover, UC-MSCs reduce the level of proinflammatory cytokines (TGF-β, INF-γ, macrophage MIF, and TNF-α) via a cell-to-cell contact mechanism. The ability of UC-MSCs to produce and secrete VEGF also involves in the regeneration of the immature lung through enhanced angiogenesis. b Upon an exacerbation of chronic obstructive pulmonary disease (COPD), UC-MSCs respond to the surrounding stimuli by reducing IL-8 and TNF-α levels, resulting in the inhibition of the inflammatory response but an increase in the secretion of growth factors participating in the protection of alveoli, fluid clearance and reduced oxidative stress and lung fibrosis, including HGF, TGF-β, IGF-1, and exosomes. c In a similar manner, UC-MSCs prevent the formation of cytokine storms in coronavirus disease 2019 (COVID-19) by inhibiting CD34+ T-cell differentiation into Th17 cells and enhancing the number of regulatory T cells. Moreover, UC-MSCs also have antibacterial activity by secreting LL-3717 and lipocalin. Figure was created with BioRender.com

Preclinical studies and clinical trials have revealed the therapeutic effects of MSCs, in general, and AT-MSCs, in particular, in the management of POF, with relatively high efficacy and enhanced regeneration of the ovaries. Understanding the molecular and cellular mechanisms underlying these effects is the first step in the development of suitable MSC-based therapies for POF. One of the mechanisms by which MSCs exert their therapeutic effects is their ability to migrate to sites of injury, a process known as “homing”. Studies have shown that MSCs from different sources have the ability to migrate to different compartments of the injured ovary. For example, BM-MSCs administered through IV routes migrated mostly to the ovarian hilum and medulla, 416 whereas a significant number of UC-MSCs were found in the medulla. 417 Interestingly, AT-MSCs were found to be engrafted in the theca layers of the ovary but not in the follicles, where they acted as supportive cells to promote follicular growth and the regeneration of thecal layers. 418 The structure and function of the thecal layer have a great impact on fertility, which has been reviewed elsewhere. 419 In brief, the thecal layer consists of two distinct parts, the theca interna, which contains endocrine cells, and the theca externa, which is an outer fibrous layer. The thecal layer contains not only endocrine-derived cells but also vascular- and immune-derived cells, whose functions are to maintain the structural integrity of the follicles, transport nutrients to the inner compartment of the ovary and produce key reproductive hormones such as androgens (testosterone and dihydrotestosterone) and growth factors (morphogenic proteins, e.g., BMPs and TGF-β). 420 As AT-MSCs originate from an endocrine organ, their ability to sense signals and migrate to the thecal layer is anticipated. Additionally, secretome analysis of AT-MSCs showed a wide range of growth factors, including HGF, TBG-β, VEGF, insulin-like growth factor-1 (IGF-1), and EGF, 421 that are directly involved in the restoration of the structure and function of damaged ovaries by stimulating cell proliferation and reducing the aging process of oocytes via the activation of the SIRT1/FOXO1 pathway, a key regulator of vascular endothelial homeostasis. 422 , 423 In POF pathology, autophagy and its correlated oxidative stress contribute to the development of POF throughout a patient’s life. Recently, AT-MSCs were shown to be able to improve the structure and function of mouse ovaries by reducing oxidative stress and inflammation, providing essential data supporting the mechanism of AT-MSCs in the treatment of POF. 424 Several studies have illustrated that AT-MSCs secrete biologically active EVs that regulate the proliferation of ovarian granulosa cells via the PI3K/AKT pathway, resulting in the enhancement of ovarian function. 425 Direct regulation of ovarian cell proliferation modulates the state of these cells, which in turn restores the ovarian reserve. 426 Other mechanisms supporting the effectiveness of MSCs have been carefully reviewed, confirming the therapeutic potential of MSCs derived from different sources 426 (Fig. 5c ).

In the last decade, the number of clinical trials using AT-MSCs in the treatment of chronic skin wounds and skin regeneration has exponentially increased, with data supporting the enhancement of the skin healing processes, the reduction of scar formation, and improvements in skin structure and quality. Several mechanisms are directly linked to the origin of AT-MSCs, including differentiation ability, neovascularization, anti-apoptosis, and immunological regulation. AT is a connective and supportive tissue positioned just beneath the skin layers. AT-MSCs have a strong ability to differentiate into adipocytes, endothelial cells, 427 epithelial cells 428 and muscle cells. 429 The adipogenic differentiation of AT-MSCs is one of the three mesoderm lineages that defines MSC features, and AT-MSCs are likely to be the best MSC type harboring this ability compared to BM- and UC-MSCs. Recent reports detailed that AT-MSCs accelerated diabetic wound tissue closure through the recruitment and differentiation of endothelial cell progenitor cells into endothelial cells mediated by the VEGF-PLCγ-ERK1/ERK2 pathway. 430 Upon injury, the skin must be healed as quickly as possible to prevent inflammation and excessive blood loss. The reparation process occurs through distinct overlapping phases and involves various cell types and processes, including endothelial cells, keratinocyte proliferation, stem cell differentiation, and the restoration of skin homeostasis. 431 Hence, the differentiation ability of AT-MSCs plays a critical role in their therapeutic effect on skin wound regeneration and healing processes. AT-MSCs accelerate wound healing via the production of exosomes that serve as paracrine factors. It was reported that AT-MSCs responded to skin wound injury stimuli by increasing their expression of the lncRNA H19 exosome, which upregulated SOX9 expression via miR-19b, resulting in the acceleration of human skin fibroblast proliferation, migration, and invasion. 432 In addition, the engraftment of AT-MSCs supported wound bed blood flow and epithelialization processes. 433 Anti-apoptosis plays a critical role in AT-MSC-based therapy, as without a microvascular supply network established within 4 days post injury, adipocytes undergo apoptosis and degenerate. Exogenous sources of AT-MSCs mediate anti-apoptosis via IGF-1 and exosome secretion by triggering the activation of PI3K signaling pathways. 434 Another mechanism supporting the therapeutic potential of AT-MSCs is their anti-inflammatory function, which results in the reduction of proinflammatory factors, such as tumor necrosis factor (TNF) and interferon-γ (IFN-γ), and increases the production of the anti-inflammatory factors IL-10 and IL-4. Exosomes from AT-MSCs in response to a wound environment were found to contain high levels of Nrf2, which downregulated wound NADPH oxidase isoform 1 (NOX1), NADPH oxidase isoform 4 (NOX4), IL-1β, IL-6, and TNF-α expression. The anti-inflammatory functions of AT-MSCs are also regulated by their immunomodulatory ability, partially through the inhibition of NF-κB activation in T cells via the PD-L1/PD-1 and Gal-9/TIM-3 pathways, providing a novel target for the acceleration of wound healing 435 (Fig. 5d ).

Therefore, as an endocrine organ in the human body, AT and its derivative stem cells, including AT-MSCs, have shown great potential in the treatment of reproductive disorders and skin diseases. Their potential is supported by mechanisms that are directly related to the nature of AT-MSCs in the maintenance of tissue homeostasis, angiogenesis, anti-apoptosis, and the regulation of inflammatory responses.

The current challenges for MSC-based therapies

Over the past decades, MSC-based research and therapy have made tremendous advancements due to their advantages, including immune evasion, diverse tissue sources for harvesting, ease of isolation, rapid expansion, and cryopreservation as “off-the-shelf” products. However, several important challenges have to be addressed to further enhance the safety profile and efficacy of MSC-based therapy. In our opinion, the most important challenge of MSC-based therapy is the fate of these cells post administration, especially the long-term survival of allogeneic cells in the treatment of certain diseases. Although reported data confirm that the majority of MSCs are trapped in the lung and rapidly removed from the circulation, caution has been raised related to the occurrence of embolism events post infusion, which was proven to be related to MSC-induced innate immune attack (called instant blood-mediated inflammatory reaction). 436 Another related challenge is the homing ability of infused cells, as successful homing at targeted tissue might result in long-term benefits to patients. Other concerns related to MSC-based therapy are the number of dead cells infused into the patients. An interesting study reported that dead MSCs alone still exerted the same immunomodulatory property as live MSCs by releasing phosphatidylserine. 437 This is an interesting observation, as there is always a certain number of dead cells present in the cell-based product, and concerns are always raised related to their effects on the patient’s health. Finally, the hypothesis presented in this review is also a great challenge of the field, which has been proposed for future studies to answer the question: “What is the impact of MSC sources on their downstream application?”. Tables 5 and 6 illustrate the comparative studies that were conducted in preclinical and clinical settings to address the MSC source challenge. Other challenges of MSC-based therapies have been discussed in several reviews and systematic studies, 135 , 185 , 438 , 439 which are highly recommended.

Limitations of the current hypothesis

The proposed hypothesis presented in this review was made based on (1) the calculated number of recovered patients from published clinical trials; (2) the empirical experience of the authors in the treatment of brain-related diseases, 440 pulmonary disorders, 215 and endocrinological conditions; 271 , 441 and (3) the proposed mechanisms by which each type of MSC exhibits its best potential for downstream applications. The authors understand that the approach that we used has a certain level of research bias, as a comprehensive meta-analysis is needed to first confirm the correlation between the origins of MSCs and their downstream clinical outcomes before a complete hypothesis can be made. However, to date, a limited number of clinical trials have been conducted to directly compare the efficacy of MSCs from different sources in treating the same disease, which in turn dampened our analysis to prove this hypothesis. In addition, MSC-based therapy is still in its early stages, as controversy and arguments are still present in the field, including (1) the name of MSCs (medicinal signaling cells vs. MSCs or mesenchymal stromal cells), 442 , 443 (2) the existence of “magic cells” (one cell type for the treatment of all diseases), 444 , 445 (3) the conflicting results from large-scale clinical trials, 135 and (4) the dangerous issues of unauthorized, unproven stem cell therapies and clinics. 446 , 447 Therefore, our hypothesis is proposed at this time to encourage active researchers and clinicians to either prove or disprove it so that future research can strengthen the uses of MSC-based therapies with solid mechanistic study results and clarify results for “one cell type for the treatment of all diseases”.

Another limitation is the knowledge coverage in the field of MSC-based regenerative medicine, as discussed in this study. First, the abovementioned diseases were narrowed to four major disease categories for which MSC-based therapy is widely applied, including neuronal, pulmonary, cardiovascular, and endocrinological conditions. In fact, other diseases also receive great benefits from MSC therapy, including liver cirrhosis, 448 bone regeneration, 360 plastic surgery, 449 autoimmune disease, 450 etc., which are not fully discussed in this review and included in our hypothesis. Recently, the secretome profile of MSCs and its potential application in clinical settings have emerged as a new player in the field, with a recently published comprehensive review including MSC-derived exosomes. 451 , 452 To date, the therapeutic potential of MSCs is believed to be strongly influenced by their secretomes, including growth factors, cytokines, chemokines, and exosomes. 453 However, this body of knowledge is also not fully included in our discussion, as this review focuses on the function and potency of MSCs as a whole with considerations derived from published clinical data. Therefore, the authors believe in and support the future applications of the secreted components derived from MSCs, including exosomes, in the treatment of human diseases. In fact, this potential approach could elevate the uses of MSCs to the next level, where the sources of MSCs could be neglected with advancements in the development of protocols that allow strict control of the secretome profiles of MSCs under specific conditions. 454 , 455 , 456 Finally, strategies that could potentially enhance the therapeutic outcomes of MSC-based therapy, such as the “priming” process, are not discussed in this review. The idea of “priming” MSCs is based on the nature of MSCs, which is similar to the immune cells, 457 that MSCs have proven to be able to “remember” the stimulus from the surrounding environment. 458 , 459 Thus, activating or priming MSCs using certain conditions, such as hypoxia, matrix mechanics, 3D environment, hormones, or inflammatory cytokines, could trigger the memory mechanism of the MSCs in vitro so that these cells are ready to function towards specific therapeutic activities without the need for in vivo activation. 3 , 460

From a cellular and molecular perspective and from our own experience in a clinical trial setting, AD-, BM- and UC-MSCs exhibit different functional activities and treatment effectiveness across a wide range of human diseases. In this paper, we have provided up-to-date data from the most recently published clinical trials conducted in neuronal diseases, endocrine and reproductive disorders, skin regeneration, pulmonary dysplasia, and cardiovascular diseases. The implications of the results and discussions presented in this review and in a very large body of comprehensive and excellent reviews as well as systematic analyses in the literature provide a different aspect and perspective on the use of MSCs from different sources in the treatment of human diseases. We strongly believe that the field of regenerative medicine and MSC-based therapy will benefit from active discussion, which in turn will significantly advance our knowledge of MSCs. Based on the proposed mechanisms presented in this review, we suggest several key mechanistic issues and questions that need to be addressed in the future:

The confirmation and demonstration of the mechanism of action prove that tissue origin plays a significant role in the downstream applications of the originated MSCs.

Is it required that MSCs derived from particular cell sources need to have certain functionalities that are unique to or superior in the original tissue sources?

As mechanisms may rely on the secretion of factors from MSCs, it is important to identify the specific stimuli from the wound environments to understand how MSCs from different sources can exhibit similar functions in the same disease and whether or not MSCs derived from a particular source have stronger effects than their counterparts derived from other tissue sources.

Should we create “universal” MSCs that could be functionally equal in the treatment of all diseases regardless of their origin by modeling their genetic materials?

Can new sources of MSCs from either perinatal or adult tissues better stimulate the innate mechanisms of specific cell types in our body, providing a better tool for MSC-based treatment?

A potential ‘priming’ protocol that allows priming, activating, and switching the potency of MSCs from one source to another with a more appropriate clinical phenotype to treat certain diseases. This idea is potentially relevant to our suggestion that each MSC type could be more beneficial in downstream applications, and the development of such a “priming” protocol would allow us to expand the bioavailability of specific MSC types.

From our clinical perspective, the underlying proposal in our review is to no longer use MSCs for applications while disregarding their sources but rather to match the MSC tissue source to the application, shifting from one cell type for the treatment of all diseases to cell source-specific disease treatments. Whether the application of MSCs from different sources still shows their effectiveness to a certain extent in the treatment of diseases or not, the transplantation of MSCs derived from different sources for each particular disease needs to be further investigated, and protocols need to be established via multicentre, randomized, placebo-controlled phase II and III clinical trials (Fig. 6 ).

figure 6

The tissue sources of mesenchymal stem cells (MSCs) contribute greatly to their therapeutic potential, as all MSC types share safety profiles and overlapping efficacy. Although a large body of data and their review and systematic analysis indicated the shared safety and potential efficacy of MSCs derived from different tissue sources, targeted therapies considering MSC origin as an important factor are imperative to enhance the downstream therapeutic effects of MSCs. We suggest that bone marrow-derived MSCs (BM-MSCs) are good candidates for the treatment of brain and spinal cord injury, adipose tissue-derived MSCs (AT-MSCs) are suitable for the treatment of reproductive disorders and skin regeneration, and umbilical cord-derived MSCs (UC-MSCs) could be alternatives for the treatment of pulmonary diseases and acute respiratory distress syndrome (ARDS). Figure was created with BioRender.com

Data availability

All data generated or analyzed in this study are included in this published article.

Mellman, I., Coukos, G. & Dranoff, G. Cancer immunotherapy comes of age. Nature 480 , 480–489 (2011).

Article   PubMed   PubMed Central   CAS   Google Scholar  

Ancans, J. Cell therapy medicinal product regulatory framework in Europe and its application for MSC-based therapy development. Front. Immunol. 3 , 253 (2012).

Article   PubMed   PubMed Central   Google Scholar  

Yin, J. Q., Zhu, J. & Ankrum, J. A. Manufacturing of primed mesenchymal stromal cells for therapy. Nat. Biomed. Eng. 3 , 90–104 (2019).

Article   PubMed   CAS   Google Scholar  

O’Brien, T. & Barry, F. P. Stem cell therapy and regenerative medicine. Mayo Clin. Proc. 84 , 859–861 (2009).

Mousaei Ghasroldasht, M., Seok, J., Park, H. S., Liakath Ali, F. B. & Al-Hendy, A. Stem cell therapy: from idea to clinical practice. Int. J. Mol. Sci . 23 , 2850 (2022).

Kuriyan, A. E. et al. Vision loss after intravitreal injection of autologous “stem cells” for AMD. N. Engl. J. Med. 376 , 1047–1053 (2017).

Biehl, J. K. & Russell, B. Introduction to stem cell therapy. J. Cardiovasc. Nurs. 24 , 98–103 (2009). quiz 104-105.

Srijaya, T. C., Ramasamy, T. S. & Kasim, N. H. Advancing stem cell therapy from bench to bedside: lessons from drug therapies. J. Transl. Med. 12 , 243 (2014).

Ramalho-Santos, M. & Willenbring, H. On the origin of the term “stem cell”. Cell Stem Cell 1 , 35–38 (2007).

Konstantinov, I. E. In search of Alexander A. Maximow: the man behind the unitarian theory of hematopoiesis. Perspect. Biol. Med. 43 , 269–276 (2000).

Droscher, A. Images of cell trees, cell lines, and cell fates: the legacy of Ernst Haeckel and August Weismann in stem cell research. Hist. Philos. Life Sci. 36 , 157–186 (2014).

Article   PubMed   Google Scholar  

Jansen, J. The first successful allogeneic bone-marrow transplant: Georges Mathe. Transfus. Med. Rev. 19 , 246–248 (2005).

Blume, K. G. & Weissman, I. L. E. Donnall Thomas (1920-2012). Proc. Natl Acad. Sci. USA 109 , 20777–20778 (2012).

Cheng, M. Hartmann Stahelin (1925-2011) and the contested history of cyclosporin A. Clin. Transpl. 27 , 326–329 (2013).

Article   CAS   Google Scholar  

Thomas, E. D. et al. Aplastic anaemia treated by marrow transplantation. Lancet 1 , 284–289 (1972).

Friedenstein, A. J., Chailakhyan, R. K. & Gerasimov, U. V. Bone marrow osteogenic stem cells: in vitro cultivation and transplantation in diffusion chambers. Cell Tissue Kinet. 20 , 263–272 (1987).

PubMed   CAS   Google Scholar  

Friedenstein, A. J., Chailakhjan, R. K. & Lalykina, K. S. The development of fibroblast colonies in monolayer cultures of guinea-pig bone marrow and spleen cells. Cell Tissue Kinet. 3 , 393–403 (1970).

Caplan, A. I. Mesenchymal stem cells. J. Orthop. Res. 9 , 641–650 (1991).

Bolli, R., Tang, X. L., Guo, Y. & Li, Q. After the storm: an objective appraisal of the efficacy of c-kit+ cardiac progenitor cells in preclinical models of heart disease. Can. J. Physiol. Pharm. 99 , 129–139 (2021).

Liu, C., Han, D., Liang, P., Li, Y. & Cao, F. The current dilemma and breakthrough of stem cell therapy in ischemic heart disease. Front. Cell Dev. Biol. 9 , 636136 (2021).

Zhang, J. et al. Basic and translational research in cardiac repair and regeneration: JACC state-of-the-art review. J. Am. Coll. Cardiol. 78 , 2092–2105 (2021).

Gyongyosi, M., Wojakowski, W., Navarese, E. P., Moye, L. A. & Investigators, A. Meta-analyses of human cell-based cardiac regeneration therapies: controversies in meta-analyses results on cardiac cell-based regenerative studies. Circ. Res. 118 , 1254–1263 (2016).

Okamoto, R., Matsumoto, T. & Watanabe, M. Regeneration of the intestinal epithelia: regulation of bone marrow-derived epithelial cell differentiation towards secretory lineage cells. Hum. Cell 19 , 71–75 (2006).

Gehart, H. & Clevers, H. Tales from the crypt: new insights into intestinal stem cells. Nat. Rev. Gastroenterol. Hepatol. 16 , 19–34 (2019).

Santos, A. J. M., Lo, Y. H., Mah, A. T. & Kuo, C. J. The intestinal stem cell niche: homeostasis and adaptations. Trends Cell Biol. 28 , 1062–1078 (2018).

Roda, G. et al. Crohn’s disease. Nat. Rev. Dis. Prim. 6 , 22 (2020).

Kobayashi, T. et al. Ulcerative colitis. Nat. Rev. Dis. Prim. 6 , 74 (2020).

Gratwohl, A. et al. Autologous hematopoietic stem cell transplantation for autoimmune diseases. Bone Marrow Transpl. 35 , 869–879 (2005).

Kashyap, A. & Forman, S. J. Autologous bone marrow transplantation for non-Hodgkin’s lymphoma resulting in long-term remission of coincidental Crohn’s disease. Br. J. Haematol. 103 , 651–652 (1998).

Hurley, J. M., Lee, S. G., Andrews, R. E. Jr., Klowden, M. J. & Bulla, L. A. Jr. Separation of the cytolytic and mosquitocidal proteins of Bacillus thuringiensis subsp. israelensis. Biochem Biophys. Res. Commun. 126 , 961–965 (1985).

Oyama, Y. et al. Autologous hematopoietic stem cell transplantation in patients with refractory Crohn’s disease. Gastroenterology 128 , 552–563 (2005).

Burt, R. K. et al. Autologous nonmyeloablative hematopoietic stem cell transplantation in patients with severe anti-TNF refractory Crohn disease: long-term follow-up. Blood 116 , 6123–6132 (2010).

Hasselblatt, P. et al. Remission of refractory Crohn’s disease by high-dose cyclophosphamide and autologous peripheral blood stem cell transplantation. Aliment Pharm. Ther. 36 , 725–735 (2012).

Hawkey, C. J. et al. Autologous hematopoetic stem cell transplantation for refractory Crohn disease: a randomized clinical trial. J. Am. Med. Assoc. 314 , 2524–2534 (2015).

Lindsay, J. O. et al. Autologous stem-cell transplantation in treatment-refractory Crohn’s disease: an analysis of pooled data from the ASTIC trial. Lancet Gastroenterol. Hepatol. 2 , 399–406 (2017).

Wang, R. et al. Stem cell therapy for Crohn’s disease: systematic review and meta-analysis of preclinical and clinical studies. Stem Cell Res Ther. 12 , 463 (2021).

Hawkey, C. J. Hematopoietic stem cell transplantation in Crohn’s disease: state-of-the-art treatment. Dig. Dis. 35 , 107–114 (2017).

Si-Tayeb, K., Lemaigre, F. P. & Duncan, S. A. Organogenesis and development of the liver. Dev. Cell 18 , 175–189 (2010).

Xue, R. et al. Clinical performance of stem cell therapy in patients with acute-on-chronic liver failure: a systematic review and meta-analysis. J. Transl. Med. 16 , 126 (2018).

Shi, M. et al. Human mesenchymal stem cell transfusion is safe and improves liver function in acute-on-chronic liver failure patients. Stem Cells Transl. Med. 1 , 725–731 (2012).

Liu, Y., Dong, Y., Wu, X., Xu, X. & Niu, J. The assessment of mesenchymal stem cells therapy in acute on chronic liver failure and chronic liver disease: a systematic review and meta-analysis of randomized controlled clinical trials. Stem Cell Res. Ther. 13 , 204 (2022).

Lin, B. L. et al. Allogeneic bone marrow-derived mesenchymal stromal cells for hepatitis B virus-related acute-on-chronic liver failure: a randomized controlled trial. Hepatology 66 , 209–219 (2017).

Gordon, M. Y. et al. Characterization and clinical application of human CD34+ stem/progenitor cell populations mobilized into the blood by granulocyte colony-stimulating factor. Stem Cells 24 , 1822–1830 (2006).

Arroyo, V. et al. Acute-on-chronic liver failure in cirrhosis. Nat. Rev. Dis. Prim. 2 , 16041 (2016).

Zhang, Z. et al. Human umbilical cord mesenchymal stem cells improve liver function and ascites in decompensated liver cirrhosis patients. J. Gastroenterol. Hepatol. 27 (Suppl 2), 112–120 (2012).

El-Ansary, M. et al. Phase II trial: undifferentiated versus differentiated autologous mesenchymal stem cells transplantation in Egyptian patients with HCV induced liver cirrhosis. Stem Cell Rev. Rep. 8 , 972–981 (2012).

Xu, L. et al. Randomized trial of autologous bone marrow mesenchymal stem cells transplantation for hepatitis B virus cirrhosis: regulation of Treg/Th17 cells. J. Gastroenterol. Hepatol. 29 , 1620–1628 (2014).

Suk, K. T. et al. Transplantation with autologous bone marrow-derived mesenchymal stem cells for alcoholic cirrhosis: Phase 2 trial. Hepatology 64 , 2185–2197 (2016).

Fang, X. et al. A study about immunomodulatory effect and efficacy and prognosis of human umbilical cord mesenchymal stem cells in patients with chronic hepatitis B-induced decompensated liver cirrhosis. J. Gastroenterol. Hepatol. 33 , 774–780 (2018).

Mohamadnejad, M. et al. Randomized placebo-controlled trial of mesenchymal stem cell transplantation in decompensated cirrhosis. Liver Int. 33 , 1490–1496 (2013).

Nguyen, T. L. et al. Autologous bone marrow mononuclear cell infusion for liver cirrhosis after the Kasai operation in children with biliary atresia. Stem Cell Res. Ther. 13 , 108 (2022).

Bai, Y. Q. et al. Outcomes of autologous bone marrow mononuclear cell transplantation in decompensated liver cirrhosis. World J. Gastroenterol. 20 , 8660–8666 (2014).

Guo, C. et al. Long-term outcomes of autologous peripheral blood stem cell transplantation in patients with cirrhosis. Clin. Gastroenterol. Hepatol. 17 , 1175–1182 e1172 (2019).

Newsome, P. N. et al. Granulocyte colony-stimulating factor and autologous CD133-positive stem-cell therapy in liver cirrhosis (REALISTIC): an open-label, randomised, controlled phase 2 trial. Lancet Gastroenterol. Hepatol. 3 , 25–36 (2018).

Spahr, L. et al. Autologous bone marrow mononuclear cell transplantation in patients with decompensated alcoholic liver disease: a randomized controlled trial. PLoS ONE 8 , e53719 (2013).

Maurice, J. & Manousou, P. Non-alcoholic fatty liver disease. Clin. Med. 18 , 245–250 (2018).

Article   Google Scholar  

Huang, T. D., Behary, J. & Zekry, A. Non-alcoholic fatty liver disease: a review of epidemiology, risk factors, diagnosis and management. Intern. Med. J. 50 , 1038–1047 (2020).

Sakai, Y. et al. Clinical trial of autologous adipose tissue-derived regenerative (stem) cells therapy for exploration of its safety and efficacy. Regen. Ther. 18 , 97–101 (2021).

Mieli-Vergani, G. et al. Autoimmune hepatitis. Nat. Rev. Dis. Primers 4 , 18018 (2018).

Calore, E. et al. Haploidentical stem cell transplantation cures autoimmune hepatitis and cerebrovascular disease in a patient with sickle cell disease. Bone Marrow Transpl. 53 , 644–646 (2018).

Vento, S., Cainelli, F., Renzini, C., Ghironzi, G. & Concia, E. Resolution of autoimmune hepatitis after bone-marrow transplantation. Lancet 348 , 544–545 (1996).

Terziroli Beretta-Piccoli, B., Mieli-Vergani, G. & Vergani, D. Autoimmmune hepatitis. Cell Mol. Immunol. 19 , 158–176 (2022).

Wang, L. et al. Pilot study of umbilical cord-derived mesenchymal stem cell transfusion in patients with primary biliary cirrhosis. J. Gastroenterol. Hepatol. 28 (Suppl 1), 85–92 (2013).

Wang, L. et al. Allogeneic bone marrow mesenchymal stem cell transplantation in patients with UDCA-resistant primary biliary cirrhosis. Stem Cells Dev. 23 , 2482–2489 (2014).

Martel-Pelletier, J. et al. Osteoarthritis. Nat. Rev. Dis. Prim. 2 , 16072 (2016).

Olsson, S., Akbarian, E., Lind, A., Razavian, A. S. & Gordon, M. Automating classification of osteoarthritis according to Kellgren-Lawrence in the knee using deep learning in an unfiltered adult population. BMC Musculoskelet. Disord. 22 , 844 (2021).

Mahmoudian, A., Lohmander, L. S., Mobasheri, A., Englund, M. & Luyten, F. P. Early-stage symptomatic osteoarthritis of the knee—time for action. Nat. Rev. Rheumatol. 17 , 621–632 (2021).

Kubsik-Gidlewska, A. et al. CD34+ stem cell treatment for knee osteoarthritis: a treatment and rehabilitation algorithm. J. Rehabil. Med Clin. Commun. 3 , 1000012 (2018).

Jevotovsky, D. S., Alfonso, A. R., Einhorn, T. A. & Chiu, E. S. Osteoarthritis and stem cell therapy in humans: a systematic review. Osteoarthr. Cartil. 26 , 711–729 (2018).

Wiggers, T. G., Winters, M., Van den Boom, N. A., Haisma, H. J. & Moen, M. H. Autologous stem cell therapy in knee osteoarthritis: a systematic review of randomised controlled trials. Br. J. Sports Med 55 , 1161–1169 (2021).

Han, S. B., Seo, I. W. & Shin, Y. S. Intra-articular injections of hyaluronic acid or steroids associated with better outcomes than platelet-rich plasma, adipose mesenchymal stromal cells, or placebo in knee osteoarthritis: a network meta-analysis. Arthroscopy 37 , 292–306 (2021).

Bastos, R. et al. Intra-articular injections of expanded mesenchymal stem cells with and without addition of platelet-rich plasma are safe and effective for knee osteoarthritis. Knee Surg. Sports Traumatol. Arthrosc. 26 , 3342–3350 (2018).

Molnar, V. et al. Mesenchymal stem cell mechanisms of action and clinical effects in osteoarthritis: a narrative review. Genes 13 , 949 (2022).

Barisic, S. & Childs, R. W. Graft-versus-solid-tumor effect: from hematopoietic stem cell transplantation to adoptive cell therapies. Stem Cells 40 , 556–563 (2022).

Mello, M. M. & Brennan, T. A. The controversy over high-dose chemotherapy with autologous bone marrow transplant for breast cancer. Health Aff. (Millwood) 20 , 101–117 (2001).

Sissung, T. M. & Figg, W. D. Stem cell clinics: risk of proliferation. Lancet Oncol. 21 , 205–206 (2020).

Fu, X. et al. Mesenchymal stem cell migration and tissue repair. Cells 8 , 784 (2019).

Zachar, L., Bacenkova, D. & Rosocha, J. Activation, homing, and role of the mesenchymal stem cells in the inflammatory environment. J. Inflamm. Res. 9 , 231–240 (2016).

de Araujo Farias, V., Carrillo-Galvez, A. B., Martin, F. & Anderson, P. TGF-beta and mesenchymal stromal cells in regenerative medicine, autoimmunity and cancer. Cytokine Growth Factor Rev. 43 , 25–37 (2018).

Ding, W. et al. Platelet-derived growth factor (PDGF)-PDGF receptor interaction activates bone marrow-derived mesenchymal stromal cells derived from chronic lymphocytic leukemia: implications for an angiogenic switch. Blood 116 , 2984–2993 (2010).

Ritter, E. et al. Breast cancer cell-derived fibroblast growth factor 2 and vascular endothelial growth factor are chemoattractants for bone marrow stromal stem cells. Ann. Surg. 247 , 310–314 (2008).

Cronwright, G. et al. Cancer/testis antigen expression in human mesenchymal stem cells: down-regulation of SSX impairs cell migration and matrix metalloproteinase 2 expression. Cancer Res. 65 , 2207–2215 (2005).

Aldinucci, D., Borghese, C. & Casagrande, N. The CCL5/CCR5 axis in cancer progression. Cancers 12 , 1765 (2020).

Karnoub, A. E. et al. Mesenchymal stem cells within tumour stroma promote breast cancer metastasis. Nature 449 , 557–563 (2007).

Kucerova, L., Matuskova, M., Hlubinova, K., Altanerova, V. & Altaner, C. Tumor cell behaviour modulation by mesenchymal stromal cells. Mol. Cancer 9 , 129 (2010).

Schmohl, K. A., Muller, A. M., Nelson, P. J. & Spitzweg, C. Thyroid hormone effects on mesenchymal stem cell biology in the tumour microenvironment. Exp. Clin. Endocrinol. Diabetes 128 , 462–468 (2020).

Mishra, P. J. et al. Carcinoma-associated fibroblast-like differentiation of human mesenchymal stem cells. Cancer Res. 68 , 4331–4339 (2008).

Liu, J., Han, G., Liu, H. & Qin, C. Suppression of cholangiocarcinoma cell growth by human umbilical cord mesenchymal stem cells: a possible role of Wnt and Akt signaling. PLoS ONE 8 , e62844 (2013).

Ho, I. A. et al. Human bone marrow-derived mesenchymal stem cells suppress human glioma growth through inhibition of angiogenesis. Stem Cells 31 , 146–155 (2013).

Sun, Z., Wang, S. & Zhao, R. C. The roles of mesenchymal stem cells in tumor inflammatory microenvironment. J. Hematol. Oncol. 7 , 14 (2014).

Rhee, K. J., Lee, J. I. & Eom, Y. W. Mesenchymal stem cell-mediated effects of tumor support or suppression. Int. J. Mol. Sci. 16 , 30015–30033 (2015).

Liang, W. et al. Mesenchymal stem cells as a double-edged sword in tumor growth: focusing on MSC-derived cytokines. Cell Mol. Biol. Lett. 26 , 3 (2021).

Hmadcha, A., Martin-Montalvo, A., Gauthier, B. R., Soria, B. & Capilla-Gonzalez, V. Therapeutic potential of mesenchymal stem cells for cancer therapy. Front. Bioeng. Biotechnol. 8 , 43 (2020).

Cao, G. D. et al. The oncolytic virus in cancer diagnosis and treatment. Front. Oncol. 10 , 1786 (2020).

Melen, G. J. et al. Influence of carrier cells on the clinical outcome of children with neuroblastoma treated with high dose of oncolytic adenovirus delivered in mesenchymal stem cells. Cancer Lett. 371 , 161–170 (2016).

Garcia-Castro, J. et al. Treatment of metastatic neuroblastoma with systemic oncolytic virotherapy delivered by autologous mesenchymal stem cells: an exploratory study. Cancer Gene Ther. 17 , 476–483 (2010).

Draganov, D. D. et al. Delivery of oncolytic vaccinia virus by matched allogeneic stem cells overcomes critical innate and adaptive immune barriers. J. Transl. Med. 17 , 100 (2019).

Cyranoski, D. How human embryonic stem cells sparked a revolution. Nature 555 , 428–430 (2018).

Thomson, J. A. et al. Embryonic stem cell lines derived from human blastocysts. Science 282 , 1145–1147 (1998).

Takahashi, K. & Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126 , 663–676 (2006).

Takahashi, K. et al. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131 , 861–872 (2007).

Gepstein, L. Derivation and potential applications of human embryonic stem cells. Circ. Res. 91 , 866–876 (2002).

Andrews, P. W. From teratocarcinomas to embryonic stem cells. Philos. Trans. R. Soc. Lond. B Biol. Sci. 357 , 405–417 (2002).

Finch, B. W. & Ephrussi, B. Retention of multiple developmental potentialities by cells of a mouse testicular teratocarcinoma during prolonged culture in vitro and their extinction upon hybridization with cells of permanent lines. Proc. Natl Acad. Sci. USA 57 , 615–621 (1967).

Ried, T. et al. The consequences of chromosomal aneuploidy on the transcriptome of cancer cells. Biochim Biophys. Acta 1819 , 784–793 (2012).

Evans, M. J. & Kaufman, M. H. Establishment in culture of pluripotential cells from mouse embryos. Nature 292 , 154–156 (1981).

Martin, G. R. Isolation of a pluripotent cell line from early mouse embryos cultured in medium conditioned by teratocarcinoma stem cells. Proc. Natl Acad. Sci. USA 78 , 7634–7638 (1981).

Lo, B. & Parham, L. Ethical issues in stem cell research. Endocr. Rev. 30 , 204–213 (2009).

Wilmut, I., Schnieke, A. E., McWhir, J., Kind, A. J. & Campbell, K. H. Viable offspring derived from fetal and adult mammalian cells. Nature 385 , 810–813 (1997).

Schwartz, S. D. et al. Embryonic stem cell trials for macular degeneration: a preliminary report. Lancet 379 , 713–720 (2012).

Atala, A. Human embryonic stem cells: early hints on safety and efficacy. Lancet 379 , 689–690 (2012).

Schwartz, S. D. et al. Human embryonic stem cell-derived retinal pigment epithelium in patients with age-related macular degeneration and Stargardt’s macular dystrophy: follow-up of two open-label phase 1/2 studies. Lancet 385 , 509–516 (2015).

Song, W. K. et al. Treatment of macular degeneration using embryonic stem cell-derived retinal pigment epithelium: preliminary results in Asian patients. Stem Cell Rep. 4 , 860–872 (2015).

Liu, Y. et al. Human embryonic stem cell-derived retinal pigment epithelium transplants as a potential treatment for wet age-related macular degeneration. Cell Discov. 4 , 50 (2018).

Limnios, I. J., Chau, Y. Q., Skabo, S. J., Surrao, D. C. & O’Neill, H. C. Efficient differentiation of human embryonic stem cells to retinal pigment epithelium under defined conditions. Stem Cell Res. Ther. 12 , 248 (2021).

Foltz, L. P. & Clegg, D. O. Rapid, directed differentiation of retinal pigment epithelial cells from human embryonic or induced pluripotent stem cells. J. Vis. Exp. 128 , e56274 (2017).

Kuroda, T., Ando, S., Takeno, Y., Kishino, A. & Kimura, T. Robust induction of retinal pigment epithelium cells from human induced pluripotent stem cells by inhibiting FGF/MAPK signaling. Stem Cell Res 39 , 101514 (2019).

Dewell, T. E. et al. Transcription factor overexpression drives reliable differentiation of retinal pigment epithelium from human induced pluripotent stem cells. Stem Cell Res. 53 , 102368 (2021).

Dehghan, S., Mirshahi, R., Shoae-Hassani, A. & Naseripour, M. Human-induced pluripotent stem cells-derived retinal pigmented epithelium, a new horizon for cells-based therapies for age-related macular degeneration. Stem Cell Res. Ther. 13 , 217 (2022).

Menasche, P. et al. Human embryonic stem cell-derived cardiac progenitors for severe heart failure treatment: first clinical case report. Eur. Heart J. 36 , 2011–2017 (2015).

Menasche, P. et al. Transplantation of human embryonic stem cell-derived cardiovascular progenitors for severe ischemic left ventricular dysfunction. J. Am. Coll. Cardiol. 71 , 429–438 (2018).

Cyranoski, D. ‘Reprogrammed’ stem cells approved to mend human hearts for the first time. Nature 557 , 619–620 (2018).

Povsic, T. J. & Gersh, B. J. Stem cells in cardiovascular diseases: 30,000-foot view. Cells 10 , 600 (2021).

Romito, A. & Cobellis, G. Pluripotent stem cells: current understanding and future directions. Stem Cells Int. 2016 , 9451492 (2016).

McKenna, S. L. et al. Ten-year safety of pluripotent stem cell transplantation in acute thoracic spinal cord injury. J. Neurosurg. Spine 1 , 1–10 (2022).

Deinsberger, J., Reisinger, D. & Weber, B. Global trends in clinical trials involving pluripotent stem cells: a systematic multi-database analysis. NPJ Regen. Med. 5 , 15 (2020).

Kim, J. Y., Nam, Y., Rim, Y. A. & Ju, J. H. Review of the current trends in clinical trials involving induced pluripotent stem cells. Stem Cell Rev. Rep. 18 , 142–154 (2022).

Ji, P., Manupipatpong, S., Xie, N. & Li, Y. Induced pluripotent stem cells: generation strategy and epigenetic mystery behind reprogramming. Stem Cells Int. 2016 , 8415010 (2016).

Fu, X. The immunogenicity of cells derived from induced pluripotent stem cells. Cell Mol. Immunol. 11 , 14–16 (2014).

Lee, A. S., Tang, C., Rao, M. S., Weissman, I. L. & Wu, J. C. Tumorigenicity as a clinical hurdle for pluripotent stem cell therapies. Nat. Med. 19 , 998–1004 (2013).

Friedenstein, A. J., Piatetzky, S. II & Petrakova, K. V. Osteogenesis in transplants of bone marrow cells. J. Embryol. Exp. Morphol. 16 , 381–390 (1966).

Pittenger, M. F. et al. Multilineage potential of adult human mesenchymal stem cells. Science 284 , 143–147 (1999).

Nombela-Arrieta, C., Ritz, J. & Silberstein, L. E. The elusive nature and function of mesenchymal stem cells. Nat. Rev. Mol. Cell Biol. 12 , 126–131 (2011).

Dominici, M. et al. Minimal criteria for defining multipotent mesenchymal stromal cells. The International Society for Cellular Therapy position statement. Cytotherapy 8 , 315–317 (2006).

Zhou, T. et al. Challenges and advances in clinical applications of mesenchymal stromal cells. J. Hematol. Oncol. 14 , 24 (2021).

Ankrum, J. & Karp, J. M. Mesenchymal stem cell therapy: two steps forward, one step back. Trends Mol. Med. 16 , 203–209 (2010).

Tuan, R. S., Boland, G. & Tuli, R. Adult mesenchymal stem cells and cell-based tissue engineering. Arthritis Res. Ther. 5 , 32–45 (2003).

Witkowska-Zimny, M. & Wrobel, E. Perinatal sources of mesenchymal stem cells: Wharton’s jelly, amnion and chorion. Cell Mol. Biol. Lett. 16 , 493–514 (2011).

Alkhalil, M., Smajilagic, A. & Redzic, A. Human dental pulp mesenchymal stem cells isolation and osteoblast differentiation. Med. Glas. (Zenica) 12 , 27–32 (2015).

Google Scholar  

Ouryazdanpanah, N., Dabiri, S., Derakhshani, A., Vahidi, R. & Farsinejad, A. Peripheral blood-derived mesenchymal stem cells: growth factor-free isolation, molecular characterization and differentiation. Iran. J. Pathol. 13 , 461–466 (2018).

PubMed   PubMed Central   Google Scholar  

Francis, M. P., Sachs, P. C., Elmore, L. W. & Holt, S. E. Isolating adipose-derived mesenchymal stem cells from lipoaspirate blood and saline fraction. Organogenesis 6 , 11–14 (2010).

Gong, X. et al. Isolation and characterization of lung resident mesenchymal stem cells capable of differentiating into alveolar epithelial type II cells. Cell Biol. Int. 38 , 405–411 (2014).

Wang, B. et al. Human hair follicle-derived mesenchymal stem cells: Isolation, expansion, and differentiation. World J. Stem Cells 12 , 462–470 (2020).

Pilato, C. A. et al. Isolation and characterization of cardiac mesenchymal stromal cells from endomyocardial bioptic samples of arrhythmogenic cardiomyopathy patients. J. Vis. Exp . 132 , e57263 (2018).

Mannino, G. et al. Adult stem cell niches for tissue homeostasis. J. Cell Physiol. 237 , 239–257 (2022).

Pavlushina, S. V., Orlova, T. G. & Tabagari, D. Z. [Isolation of mononuclear cells from the bone marrow of patients with hemoblastoses using one-step ficoll-verographin density gradient separation]. Eksp. Onkol. 6 , 68–70 (1984).

Schneider, S., Unger, M., van Griensven, M. & Balmayor, E. R. Adipose-derived mesenchymal stem cells from liposuction and resected fat are feasible sources for regenerative medicine. Eur. J. Med Res. 22 , 17 (2017).

Torre, P. & Flores, A. I. Current status and future prospects of perinatal stem cells. Genes 12 , 6 (2020).

Hoang, V. T. et al. Standardized xeno- and serum-free culture platform enables large-scale expansion of high-quality mesenchymal stem/stromal cells from perinatal and adult tissue sources. Cytotherapy 23 , 88–99 (2020).

Mohamed-Ahmed, S. et al. Adipose-derived and bone marrow mesenchymal stem cells: a donor-matched comparison. Stem Cell Res. Ther. 9 , 168 (2018).

Zuk, P. A. et al. Human adipose tissue is a source of multipotent stem cells. Mol. Biol. Cell 13 , 4279–4295 (2002).

Li, Z. CD133: a stem cell biomarker and beyond. Exp. Hematol. Oncol. 2 , 17 (2013).

Petrenko, Y. et al. A comparative analysis of multipotent mesenchymal stromal cells derived from different sources, with a focus on neuroregenerative potential. Sci. Rep. 10 , 4290 (2020).

Wang, Z. & Yan, X. CD146, a multi-functional molecule beyond adhesion. Cancer Lett. 330 , 150–162 (2013).

Xu, L. et al. Tissue source determines the differentiation potentials of mesenchymal stem cells: a comparative study of human mesenchymal stem cells from bone marrow and adipose tissue. Stem Cell Res Ther. 8 , 275 (2017).

Han, I., Kwon, B. S., Park, H. K. & Kim, K. S. Differentiation potential of mesenchymal stem cells is related to their intrinsic mechanical properties. Int. Neurourol. J. 21 , S24–S31 (2017).

Song, Y. et al. Human mesenchymal stem cells derived from umbilical cord and bone marrow exert immunomodulatory effects in different mechanisms. World J. Stem Cells 12 , 1032–1049 (2020).

Lee, R. H. et al. Intravenous hMSCs improve myocardial infarction in mice because cells embolized in lung are activated to secrete the anti-inflammatory protein TSG-6. Cell Stem Cell 5 , 54–63 (2009).

Allers, C. et al. Dynamic of distribution of human bone marrow-derived mesenchymal stem cells after transplantation into adult unconditioned mice. Transplantation 78 , 503–508 (2004).

Devine, S. M., Cobbs, C., Jennings, M., Bartholomew, A. & Hoffman, R. Mesenchymal stem cells distribute to a wide range of tissues following systemic infusion into nonhuman primates. Blood 101 , 2999–3001 (2003).

Fischer, U. M. et al. Pulmonary passage is a major obstacle for intravenous stem cell delivery: the pulmonary first-pass effect. Stem Cells Dev. 18 , 683–692 (2009).

Sierra-Parraga, J. M. et al. Mesenchymal stromal cells are retained in the porcine renal cortex independently of their metabolic state after renal intra-arterial infusion. Stem Cells Dev. 28 , 1224–1235 (2019).

Henriksson, H. B. et al. The traceability of mesenchymal stromal cells after injection into degenerated discs in patients with low back pain. Stem Cells Dev. 28 , 1203–1211 (2019).

Sokal, E. M. et al. Biodistribution of liver-derived mesenchymal stem cells after peripheral injection in a hemophilia A patient. Transplantation 101 , 1845–1851 (2017).

Sood, V. et al. Biodistribution of 18F-FDG-labeled autologous bone marrow-derived stem cells in patients with type 2 diabetes mellitus: exploring targeted and intravenous routes of delivery. Clin. Nucl. Med. 40 , 697–700 (2015).

Sanchez-Diaz, M. et al. Biodistribution of mesenchymal stromal cells after administration in animal models and humans: a systematic review. J. Clin. Med. 10 , 2925 (2021).

Sensebe, L. & Fleury-Cappellesso, S. Biodistribution of mesenchymal stem/stromal cells in a preclinical setting. Stem Cells Int. 2013 , 678063 (2013).

Zhuang, W. Z. et al. Mesenchymal stem/stromal cell-based therapy: mechanism, systemic safety and biodistribution for precision clinical applications. J. Biomed. Sci. 28 , 28 (2021).

Wei, X. et al. Mesenchymal stem cells: a new trend for cell therapy. Acta Pharm. Sin. 34 , 747–754 (2013).

Kouchakian, M. R. et al. The clinical trials of mesenchymal stromal cells therapy. Stem Cells Int. 2021 , 1634782 (2021).

Chen, L. et al. Mesenchymal stem cell-based treatments for COVID-19: status and future perspectives for clinical applications. Cell Mol. Life Sci. 79 , 142 (2022).

Borow, K. M., Yaroshinsky, A., Greenberg, B. & Perin, E. C. Phase 3 DREAM-HF trial of mesenchymal precursor cells in chronic heart failure. Circ. Res. 125 , 265–281 (2019).

Zheng, H. et al. Mesenchymal stem cell therapy in stroke: a systematic review of literature in pre-clinical and clinical research. Cell Transpl. 27 , 1723–1730 (2018).

Rodriguez-Fuentes, D. E. et al. Mesenchymal stem cells current clinical applications: a systematic review. Arch. Med. Res. 52 , 93–101 (2021).

Shi, L. et al. Mesenchymal stem cell therapy for severe COVID-19. Signal Transduct. Target Ther. 6 , 339 (2021).

Carney, B. J. & Shah, K. Migration and fate of therapeutic stem cells in different brain disease models. Neuroscience 197 , 37–47 (2011).

Yao, P., Zhou, L., Zhu, L., Zhou, B. & Yu, Q. Mesenchymal stem cells: a potential therapeutic strategy for neurodegenerative diseases. Eur. Neurol. 83 , 235–241 (2020).

Bonaventura, G. et al. Stem cells: innovative therapeutic options for neurodegenerative diseases? Cells 10 , 1992 (2021).

Mansoor, S. R., Zabihi, E. & Ghasemi-Kasman, M. The potential use of mesenchymal stem cells for the treatment of multiple sclerosis. Life Sci. 235 , 116830 (2019).

Chung, J. W. et al. Efficacy and safety of intravenous mesenchymal stem cells for ischemic stroke. Neurology 96 , e1012–e1023 (2021).

Yamazaki, K., Kawabori, M., Seki, T. & Houkin, K. Clinical trials of stem cell treatment for spinal cord injury. Int. J. Mol. Sci . 21 , 3994 (2020).

Xie, B., Chen, M., Hu, R., Han, W. & Ding, S. Therapeutic evidence of human mesenchymal stem cell transplantation for cerebral palsy: a meta-analysis of randomized controlled trials. Stem Cells Int. 2020 , 5701920 (2020).

McDonald, C. A. et al. Intranasal delivery of mesenchymal stromal cells protects against neonatal hypoxic(-)ischemic brain injury. Int. J. Mol. Sci . 20 , 2449 (2019).

Liu, Q. et al. Rational use of mesenchymal stem cells in the treatment of autism spectrum disorders. World J. Stem Cells 11 , 55–72 (2019).

Fricova, D., Korchak, J. A. & Zubair, A. C. Challenges and translational considerations of mesenchymal stem/stromal cell therapy for Parkinson’s disease. npj Regen. Med. 5 , 20 (2020).

Bang, O. Y., Lee, J. S., Lee, P. H. & Lee, G. Autologous mesenchymal stem cell transplantation in stroke patients. Ann. Neurol. 57 , 874–882 (2005).

Lee, J. S. et al. A long-term follow-up study of intravenous autologous mesenchymal stem cell transplantation in patients with ischemic stroke. Stem Cells 28 , 1099–1106 (2010).

Honmou, O. et al. Intravenous administration of auto serum-expanded autologous mesenchymal stem cells in stroke. Brain 134 , 1790–1807 (2011).

Bhasin, A. et al. Autologous mesenchymal stem cells in chronic stroke. Cerebrovasc. Dis. Extra 1 , 93–104 (2011).

Jaillard, A. et al. Autologous mesenchymal stem cells improve motor recovery in subacute ischemic stroke: a randomized clinical trial. Transl. Stroke Res. 11 , 910–923 (2020).

Lee, J. et al. Efficacy of intravenous mesenchymal stem cells for motor recovery after ischemic stroke: a neuroimaging study. Stroke 53 , 20–28 (2022).

Levy, M. L. et al. Phase I/II study of safety and preliminary efficacy of intravenous allogeneic mesenchymal stem cells in chronic stroke. Stroke 50 , 2835–2841 (2019).

Xu, P. & Yang, X. The efficacy and safety of mesenchymal stem cell transplantation for spinal cord injury patients: a meta-analysis and systematic review. Cell Transpl. 28 , 36–46 (2019).

Liau, L. L. et al. Treatment of spinal cord injury with mesenchymal stem cells. Cell Biosci. 10 , 112 (2020).

Liu, X. et al. Comparative analysis of curative effect of bone marrow mesenchymal stem cell and bone marrow mononuclear cell transplantation for spastic cerebral palsy. J. Transl. Med. 15 , 1–9 (2017).

Sharma, A. K. et al. Cell transplantation as a novel therapeutic strategy for autism spectrum disorders: a clinical study. Am J. Stem Cells 9 , 89 (2020).

Ballen, K. & Kurtzberg, J. Exploring new therapies for children with autism: “Do no harm” does not mean do not try. Stem Cells Transl. Med. 10 , 823–825 (2021).

Reyhani, S., Abbaspanah, B. & Mousavi, S. H. Umbilical cord-derived mesenchymal stem cells in neurodegenerative disorders: from literature to clinical practice. Regen. Med. 15 , 1561–1578 (2020).

Gu, J. et al. Therapeutic evidence of umbilical cord-derived mesenchymal stem cell transplantation for cerebral palsy: a randomized, controlled trial. Stem Cell Res Ther. 11 , 43 (2020).

Retraction. Stem Cells Transl. Med. 10 , 1717 (2021).

Sun, J. M. et al. Infusion of human umbilical cord tissue mesenchymal stromal cells in children with autism spectrum disorder. Stem Cells Transl. Med. 9 , 1137–1146 (2020).

Yang, Y. et al. Repeated subarachnoid administrations of allogeneic human umbilical cord mesenchymal stem cells for spinal cord injury: a phase 1/2 pilot study. Cytotherapy 23 , 57–64 (2021).

Liu, J. et al. Clinical analysis of the treatment of spinal cord injury with umbilical cord mesenchymal stem cells. Cytotherapy 15 , 185–191 (2013).

Przekora, A. & Juszkiewicz, L. The effect of autologous adipose tissue-derived mesenchymal stem cells’ therapy in the treatment of chronic posttraumatic spinal cord injury in a domestic ferret patient. Cell Transpl. 29 , 963689720928982 (2020).

Hur, J. W. et al. Intrathecal transplantation of autologous adipose-derived mesenchymal stem cells for treating spinal cord injury: a human trial. J. Spinal Cord. Med. 39 , 655–664 (2016).

de Celis-Ruiz, E. et al. Final results of allogeneic adipose tissue-derived mesenchymal stem cells in acute ischemic stroke (AMASCIS): a phase II, randomized, double-blind, placebo-controlled, single-center, pilot clinical trial. Cell Transpl. 31 , 9636897221083863 (2022).

Yang, Y. et al. Human umbilical cord mesenchymal stem cells to treat spinal cord injury in the early chronic phase: study protocol for a prospective, multicenter, randomized, placebo-controlled, single-blinded clinical trial. Neural Regen. Res. 15 , 1532–1538 (2020).

de Celis-Ruiz, E. et al. Allogeneic adipose tissue-derived mesenchymal stem cells in ischaemic stroke (AMASCIS-02): a phase IIb, multicentre, double-blind, placebo-controlled clinical trial protocol. BMJ Open 11 , e051790 (2021).

Murray, C. J. L. COVID-19 will continue but the end of the pandemic is near. Lancet 399 , 417–419 (2022).

Thebaud, B. et al. Bronchopulmonary dysplasia. Nat. Rev. Dis. Prim. 5 , 78 (2019).

Mohamed, T., Abdul-Hafez, A., Gewolb, I. H. & Uhal, B. D. Oxygen injury in neonates: which is worse? hyperoxia, hypoxia, or alternating hyperoxia/hypoxia. J. Lung Pulm. Respir. Res. 7 , 4–13 (2020).

Omar, S. A. et al. Stem-cell therapy for bronchopulmonary dysplasia (BPD) in newborns. Cells 11 , 1275 (2022).

Chang, Y. S. et al. Mesenchymal stem cells for bronchopulmonary dysplasia: phase 1 dose-escalation clinical trial. J. Pediatr. 164 , 966–972 e966 (2014).

Powell, S. B. & Silvestri, J. M. Safety of intratracheal administration of human umbilical cord blood derived mesenchymal stromal cells in extremely low birth weight preterm infants. J. Pediatr. 210 , 209–213 e202 (2019).

Nguyen, L. T. et al. Allogeneic administration of human umbilical cord-derived mesenchymal stem/stromal cells for bronchopulmonary dysplasia: preliminary outcomes in four Vietnamese infants. J. Transl. Med. 18 , 398 (2020).

Ahn, S. Y. et al. Stem cells for bronchopulmonary dysplasia in preterm infants: a randomized controlled phase II trial. Stem Cells Transl. Med. 10 , 1129–1137 (2021).

Averyanov, A. et al. First-in-human high-cumulative-dose stem cell therapy in idiopathic pulmonary fibrosis with rapid lung function decline. Stem Cells Transl. Med. 9 , 6–16 (2020).

Ribeiro-Paes, J. T. et al. Unicentric study of cell therapy in chronic obstructive pulmonary disease/pulmonary emphysema. Int. J. Chron. Obstruct Pulmon Dis. 6 , 63–71 (2011).

Stessuk, T. et al. Phase I clinical trial of cell therapy in patients with advanced chronic obstructive pulmonary disease: follow-up of up to 3 years. Rev. Bras. Hematol. Hemoter. 35 , 352–357 (2013).

Weiss, D. J., Casaburi, R., Flannery, R., LeRoux-Williams, M. & Tashkin, D. P. A placebo-controlled, randomized trial of mesenchymal stem cells in COPD. Chest 143 , 1590–1598 (2013).

de Oliveira, H. G. et al. Combined bone marrow-derived mesenchymal stromal cell therapy and one-way endobronchial valve placement in patients with pulmonary emphysema: a phase I clinical trial. Stem Cells Transl. Med. 6 , 962–969 (2017).

Stolk, J. et al. A phase I study for intravenous autologous mesenchymal stromal cell administration to patients with severe emphysema. QJM 109 , 331–336 (2016).

Armitage, J. et al. Mesenchymal stromal cell infusion modulates systemic immunological responses in stable COPD patients: a phase I pilot study. Eur. Respir. J. 51 , 1702369 (2018).

Comella, K. et al. Autologous stromal vascular fraction in the intravenous treatment of end-stage chronic obstructive pulmonary disease: a phase I trial of safety and tolerability. J. Clin. Med. Res. 9 , 701–708 (2017).

Tzilas, V. et al . Prospective phase 1 open clinical trial to study the safety of adipose derived mesenchymal stem cells (ADMSCs) in COPD and combined pulmonary fibrosis and emphysema (CPFE). Eur. Respir. J. 46 , (2015).

Glassberg, M. K., Csete, I., Simonet, E. & Elliot, S. J. Stem cell therapy for COPD: hope and exploitation. Chest 160 , 1271–1281 (2021).

Le Thi Bich, P. et al. Allogeneic umbilical cord-derived mesenchymal stem cell transplantation for treating chronic obstructive pulmonary disease: a pilot clinical study. Stem Cell Res. Ther. 60 , 11 (2020).

Karaoz, E., Kalemci, S. & Ece, F. Improving effects of mesenchymal stem cells on symptoms of chronic obstructive pulmonary disease. Bratisl. Lek. Listy. 121 , 188–191 (2020).

Hoang, D. M., Nguyen, K. T., Nguyen, A. H., Nguyen, B. N. & Nguyen, L. T. Allogeneic human umbilical cord-derived mesenchymal stem/stromal cells for chronic obstructive pulmonary disease (COPD): study protocol for a matched case-control, phase I/II trial. BMJ Open 11 , e045788 (2021).

Xu, R., Feng, Z. & Wang, F. S. Mesenchymal stem cell treatment for COVID-19. EBioMedicine 77 , 103920 (2022).

Khoury, M. et al. Current status of cell-based therapies for respiratory virus infections: applicability to COVID-19. Eur. Respir. J. 55 , 2000858 (2020).

Jamilloux, Y. et al. Should we stimulate or suppress immune responses in COVID-19? Cytokine and anti-cytokine interventions. Autoimmun. Rev. 19 , 102567 (2020).

Feng, Y. et al. Safety and feasibility of umbilical cord mesenchymal stem cells in patients with COVID-19 pneumonia: a pilot study. Cell Prolif. 53 , e12947 (2020).

Primorac, D. et al. Mesenchymal stromal cells: potential option for COVID-19 treatment. Pharmaceutic 13 , 1481 (2021).

Zhang, Y. et al. Intravenous infusion of human umbilical cord Wharton’s jelly-derived mesenchymal stem cells as a potential treatment for patients with COVID-19 pneumonia. Stem Cell Res. Ther. 11 , 207 (2020).

Shu, L. et al. Treatment of severe COVID-19 with human umbilical cord mesenchymal stem cells. Stem Cell Res. Ther. 11 , 361 (2020).

Tao, J. et al. Umbilical cord blood-derived mesenchymal stem cells in treating a critically ill COVID-19 patient. J. Infect. Dev. Ctries 14 , 1138–1145 (2020).

Saleh, M. et al. Cell therapy in patients with COVID-19 using Wharton’s jelly mesenchymal stem cells: a phase 1 clinical trial. Stem Cell Res. Ther. 12 , 410 (2021).

Leng, Z. et al. Transplantation of ACE2(-) mesenchymal stem cells improves the outcome of patients with COVID-19 pneumonia. Aging Dis. 11 , 216–228 (2020).

Guo, Z. et al. Administration of umbilical cord mesenchymal stem cells in patients with severe COVID-19 pneumonia. Crit. Care 24 , 420 (2020).

Meng, F. et al. Human umbilical cord-derived mesenchymal stem cell therapy in patients with COVID-19: a phase 1 clinical trial. Signal Transduct. Target Ther. 5 , 172 (2020).

Shi, L. et al. Human mesenchymal stem cells treatment for severe COVID-19: 1-year follow-up results of a randomized, double-blind, placebo-controlled trial. EBioMedicine 75 , 103789 (2021).

Adas, G. et al. The systematic effect of mesenchymal stem cell therapy in critical COVID-19 patients: a prospective double controlled trial. Cell Transpl. 30 , 9636897211024942 (2021).

Shi, L. et al. Effect of human umbilical cord-derived mesenchymal stem cells on lung damage in severe COVID-19 patients: a randomized, double-blind, placebo-controlled phase 2 trial. Signal Transduct. Targeted Ther. 6 , 58 (2021).

Lanzoni, G. et al. Umbilical cord mesenchymal stem cells for COVID-19 acute respiratory distress syndrome: a double-blind, phase 1/2a, randomized controlled trial. Stem Cells Transl. Med. 10 , 660–673 (2021).

Hashemian, S. R. et al. Mesenchymal stem cells derived from perinatal tissues for treatment of critically ill COVID-19-induced ARDS patients: a case series. Stem Cell Res Ther. 12 , 91 (2021).

Zhu, R. et al. Mesenchymal stem cell treatment improves outcome of COVID-19 patients via multiple immunomodulatory mechanisms. Cell Res. 31 , 1244–1262 (2021).

Shi, L. et al. Effect of human umbilical cord-derived mesenchymal stem cells on lung damage in severe COVID-19 patients: a randomized, double-blind, placebo-controlled phase 2 trial. Signal Transduct. Target Ther. 6 , 58 (2021).

N, O. E., Pekkoc-Uyanik, K. C., Alpaydin, N., Gulay, G. R. & Simsek, M. Clinical experience on umbilical cord mesenchymal stem cell treatment in 210 severe and critical COVID-19 cases in Turkey. Stem Cell Rev. Rep. 17 , 1917–1925 (2021).

Gentile, P., Sterodimas, A., Pizzicannella, J., Calabrese, C. & Garcovich, S. Research progress on mesenchymal stem cells (MSCs), adipose-derived mesenchymal stem cells (AD-MSCs), drugs, and vaccines in inhibiting COVID-19 disease. Aging Dis. 11 , 1191–1201 (2020).

Copcu, H. E. Potential using of fat-derived stromal cells in the treatment of active disease, and also, in both pre- and post-periods in COVID-19. Aging Dis. 11 , 730–736 (2020).

Gentile, P. & Sterodimas, A. Adipose-derived stromal stem cells (ASCs) as a new regenerative immediate therapy combating coronavirus (COVID-19)-induced pneumonia. Expert Opin. Biol. Ther. 20 , 711–716 (2020).

Matthay, M. A. et al. Treatment with allogeneic mesenchymal stromal cells for moderate to severe acute respiratory distress syndrome (START study): a randomised phase 2a safety trial. Lancet Respir. Med. 7 , 154–162 (2019).

Álvarez-Fuente, M. et al. Off-label mesenchymal stromal cell treatment in two infants with severe bronchopulmonary dysplasia: clinical course and biomarkers profile. Cytotherapy 20 , 1337–1344 (2018).

Zheng, G. et al. Treatment of acute respiratory distress syndrome with allogeneic adipose-derived mesenchymal stem cells: a randomized, placebo-controlled pilot study. Respir. Res. 15 , 39 (2014).

Simonson, O. E. et al. In vivo effects of mesenchymal stromal cells in two patients with severe acute respiratory distress syndrome. Stem Cells Transl. Med. 4 , 1199–1213 (2015).

Wilson, J. G. et al. Mesenchymal stem (stromal) cells for treatment of ARDS: a phase 1 clinical trial. Lancet Respir. Med. 3 , 24–32 (2015).

Yip, H. K. et al. Human umbilical cord-derived mesenchymal stem cells for acute respiratory distress syndrome. Crit. Care Med 48 , e391–e399 (2020).

Gorman, E. et al. Repair of acute respiratory distress syndrome by stromal cell administration (REALIST) trial: a phase 1 trial. EClinicalMedicine 41 , 101167 (2021).

Le Thi Bich, P. et al. Allogeneic umbilical cord-derived mesenchymal stem cell transplantation for treating chronic obstructive pulmonary disease: a pilot clinical study. Stem Cell Res. Ther. 11 , 60 (2020).

Wang, M. Y. et al. Current therapeutic strategies for respiratory diseases using mesenchymal stem cells. MedComm 2 , 351–380 (2021).

Carlsson, P. O., Schwarcz, E., Korsgren, O. & Le Blanc, K. Preserved beta-cell function in type 1 diabetes by mesenchymal stromal cells. Diabetes 64 , 587–592 (2015).

Dantas, J. R. et al. Adipose tissue-derived stromal/stem cells + cholecalciferol: a pilot study in recent-onset type 1 diabetes patients. Arch. Endocrinol. Metab. 65 , 342–351 (2021).

PubMed   Google Scholar  

Joseph, U. A. & Jhingran, S. G. Technetium-99m labeled RBC imaging in gastrointestinal bleeding from gastric leiomyoma. Clin. Nucl. Med. 13 , 23–25 (1988).

Hu, J. et al. Long term effects of the implantation of Wharton’s jelly-derived mesenchymal stem cells from the umbilical cord for newly-onset type 1 diabetes mellitus. Endocr. J. 60 , 347–357 (2013).

Cai, J. et al. Umbilical cord mesenchymal stromal cell with autologous bone marrow cell transplantation in established type 1 diabetes: a pilot randomized controlled open-label clinical study to assess safety and impact on insulin secretion. Diabetes Care 39 , 149–157 (2016).

Huang, Q., Huang, Y. & Liu, J. Mesenchymal stem cells: an excellent candidate for the treatment of diabetes mellitus. Int. J. Endocrinol. 2021 , 9938658 (2021).

Nguyen, L. T. et al. Type 2 diabetes mellitus duration and obesity alter the efficacy of autologously transplanted bone marrow-derived mesenchymal stem/stromal cells. Stem Cells Transl. Med. 10 , 1266–1278 (2021).

Alicka, M., Major, P., Wysocki, M. & Marycz, K. Adipose-derived mesenchymal stem cells isolated from patients with type 2 diabetes show reduced “stemness” through an altered secretome profile, impaired anti-oxidative protection, and mitochondrial dynamics deterioration. J. Clin. Med. 8 , 765 (2019).

Agarwal, A. et al. Male infertility. Lancet 397 , 319–333 (2021).

Farquhar, C. & Marjoribanks, J. Assisted reproductive technology: an overview of Cochrane reviews. Cochrane Database Syst. Rev. 8 , CD010537 (2018).

Chang, Z. et al. Mesenchymal stem cells in preclinical infertility cytotherapy: a retrospective review. Stem Cells Int. 2021 , 8882368 (2021).

Fenton, A. J. Premature ovarian insufficiency: pathogenesis and management. J. Midlife Health 6 , 147–153 (2015).

Coulam, C. B. Premature gonadal failure. Fertil. Steril. 38 , 645–655 (1982).

Huhtaniemi, I. et al. Advances in the molecular pathophysiology, genetics, and treatment of primary ovarian insufficiency. Trends Endocrinol. Metab. 29 , 400–419 (2018).

Torrealday, S., Kodaman, P. & Pal, L. Premature ovarian Insufficiency—an update on recent advances in understanding and management. F1000Res 6 , 2069 (2017).

Gupta, S., Lodha, P., Karthick, M. S. & Tandulwadkar, S. R. Role of autologous bone marrow-derived stem cell therapy for follicular recruitment in premature ovarian insufficiency: review of literature and a case report of world’s first baby with ovarian autologous stem cell therapy in a perimenopausal woman of age 45 year. J. Hum. Reprod. Sci. 11 , 125–130 (2018).

Igboeli, P. et al. Intraovarian injection of autologous human mesenchymal stem cells increases estrogen production and reduces menopausal symptoms in women with premature ovarian failure: two case reports and a review of the literature. J. Med. Case Rep. 14 , 108 (2020).

Ulin, M. et al. Human mesenchymal stem cell therapy and other novel treatment approaches for premature ovarian insufficiency. Reprod. Sci. 28 , 1688–1696 (2021).

Herraiz, S. et al. Autologous stem cell ovarian transplantation to increase reproductive potential in patients who are poor responders. Fertil. Steril. 110 , 496–505 e491 (2018).

Ding, L. et al. Transplantation of UC-MSCs on collagen scaffold activates follicles in dormant ovaries of POF patients with long history of infertility. Sci. China Life Sci. 61 , 1554–1565 (2018).

Wang, M. Y., Wang, Y. X., Li-Ling, J. & Xie, H. Q. Adult stem cell therapy for premature ovarian failure: from bench to bedside. Tissue Eng. Part B Rev. 28 , 63–78 (2022).

Kaddoura, I., Abu-Sittah, G., Ibrahim, A., Karamanoukian, R. & Papazian, N. Burn injury: review of pathophysiology and therapeutic modalities in major burns. Ann. Burns Fire Disasters 30 , 95–102 (2017).

PubMed   PubMed Central   CAS   Google Scholar  

Jeschke, M. G. et al. Burn injury. Nat. Rev. Dis. Prim. 6 , 11 (2020).

Rasulov, M. F. et al. First experience of the use bone marrow mesenchymal stem cells for the treatment of a patient with deep skin burns. Bull. Exp. Biol. Med. 139 , 141–144 (2005).

Mansilla, E. et al. Cadaveric bone marrow mesenchymal stem cells: first experience treating a patient with large severe burns. Burns Trauma 3 , 17 (2015).

Xu, Y., Huang, S. & Fu, X. Autologous transplantation of bone marrow-derived mesenchymal stem cells: a promising therapeutic strategy for prevention of skin-graft contraction. Clin. Exp. Dermatol. 37 , 497–500 (2012).

Yoshikawa, T. et al. Wound therapy by marrow mesenchymal cell transplantation. Plast. Reconstr. Surg. 121 , 860–877 (2008).

Abo-Elkheir, W. et al. Role of cord blood and bone marrow mesenchymal stem cells in recent deep burn: a case-control prospective study. Am. J. Stem Cells 6 , 23–35 (2017).

Li, L. et al. Conditioned medium from human adipose-derived mesenchymal stem cell culture prevents UVB-induced skin aging in human keratinocytes and dermal fibroblasts. Int. J. Mol. Sci. 21 , 49 (2019).

Lotfi, M. et al. Adipose tissue-derived mesenchymal stem cells and keratinocytes co-culture on gelatin/chitosan/beta-glycerol phosphate nanoscaffold in skin regeneration. Cell Biol. Int. 43 , 1365–1378 (2019).

Yang, J. A., Chung, H. M., Won, C. H. & Sung, J. H. Potential application of adipose-derived stem cells and their secretory factors to skin: discussion from both clinical and industrial viewpoints. Expert Opin. Biol. Ther. 10 , 495–503 (2010).

Zhou, Y. et al. Combined topical and systemic administration with human adipose-derived mesenchymal stem cells (hADSC) and hADSC-derived exosomes markedly promoted cutaneous wound healing and regeneration. Stem Cell Res. Ther. 12 , 257 (2021).

Arjmand, B. et al. Regenerative medicine for the treatment of ischemic heart disease; status and future perspectives. Front. Cell Dev. Biol. 9 , 704903 (2021).

Denning, C. et al. Cardiomyocytes from human pluripotent stem cells: from laboratory curiosity to industrial biomedical platform. Biochim Biophys. Acta 1863 , 1728–1748 (2016).

Wu, R., Hu, X. & Wang, J. Concise review: optimized strategies for stem cell-based therapy in myocardial repair: clinical translatability and potential limitation. Stem Cells 36 , 482–500 (2018).

Chong, J. J. et al. Human embryonic-stem-cell-derived cardiomyocytes regenerate non-human primate hearts. Nature 510 , 273–277 (2014).

Bagno, L., Hatzistergos, K. E., Balkan, W. & Hare, J. M. Mesenchymal stem cell-based therapy for cardiovascular disease: progress and challenges. Mol. Ther. 26 , 1610–1623 (2018).

Demurtas, J. et al. Stem cells for treatment of cardiovascular diseases: an umbrella review of randomized controlled trials. Ageing Res. Rev. 67 , 101257 (2021).

Gubert, F. et al. Mesenchymal stem cells therapies on fibrotic heart diseases. Int. J. Mol. Sci. 22 , 7447 (2021).

da Silva, J. S. et al. Mesenchymal stem cell therapy in diabetic cardiomyopathy. Cells 11 , 240 (2022).

He, X. et al. Signaling cascades in the failing heart and emerging therapeutic strategies. Signal Transduct. Target Ther. 7 , 134 (2022).

Bolli, R., Solankhi, M., Tang, X. L. & Kahlon, A. Cell therapy in patients with heart failure: a comprehensive review and emerging concepts. Cardiovasc Res. 118 , 951–976 (2022).

Bartunek, J. et al. Cardiopoietic stem cell therapy in heart failure: the C-CURE (cardiopoietic stem Cell therapy in heart failURE) multicenter randomized trial with lineage-specified biologics. J. Am. Coll. Cardiol. 61 , 2329–2338 (2013).

Bartunek, J. et al. Cardiopoietic cell therapy for advanced ischaemic heart failure: results at 39 weeks of the prospective, randomized, double blind, sham-controlled CHART-1 clinical trial. Eur. Heart J. 38 , 648–660 (2017).

Hare, J. M. et al. Comparison of allogeneic vs autologous bone marrow-derived mesenchymal stem cells delivered by transendocardial injection in patients with ischemic cardiomyopathy: the POSEIDON randomized trial. J. Am. Med. Assoc. 308 , 2369–2379 (2012).

Hare, J. M. et al. Randomized comparison of allogeneic versus autologous mesenchymal stem cells for nonischemic dilated cardiomyopathy: POSEIDON-DCM trial. J. Am. Coll. Cardiol. 69 , 526–537 (2017).

Mathiasen, A. B. et al. Bone marrow-derived mesenchymal stromal cell treatment in patients with severe ischaemic heart failure: a randomized placebo-controlled trial (MSC-HF trial). Eur. Heart J. 36 , 1744–1753 (2015).

Mathiasen, A. B. et al. Bone marrow-derived mesenchymal stromal cell treatment in patients with ischaemic heart failure: final 4-year follow-up of the MSC-HF trial. Eur. J. Heart Fail 22 , 884–892 (2020).

Florea, V. et al. Dose comparison study of allogeneic mesenchymal stem cells in patients with ischemic cardiomyopathy (The TRIDENT Study). Circ. Res. 121 , 1279–1290 (2017).

Bolli, R. et al. A Phase II study of autologous mesenchymal stromal cells and c-kit positive cardiac cells, alone or in combination, in patients with ischaemic heart failure: the CCTRN CONCERT-HF trial. Eur. J. Heart Fail 23 , 661–674 (2021).

Heldman, A. W. et al. Transendocardial mesenchymal stem cells and mononuclear bone marrow cells for ischemic cardiomyopathy: the TAC-HFT randomized trial. J. Am. Med. Assoc. 311 , 62–73 (2014).

Perin, E. C. et al. Adipose-derived regenerative cells in patients with ischemic cardiomyopathy: the PRECISE trial. Am. Heart J. 168 , 88–95 e82 (2014).

Han, S., Sun, H. M., Hwang, K. C. & Kim, S. W. Adipose-derived stromal vascular fraction cells: update on clinical utility and efficacy. Crit. Rev. Eukaryot. Gene Expr. 25 , 145–152 (2015).

Henry, T. D. et al. The Athena trials: autologous adipose-derived regenerative cells for refractory chronic myocardial ischemia with left ventricular dysfunction. Catheter Cardiovasc Inter. 89 , 169–177 (2017).

Kastrup, J. et al. Cryopreserved off-the-shelf allogeneic adipose-derived stromal cells for therapy in patients with ischemic heart disease and heart failure—a safety study. Stem Cells Transl. Med. 6 , 1963–1971 (2017).

Qayyum, A. A. et al. Adipose-derived stromal cells for treatment of patients with chronic ischemic heart disease (MyStromalCell Trial): a randomized placebo-controlled study. Stem Cells Int. 2017 , 5237063 (2017).

Qayyum, A. A. et al. Autologous adipose-derived stromal cell treatment for patients with refractory angina (MyStromalCell Trial): 3-years follow-up results. J. Transl. Med. 17 , 360 (2019).

Ngo, A. T. L. et al. Clinically relevant preservation conditions for mesenchymal stem/stromal cells derived from perinatal and adult tissue sources. J. Cell Mol. Med. 25 , 10747–10760 (2021).

Madonna, R., Cevik, C., Nasser, M. & De Caterina, R. Hepatocyte growth factor: molecular biomarker and player in cardioprotection and cardiovascular regeneration. Thromb. Haemost. 107 , 656–661 (2012).

Bartolucci, J. et al. Safety and efficacy of the intravenous infusion of umbilical cord mesenchymal stem cells in patients with heart failure: a phase 1/2 randomized controlled trial (RIMECARD Trial [randomized clinical trial of intravenous infusion umbilical cord mesenchymal stem cells on cardiopathy]). Circ. Res. 121 , 1192–1204 (2017).

Ulus, A. T. et al. Intramyocardial transplantation of umbilical cord mesenchymal stromal cells in chronic ischemic cardiomyopathy: a controlled, randomized clinical trial (HUC-HEART trial). Int. J. Stem Cells 13 , 364–376 (2020).

He, X. et al. Effect of intramyocardial grafting collagen scaffold with mesenchymal stromal cells in patients with chronic ischemic heart disease: a randomized clinical trial. JAMA Netw. Open 3 , e2016236 (2020).

Zhang, Q. et al. Signaling pathways and targeted therapy for myocardial infarction. Signal Transduct. Target Ther. 7 , 78 (2022).

Poomani, M. S. et al. Mesenchymal stem cell (MSCs) therapy for ischemic heart disease: a promising frontier. Glob. Heart 17 , 19 (2022).

Xu, W. et al. Mesenchymal stem cells from adult human bone marrow differentiate into a cardiomyocyte phenotype in vitro. Exp. Biol. Med. 229 , 623–631 (2004).

Jeong, J. O. et al. Malignant tumor formation after transplantation of short-term cultured bone marrow mesenchymal stem cells in experimental myocardial infarction and diabetic neuropathy. Circ. Res. 108 , 1340–1347 (2011).

Denu, R. A. et al. Fibroblasts and mesenchymal stromal/stem cells are phenotypically indistinguishable. Acta Haematol. 136 , 85–97 (2016).

Birbrair, A. & Frenette, P. S. Niche heterogeneity in the bone marrow. Ann. N. Y Acad. Sci. 1370 , 82–96 (2016).

Pinho, S. & Frenette, P. S. Haematopoietic stem cell activity and interactions with the niche. Nat. Rev. Mol. Cell Biol. 20 , 303–320 (2019).

Ono, N. et al. Vasculature-associated cells expressing nestin in developing bones encompass early cells in the osteoblast and endothelial lineage. Dev. Cell 29 , 330–339 (2014).

Sugiyama, T., Kohara, H., Noda, M. & Nagasawa, T. Maintenance of the hematopoietic stem cell pool by CXCL12-CXCR4 chemokine signaling in bone marrow stromal cell niches. Immunity 25 , 977–988 (2006).

Ehninger, A. & Trumpp, A. The bone marrow stem cell niche grows up: mesenchymal stem cells and macrophages move in. J. Exp. Med. 208 , 421–428 (2011).

Golan, K., Kollet, O., Markus, R. P. & Lapidot, T. Daily light and darkness onset and circadian rhythms metabolically synchronize hematopoietic stem cell differentiation and maintenance: the role of bone marrow norepinephrine, tumor necrosis factor, and melatonin cycles. Exp. Hematol. 78 , 1–10 (2019).

Cheng, X. et al. The role of SDF-1/CXCR4/CXCR7 in neuronal regeneration after cerebral ischemia. Front. Neurosci. 11 , 590 (2017).

Zou, Y. R., Kottmann, A. H., Kuroda, M., Taniuchi, I. & Littman, D. R. Function of the chemokine receptor CXCR4 in haematopoiesis and in cerebellar development. Nature 393 , 595–599 (1998).

Mao, W., Yi, X., Qin, J., Tian, M. & Jin, G. CXCL12 inhibits cortical neuron apoptosis by increasing the ratio of Bcl-2/Bax after traumatic brain injury. Int. J. Neurosci. 124 , 281–290 (2014).

Wang, Q. et al. Stromal cell-derived factor 1alpha decreases beta-amyloid deposition in Alzheimer’s disease mouse model. Brain Res. 1459 , 15–26 (2012).

Yellowley, C. CXCL12/CXCR4 signaling and other recruitment and homing pathways in fracture repair. Bonekey Rep. 2 , 300 (2013).

Li, J. et al. CXCL12 promotes spinal nerve regeneration and functional recovery after spinal cord injury. Neuroreport 32 , 450–457 (2021).

Gensel, J. C., Kigerl, K. A., Mandrekar-Colucci, S. S., Gaudet, A. D. & Popovich, P. G. Achieving CNS axon regeneration by manipulating convergent neuro-immune signaling. Cell Tissue Res. 349 , 201–213 (2012).

Matsushita, T. et al. Mesenchymal stem cells transmigrate across brain microvascular endothelial cell monolayers through transiently formed inter-endothelial gaps. Neurosci. Lett. 502 , 41–45 (2011).

Schmidt, A. et al. Mesenchymal stem cells transmigrate over the endothelial barrier. Eur. J. Cell Biol. 85 , 1179–1188 (2006).

Yarygin, K. N. et al. Cell therapy of stroke: do the intra-arterially transplanted mesenchymal stem cells cross the blood-brain barrier? Cells 10 , 2997 (2021).

Schack, L. M. et al. Expression of CD24 in human bone marrow-derived mesenchymal stromal cells is regulated by TGFbeta3 and induces a myofibroblast-like genotype. Stem Cells Int. 2016 , 1319578 (2016).

Ruster, B. et al. Mesenchymal stem cells display coordinated rolling and adhesion behavior on endothelial cells. Blood 108 , 3938–3944 (2006).

Pluchino, N. et al. CXCR4 or CXCR7 antagonists treat endometriosis by reducing bone marrow cell trafficking. J. Cell Mol. Med. 24 , 2464–2474 (2020).

Kowalski, K. et al. Stem cells migration during skeletal muscle regeneration—the role of Sdf-1/Cxcr4 and Sdf-1/Cxcr7 axis. Cell Adh. Migr. 11 , 384–398 (2017).

Liu, L. et al. From blood to the brain: can systemically transplanted mesenchymal stem cells cross the blood-brain barrier? Stem Cells Int. 2013 , 435093 (2013).

Lozito, T. P. & Tuan, R. S. Mesenchymal stem cells inhibit both endogenous and exogenous MMPs via secreted TIMPs. J. Cell Physiol. 226 , 385–396 (2011).

Lozito, T. P., Jackson, W. M., Nesti, L. J. & Tuan, R. S. Human mesenchymal stem cells generate a distinct pericellular zone of MMP activities via binding of MMPs and secretion of high levels of TIMPs. Matrix Biol. 34 , 132–143 (2014).

Menge, T. et al. Mesenchymal stem cells regulate blood-brain barrier integrity through TIMP3 release after traumatic brain injury. Sci. Transl. Med. 4 , 161ra150 (2012).

Franklin, R. J. M. & Ffrench-Constant, C. Regenerating CNS myelin—from mechanisms to experimental medicines. Nat. Rev. Neurosci. 18 , 753–769 (2017).

Brick, R. M., Sun, A. X. & Tuan, R. S. Neurotrophically induced mesenchymal progenitor cells derived from induced pluripotent stem cells enhance neuritogenesis via neurotrophin and cytokine production. Stem Cells Transl. Med. 7 , 45–58 (2018).

Zupanc, H. R. H., Alexander, P. G. & Tuan, R. S. Neurotrophic support by traumatized muscle-derived multipotent progenitor cells: role of endothelial cells and vascular endothelial growth factor-A. Stem Cell Res. Ther. 8 , 226 (2017).

Liu, Y. & Olsen, B. R. Distinct VEGF functions during bone development and homeostasis. Arch. Immunol. Ther. Exp. 62 , 363–368 (2014).

Kangari, P., Talaei-Khozani, T., Razeghian-Jahromi, I. & Razmkhah, M. Mesenchymal stem cells: amazing remedies for bone and cartilage defects. Stem Cell Res. Ther. 11 , 492 (2020).

Liu, Y. et al. Intracellular VEGF regulates the balance between osteoblast and adipocyte differentiation. J. Clin. Investig. 122 , 3101–3113 (2012).

Berendsen, A. D. & Olsen, B. R. How vascular endothelial growth factor-A (VEGF) regulates differentiation of mesenchymal stem cells. J. Histochem Cytochem. 62 , 103–108 (2014).

Garcia, K. O. et al. Therapeutic effects of the transplantation of VEGF overexpressing bone marrow mesenchymal stem cells in the hippocampus of murine model of Alzheimer’s disease. Front. Aging Neurosci. 6 , 30 (2014).

Hohman, T. J., Bell, S. P. & Jefferson, A. L., Alzheimer’s Disease Neuroimaging, I. The role of vascular endothelial growth factor in neurodegeneration and cognitive decline: exploring interactions with biomarkers of Alzheimer disease. JAMA Neurol. 72 , 520–529 (2015).

Zhang, W. et al. Neuroprotective effects of SOX5 against ischemic stroke by regulating VEGF/PI3K/AKT pathway. Gene 767 , 145148 (2021).

Jin, K. et al. Vascular endothelial growth factor (VEGF) stimulates neurogenesis in vitro and in vivo. Proc. Natl Acad. Sci. USA 99 , 11946–11950 (2002).

Bao, X. J. et al. Transplantation of Flk-1+ human bone marrow-derived mesenchymal stem cells promotes behavioral recovery and anti-inflammatory and angiogenesis effects in an intracerebral hemorrhage rat model. Int. J. Mol. Med. 31 , 1087–1096 (2013).

Bao, X. et al. Transplantation of Flk-1+ human bone marrow-derived mesenchymal stem cells promotes angiogenesis and neurogenesis after cerebral ischemia in rats. Eur. J. Neurosci. 34 , 87–98 (2011).

Pelletier, J. et al. VEGF-A promotes both pro-angiogenic and neurotrophic capacities for nerve recovery after compressive neuropathy in rats. Mol. Neurobiol. 51 , 240–251 (2015).

Hobson, M. I., Green, C. J. & Terenghi, G. VEGF enhances intraneural angiogenesis and improves nerve regeneration after axotomy. J. Anat. 197 (Pt 4), 591–605 (2000).

Hayakawa, K. et al. Vascular endothelial growth factor regulates the migration of oligodendrocyte precursor cells. J. Neurosci. 31 , 10666–10670 (2011).

Pei, G., Xu, L., Huang, W. & Yin, J. The protective role of microRNA-133b in restricting hippocampal neurons apoptosis and inflammatory injury in rats with depression by suppressing CTGF. Int. Immunopharmacol. 78 , 106076 (2020).

Xu, H. et al. Mesenchymal stem cell-derived exosomal microRNA-133b suppresses glioma progression via Wnt/beta-catenin signaling pathway by targeting EZH2. Stem Cell Res. Ther. 10 , 381 (2019).

Xin, H. et al. MiR-133b promotes neural plasticity and functional recovery after treatment of stroke with multipotent mesenchymal stromal cells in rats via transfer of exosome-enriched extracellular particles. Stem Cells 31 , 2737–2746 (2013).

Kigerl, K. A. et al. Identification of two distinct macrophage subsets with divergent effects causing either neurotoxicity or regeneration in the injured mouse spinal cord. J. Neurosci. 29 , 13435–13444 (2009).

Knoller, N. et al. Clinical experience using incubated autologous macrophages as a treatment for complete spinal cord injury: phase I study results. J. Neurosurg. Spine 3 , 173–181 (2005).

Yagura, K. et al. The enhancement of CCL2 and CCL5 by human bone marrow-derived mesenchymal stem/stromal cells might contribute to inflammatory suppression and axonal extension after spinal cord injury. PLoS ONE 15 , e0230080 (2020).

Zhong, Z. et al. Bone marrow mesenchymal stem cells upregulate PI3K/AKT pathway and down-regulate NF-kappaB pathway by secreting glial cell-derived neurotrophic factors to regulate microglial polarization and alleviate deafferentation pain in rats. Neurobiol. Dis. 143 , 104945 (2020).

Zhong, Z. et al. Adipose-derived stem cells modulate BV2 microglial M1/M2 polarization by producing GDNF. Stem Cells Dev. 29 , 714–727 (2020).

Dong, B. et al. Exosomes from human umbilical cord mesenchymal stem cells attenuate the inflammation of severe steroid-resistant asthma by reshaping macrophage polarization. Stem Cell Res. Ther. 12 , 204 (2021).

Li, X. et al. Umbilical cord tissue-derived mesenchymal stem cells induce T lymphocyte apoptosis and cell cycle arrest by expression of indoleamine 2, 3-dioxygenase. Stem Cells Int. 2016 , 7495135 (2016).

Wang, A. Y. L. et al. Human Wharton’s jelly mesenchymal stem cell-mediated sciatic nerve recovery is associated with the upregulation of regulatory T cells. Int. J. Mol. Sci. 21 , 6310 (2020).

Noone, C., Kihm, A., English, K., O’Dea, S. & Mahon, B. P. IFN-gamma stimulated human umbilical-tissue-derived cells potently suppress NK activation and resist NK-mediated cytotoxicity in vitro. Stem Cells Dev. 22 , 3003–3014 (2013).

Li, X. et al. Immunomodulatory effects of mesenchymal stem cells in peripheral nerve injury. Stem Cell Res. Ther. 13 , 18 (2022).

Shang, Y., Guan, H. & Zhou, F. Biological characteristics of umbilical cord mesenchymal stem cells and its therapeutic potential for hematological disorders. Front. Cell Dev. Biol. 9 , 570179 (2021).

Mennan, C. et al. Isolation and characterisation of mesenchymal stem cells from different regions of the human umbilical cord. Biomed. Res. Int. 2013 , 916136 (2013).

D’Addio, F. et al. The link between the PDL1 costimulatory pathway and Th17 in fetomaternal tolerance. J. Immunol. 187 , 4530–4541 (2011).

Amable, P. R., Teixeira, M. V., Carias, R. B., Granjeiro, J. M. & Borojevic, R. Protein synthesis and secretion in human mesenchymal cells derived from bone marrow, adipose tissue and Wharton’s jelly. Stem Cell Res. Ther. 5 , 53 (2014).

de Witte, S. F. H. et al. Immunomodulation by therapeutic mesenchymal stromal cells (MSC) is triggered through phagocytosis of MSC by monocytic cells. Stem Cells 36 , 602–615 (2018).

Li, Y. et al. Cell-cell contact with proinflammatory macrophages enhances the immunotherapeutic effect of mesenchymal stem cells in two abortion models. Cell Mol. Immunol. 16 , 908–920 (2019).

De Paepe, M. E., Wong, T., Chu, S. & Mao, Q. Stromal cell-derived factor-1 (SDF-1) expression in very preterm human lungs: potential relevance for stem cell therapy for bronchopulmonary dysplasia. Exp. Lung Res. 46 , 146–156 (2020).

Wynn, R. F. et al. A small proportion of mesenchymal stem cells strongly expresses functionally active CXCR4 receptor capable of promoting migration to bone marrow. Blood 104 , 2643–2645 (2004).

Ryu, C. H. et al. Migration of human umbilical cord blood mesenchymal stem cells mediated by stromal cell-derived factor-1/CXCR4 axis via Akt, ERK, and p38 signal transduction pathways. Biochem Biophys. Res. Commun. 398 , 105–110 (2010).

Yang, C. et al. The biological changes of umbilical cord mesenchymal stem cells in inflammatory environment induced by different cytokines. Mol. Cell Biochem. 446 , 171–184 (2018).

Seedorf, G. et al. Hepatocyte growth factor as a downstream mediator of vascular endothelial growth factor-dependent preservation of growth in the developing lung. Am. J. Physiol. Lung Cell Mol. Physiol. 310 , L1098–L1110 (2016).

Chen, X. Y. et al. Therapeutic potential of human umbilical cord-derived mesenchymal stem cells in recovering from murine pulmonary emphysema under cigarette smoke exposure. Front. Med. 8 , 713824 (2021).

Katsha, A. M. et al. Paracrine factors of multipotent stromal cells ameliorate lung injury in an elastase-induced emphysema model. Mol. Ther. 19 , 196–203 (2011).

Kyurkchiev, D. et al. Secretion of immunoregulatory cytokines by mesenchymal stem cells. World J. Stem Cells 6 , 552–570 (2014).

Ren, Z. et al. Human umbilical-cord mesenchymal stem cells inhibit bacterial growth and alleviate antibiotic resistance in neonatal imipenem-resistant Pseudomonas aeruginosa infection. Innate Immun. 26 , 215–221 (2020).

Liu, J. et al. Type 2 alveolar epithelial cells differentiated from human umbilical cord mesenchymal stem cells alleviate mouse pulmonary fibrosis through beta-catenin-regulated cell apoptosis. Stem Cells Dev. 30 , 660–670 (2021).

Moodley, Y. et al. Human umbilical cord mesenchymal stem cells reduce fibrosis of bleomycin-induced lung injury. Am. J. Pathol. 175 , 303–313 (2009).

Li, D. Y., Li, R. F., Sun, D. X., Pu, D. D. & Zhang, Y. H. Mesenchymal stem cell therapy in pulmonary fibrosis: a meta-analysis of preclinical studies. Stem Cell Res. Ther. 12 , 461 (2021).

Lam, G., Zhou, Y., Wang, J. X. & Tsui, Y. P. Targeting mesenchymal stem cell therapy for severe pneumonia patients. World J. Stem Cells 13 , 139–154 (2021).

Chen, K. et al. Human umbilical cord mesenchymal stem cells hUC-MSCs exert immunosuppressive activities through a PGE2-dependent mechanism. Clin. Immunol. 135 , 448–458 (2010).

Ren, G. et al. Mesenchymal stem cell-mediated immunosuppression occurs via concerted action of chemokines and nitric oxide. Cell Stem Cell 2 , 141–150 (2008).

Loy, H. et al. Therapeutic implications of human umbilical cord mesenchymal stromal cells in attenuating influenza A(H5N1) virus-associated acute lung injury. J. Infect. Dis. 219 , 186–196 (2019).

Gazdhar, A. et al. Targeted gene transfer of hepatocyte growth factor to alveolar type II epithelial cells reduces lung fibrosis in rats. Hum. Gene Ther. 24 , 105–116 (2013).

Wang, W. et al. Therapeutic mechanisms of mesenchymal stem cells in acute respiratory distress syndrome reveal potentials for Covid-19 treatment. J. Transl. Med. 19 , 198 (2021).

Chu, K. A. et al. Reversal of bleomycin-induced rat pulmonary fibrosis by a xenograft of human umbilical mesenchymal stem cells from Wharton’s jelly. Theranostics 9 , 6646–6664 (2019).

Chen, Q. H. et al. Mesenchymal stem cells regulate the Th17/Treg cell balance partly through hepatocyte growth factor in vitro. Stem Cell Res. Ther. 11 , 91 (2020).

Li, L. et al. Human umbilical cord-derived mesenchymal stem cells downregulate inflammatory responses by shifting the Treg/Th17 profile in experimental colitis. Pharmacology 92 , 257–264 (2013).

Zheng, L., Wang, S., Yang, H. & Lyu, X. [Research progress of mesenchymal stem cells attenuating acute respiratory distress syndrome by regulating the balance of M1/M2 macrophage polarization]. Zhonghua Wei Zhong Bing. Ji Jiu Yi Xue 33 , 509–512 (2021).

Fasshauer, M. & Bluher, M. Adipokines in health and disease. Trends Pharm. Sci. 36 , 461–470 (2015).

Kershaw, E. E. & Flier, J. S. Adipose tissue as an endocrine organ. J. Clin. Endocrinol. Metab. 89 , 2548–2556 (2004).

Kurylowicz, A. & Kozniewski, K. Anti-inflammatory strategies targeting metaflammation in type 2 diabetes. Molecules 25 , 2224 (2020).

Liu, J. et al. Homing and restorative effects of bone marrow-derived mesenchymal stem cells on cisplatin injured ovaries in rats. Mol. Cells 37 , 865–872 (2014).

Jalalie, L. et al. Distribution of the CM-Dil-labeled human umbilical cord vein mesenchymal stem cells migrated to the cyclophosphamide-injured ovaries in C57BL/6 mice. Iran. Biomed. J. 23 , 200–208 (2019).

Takehara, Y. et al. The restorative effects of adipose-derived mesenchymal stem cells on damaged ovarian function. Lab. Investig. 93 , 181–193 (2013).

Richards, J. S., Ren, Y. A., Candelaria, N., Adams, J. E. & Rajkovic, A. Ovarian Follicular Theca Cell Recruitment, Differentiation, and Impact on Fertility: 2017 Update. Endocr. Rev. 39 , 1–20 (2018).

Young, J. M. & McNeilly, A. S. Theca: the forgotten cell of the ovarian follicle. Reproduction 140 , 489–504 (2010).

Trzyna, A. & Banas-Zabczyk, A. Adipose-derived stem cells secretome and its potential application in “stem cell-free therapy”. Biomolecules 11 , 878 (2021).

Ding, C. et al. Human amniotic mesenchymal stem cells improve ovarian function in natural aging through secreting hepatocyte growth factor and epidermal growth factor. Stem Cell Res. Ther. 9 , 55 (2018).

Kedenko, L. et al. Genetic polymorphisms at SIRT1 and FOXO1 are associated with carotid atherosclerosis in the SAPHIR cohort. BMC Med. Genet. 15 , 112 (2014).

Shojafar, E., Soleimani Mehranjani, M. & Shariatzadeh, S. M. A. Adipose derived mesenchymal stem cells improve the structure and function of autografted mice ovaries through reducing oxidative stress and inflammation: a stereological and biochemical analysis. Tissue Cell 56 , 23–30 (2019).

Liu, M. et al. Small extracellular vesicles derived from embryonic stem cells restore ovarian function of premature ovarian failure through PI3K/AKT signaling pathway. Stem Cell Res. Ther. 11 , 3 (2020).

Li, Z., Zhang, M., Tian, Y., Li, Q. & Huang, X. Mesenchymal stem cells in premature ovarian insufficiency: mechanisms and prospects. Front. Cell Dev. Biol. 9 , 718192 (2021).

Forghani, A. et al. Differentiation of adipose tissue-derived CD34+/CD31- cells into endothelial cells in vitro. Regen. Eng. Transl. Med 6 , 101–110 (2020).

Baer, P. C. Adipose-derived stem cells and their potential to differentiate into the epithelial lineage. Stem Cells Dev. 20 , 1805–1816 (2011).

Wang, C. et al. Differentiation of adipose-derived stem cells into contractile smooth muscle cells induced by transforming growth factor-beta1 and bone morphogenetic protein-4. Tissue Eng. Part A 16 , 1201–1213 (2010).

Chen, L. et al. Adipose-derived stem cells promote diabetic wound healing via the recruitment and differentiation of endothelial progenitor cells into endothelial cells mediated by the VEGF-PLCgamma-ERK pathway. Arch. Biochem Biophys. 692 , 108531 (2020).

Dekoninck, S. & Blanpain, C. Stem cell dynamics, migration and plasticity during wound healing. Nat. Cell Biol. 21 , 18–24 (2019).

Qian, L., Pi, L., Fang, B. R. & Meng, X. X. Adipose mesenchymal stem cell-derived exosomes accelerate skin wound healing via the lncRNA H19/miR-19b/SOX9 axis. Lab. Investig. 101 , 1254–1266 (2021).

Fujiwara, O. et al. Adipose-derived stem cells improve grafted burn wound healing by promoting wound bed blood flow. Burns Trauma 8 , tkaa009 (2020).

Chen, T. et al. Efficient and sustained IGF-1 expression in the adipose tissue-derived stem cells mediated via a lentiviral vector. J. Mol. Histol. 46 , 1–11 (2015).

Zhou, K. et al. Immunosuppression of human adipose-derived stem cells on T cell subsets via the reduction of NF-kappaB activation mediated by PD-L1/PD-1 and Gal-9/TIM-3 pathways. Stem Cells Dev. 27 , 1191–1202 (2018).

Moll, G. et al. Intravascular mesenchymal stromal/stem cell therapy product diversification: time for new clinical guidelines. Trends Mol. Med. 25 , 149–163 (2019).

He, X. et al. Spontaneous apoptosis of cells in therapeutic stem cell preparation exert immunomodulatory effects through release of phosphatidylserine. Signal Transduct. Target Ther. 6 , 270 (2021).

Lukomska, B. et al. Challenges and controversies in human mesenchymal stem cell therapy. Stem Cells Int. 2019 , 9628536 (2019).

Li, C., Zhao, H. & Wang, B. Challenges for mesenchymal stem cell-based therapy for COVID-19. Drug Des. Devel Ther. 14 , 3995–4001 (2020).

Nguyen Thanh, L. et al. Outcomes of bone marrow mononuclear cell transplantation combined with interventional education for autism spectrum disorder. Stem Cells Transl. Med. 10 , 14–26 (2020).

Nguyen Thanh, L. et al. Can autologous adipose-derived mesenchymal stem cell transplantation improve sexual function in people with sexual functional deficiency? Stem Cell Rev. Rep. 17 , 2153–2163 (2021).

Caplan, A. I. Mesenchymal stem cells: time to change the name! Stem Cells Transl. Med. 6 , 1445–1451 (2017).

de Windt, T. S., Vonk, L. A. & Saris, D. B. F. Response to: Mesenchymal stem cells: time to change the name! Stem Cells Transl. Med. 6 , 1747–1748 (2017).

Boregowda, S. V., Booker, C. N. & Phinney, D. G. Mesenchymal stem cells: the moniker fits the science. Stem Cells 36 , 7–10 (2018).

Masterson, C. & O’Toole, D. The mesenchymal stromal cell magic bullet finds yet another target. Stem Cell Res. Ther. 5 , 82 (2014).

Murray, I. R. et al. Rogue stem cell clinics. Bone Jt. J. 102-B , 148–154 (2020).

Lyons, S., Salgaonkar, S. & Flaherty, G. T. International stem cell tourism: a critical literature review and evidence-based recommendations. Int. Health 14 , 132–141 (2022).

He, C. et al. Mesenchymal stem cell-based treatment in autoimmune liver diseases: underlying roles, advantages and challenges. Ther. Adv. Chronic Dis. 12 , 2040622321993442 (2021).

Bertheuil, N. et al. Adipose mesenchymal stromal cells: definition, immunomodulatory properties, mechanical isolation and interest for plastic surgery. Ann. Chir. Plast. Esthet. 64 , 1–10 (2019).

Chen, Y., Yu, Q., Hu, Y. & Shi, Y. Current research and use of mesenchymal stem cells in the therapy of autoimmune diseases. Curr. Stem Cell Res. Ther. 14 , 579–582 (2019).

Han, Y. et al. The secretion profile of mesenchymal stem cells and potential applications in treating human diseases. Signal Transduct. Target Ther. 7 , 92 (2022).

Rahmani, A. et al. Mesenchymal stem cell-derived extracellular vesicle-based therapies protect against coupled degeneration of the central nervous and vascular systems in stroke. Ageing Res. Rev. 62 , 101106 (2020).

Zhou, W. et al. Single-cell profiles and clinically useful properties of human mesenchymal stem cells of adipose and bone marrow origin. Am. J. Sports Med. 47 , 1722–1733 (2019).

Pachler, K. et al. A good manufacturing practice-grade standard protocol for exclusively human mesenchymal stromal cell-derived extracellular vesicles. Cytotherapy 19 , 458–472 (2017).

Borger, V., Staubach, S., Dittrich, R., Stambouli, O. & Giebel, B. Scaled isolation of mesenchymal stem/stromal cell-derived extracellular vesicles. Curr. Protoc. Stem Cell Biol. 55 , e128 (2020).

Nikfarjam, S., Rezaie, J., Zolbanin, N. M. & Jafari, R. Mesenchymal stem cell derived-exosomes: a modern approach in translational medicine. J. Transl. Med. 18 , 449 (2020).

Monticelli, S. & Natoli, G. Short-term memory of danger signals and environmental stimuli in immune cells. Nat. Immunol. 14 , 777–784 (2013).

Venkatesha, S. et al. Soluble endoglin contributes to the pathogenesis of preeclampsia. Nat. Med. 12 , 642–649 (2006).

Bernardo, M. E. & Fibbe, W. E. Mesenchymal stromal cells: sensors and switchers of inflammation. Cell Stem Cell 13 , 392–402 (2013).

Liu, G. Y. et al. Short-term memory of danger signals or environmental stimuli in mesenchymal stem cells: implications for therapeutic potential. Cell Mol. Immunol. 13 , 369–378 (2016).

Diez-Tejedor, E. et al. Reparative therapy for acute ischemic stroke with allogeneic mesenchymal stem cells from adipose tissue: a safety assessment: a phase II randomized, double-blind, placebo-controlled, single-center, pilot clinical trial. J. Stroke Cerebrovasc. Dis. 23 , 2694–2700 (2014).

Laskowitz, D. T. et al. Allogeneic umbilical cord blood infusion for adults with ischemic stroke: clinical outcomes from a phase I safety study. Stem Cells Transl. Med. 7 , 521–529 (2018).

Jeon, S. R. et al. Treatment of spinal cord injury with bone marrow-derived, cultured autologous mesenchymal stem cells. Tissue Eng. Regenerative Med. 7 , 316–322 (2010).

Park, J. H. et al. Long-term results of spinal cord injury therapy using mesenchymal stem cells derived from bone marrow in humans. Neurosurgery 70 , 1238–1247 (2012).

Saito, F. et al. Administration of cultured autologous bone marrow stromal cells into cerebrospinal fluid in spinal injury patients: a pilot study. Restor. Neurol. Neurosci. 30 , 127–136 (2012).

El-Kheir, W. A. et al. Autologous bone marrow-derived cell therapy combined with physical therapy induces functional improvement in chronic spinal cord injury patients. Cell Transpl. 23 , 729–745 (2014).

Karamouzian, S., Nematollahi-Mahani, S. N., Nakhaee, N. & Eskandary, H. Clinical safety and primary efficacy of bone marrow mesenchymal cell transplantation in subacute spinal cord injured patients. Clin. Neurol. Neurosurg. 114 , 935–939 (2012).

Pal, R. et al. Ex vivo-expanded autologous bone marrow-derived mesenchymal stromal cells in human spinal cord injury/paraplegia: a pilot clinical study. Cytotherapy 11 , 897–911 (2009).

Mendonca, M. V. et al. Safety and neurological assessments after autologous transplantation of bone marrow mesenchymal stem cells in subjects with chronic spinal cord injury. Stem Cell Res. Ther. 5 , 126 (2014).

Vaquero, J. et al. Intrathecal administration of autologous mesenchymal stromal cells for spinal cord injury: safety and efficacy of the 100/3 guideline. Cytotherapy 20 , 806–819 (2018).

Dai, G. et al. Transplantation of autologous bone marrow mesenchymal stem cells in the treatment of complete and chronic cervical spinal cord injury. Brain Res. 1533 , 73–79 (2013).

Jiang, P. C. et al. A clinical trial report of autologous bone marrow-derived mesenchymal stem cell transplantation in patients with spinal cord injury. Exp. Ther. Med. 6 , 140–146 (2013).

Jarocha, D., Milczarek, O., Wedrychowicz, A., Kwiatkowski, S. & Majka, M. Continuous improvement after multiple mesenchymal stem cell transplantations in a patient with complete spinal cord injury. Cell Transpl. 24 , 661–672 (2015).

Huang, L. et al. A randomized, placebo-controlled trial of human umbilical cord blood mesenchymal stem cell infusion for children with cerebral palsy. Cell Transpl. 27 , 325–334 (2018).

Karussis, D. et al. Safety and immunological effects of mesenchymal stem cell transplantation in patients with multiple sclerosis and amyotrophic lateral sclerosis. Arch. Neurol. 67 , 1187–1194 (2010).

Yamout, B. et al. Bone marrow mesenchymal stem cell transplantation in patients with multiple sclerosis: a pilot study. J. Neuroimmunol. 227 , 185–189 (2010).

Mohajeri, M., Farazmand, A., Mohyeddin Bonab, M., Nikbin, B. & Minagar, A. FOXP3 gene expression in multiple sclerosis patients pre- and post mesenchymal stem cell therapy. Iran. J. Allergy Asthma Immunol. 10 , 155–161 (2011).

Odinak, M. M. et al. [Transplantation of mesenchymal stem cells in multiple sclerosis]. Zh . Nevrol. Psikhiatr Im. S S Korsakova 111 , 72–76 (2011).

Bonab, M. M. et al. Autologous mesenchymal stem cell therapy in progressive multiple sclerosis: an open label study. Curr. Stem Cell Res Ther. 7 , 407–414 (2012).

Mohyeddin Bonab, M. et al. Evaluation of cytokines in multiple sclerosis patients treated with mesenchymal stem cells. Arch. Med Res. 44 , 266–272 (2013).

Llufriu, S. et al. Randomized placebo-controlled phase II trial of autologous mesenchymal stem cells in multiple sclerosis. PLoS ONE 9 , e113936 (2014).

Harris, V. K., Vyshkina, T. & Sadiq, S. A. Clinical safety of intrathecal administration of mesenchymal stromal cell-derived neural progenitors in multiple sclerosis. Cytotherapy 18 , 1476–1482 (2016).

Dahbour, S. et al. Mesenchymal stem cells and conditioned media in the treatment of multiple sclerosis patients: clinical, ophthalmological and radiological assessments of safety and efficacy. CNS Neurosci. Ther. 23 , 866–874 (2017).

Meng, M. et al. Umbilical cord mesenchymal stem cell transplantation in the treatment of multiple sclerosis. Am. J. Transl. Res. 10 , 212–223 (2018).

Fernandez, O. et al. Adipose-derived mesenchymal stem cells (AdMSC) for the treatment of secondary-progressive multiple sclerosis: a triple blinded, placebo controlled, randomized phase I/II safety and feasibility study. PLoS ONE 13 , e0195891 (2018).

Alvarez-Fuente, M. et al. Off-label mesenchymal stromal cell treatment in two infants with severe bronchopulmonary dysplasia: clinical course and biomarkers profile. Cytotherapy 20 , 1337–1344 (2018).

Edessy, M. et al. Autologous stem cells therapy, The first baby of idiopathic premature ovarian failure. Acta Med. Int. 3 , 19–23 (2016).

Gabr, H., Elkheir, W. & El-Gazzar, A. Autologous stem cell transplantation in patients with idiopathic premature ovarian failure. J. Tissue Sci. Eng. 7 , 27 (2016).

Bakhtiary, M. et al. Comparison of transplantation of bone marrow stromal cells (BMSC) and stem cell mobilization by granulocyte colony stimulating factor after traumatic brain injury in rat. Iran. Biomed. J. 14 , 142–149 (2010).

Zhou, Z. et al. Comparison of mesenchymal stromal cells from human bone marrow and adipose tissue for the treatment of spinal cord injury. Cytotherapy 15 , 434–448 (2013).

Yousefifard, M. et al. Human bone marrow-derived and umbilical cord-derived mesenchymal stem cells for alleviating neuropathic pain in a spinal cord injury model. Stem Cell Res. Ther. 7 , 36 (2016).

Takahashi, A. et al. Comparison of mesenchymal stromal cells isolated from murine adipose tissue and bone marrow in the treatment of spinal cord injury. Cell Transpl. 27 , 1126–1139 (2018).

Hao, T. et al. Comparison of bone marrow-vs. adipose tissue-derived mesenchymal stem cells for attenuating liver fibrosis. Exp. Ther. Med. 14 , 5956–5964 (2017).

Zare, H., Jamshidi, S., Dehghan, M. M., Saheli, M. & Piryaei, A. Bone marrow or adipose tissue mesenchymal stem cells: Comparison of the therapeutic potentials in mice model of acute liver failure. J. Cell Biochem 119 , 5834–5842 (2018).

Arminan, A. et al. Mesenchymal stem cells provide better results than hematopoietic precursors for the treatment of myocardial infarction. J. Am. Coll. Cardiol. 55 , 2244–2253 (2010).

Gaebel, R. et al. Cell origin of human mesenchymal stem cells determines a different healing performance in cardiac regeneration. PLoS ONE 6 , e15652 (2011).

Dayan, V. et al. Mesenchymal stromal cells mediate a switch to alternatively activated monocytes/macrophages after acute myocardial infarction. Basic Res. Cardiol. 106 , 1299–1310 (2011).

Lopez, Y. et al. Wharton’s jelly or bone marrow mesenchymal stromal cells improve cardiac function following myocardial infarction for more than 32 weeks in a rat model: a preliminary report. Curr. Stem Cell Res. Ther. 8 , 46–59 (2013).

Rasmussen, J. G. et al. Comparison of human adipose-derived stem cells and bone marrow-derived stem cells in a myocardial infarction model. Cell Transpl. 23 , 195–206 (2014).

Abd Emami, B. et al. Mechanical and chemical predifferentiation of mesenchymal stem cells into cardiomyocytes and their effectiveness on acute myocardial infarction. Artif. Organs 42 , E114–E126 (2018).

Omar, A. M., Meleis, A. E., Arfa, S. A., Zahran, N. M. & Mehanna, R. A. Comparative study of the therapeutic potential of mesenchymal stem cells derived from adipose tissue and bone marrow on acute myocardial infarction model. Oman Med. J. 34 , 534–543 (2019).

Download references

Acknowledgements

The authors would like to thank the Vingroup Scientific Research and Clinical Application Fund (grant number: PRO. 19.47) for supporting this work. All figures were created with Biorender.com. This work is supported by the Vingroup Scientific Research and Clinical Application Fund (Grant number: PRO.19.47).

Author information

Authors and affiliations.

Department of Research and Development, Vinmec Research Institute of Stem Cell and Gene Technology, Vinmec Healthcare System, Hanoi, Vietnam

Duc M. Hoang, Trung Q. Bach, Quyen T. Nguyen, Trang T. K. Phan, Giang H. Nguyen, Phuong T. T. Le, Van T. Hoang & Liem Thanh Nguyen

Department of Cellular Therapy, Vinmec High-Tech Center, Vinmec Healthcare System, Hanoi, Vietnam

Phuong T. Pham & Anh T. L. Ngo

Institute for Science & Technology in Medicine, Keele University, Keele, UK

Nicholas R. Forsyth

Department of Biology, Stanford University, Stanford, CA, USA

Michael Heke

You can also search for this author in PubMed   Google Scholar

Contributions

D.M.H.: conception and design, manuscript writing, administrative support, data analysis and interpretation, and final approval of the manuscript. P.T.P.: manuscript writing (BM- and UC-MSC sections) and data analysis and interpretation. T.Q.B.: manuscript writing (BM- and UC-MSC sections) and data analysis and interpretation. A.T.L.N.: manuscript writing (UC-MSC section), figure presentation, and data analysis and interpretation. Q.T.N., T.T.K.P., G.H.N., P.T.T.L., and V.T.H.: manuscript writing and data analysis and interpretation. N.R.F. and M.H.: manuscript writing and editing and data analysis and interpretation. L.T.N.: manuscript writing, administrative support, and final approval of the manuscript. All authors have read and approved the article.

Corresponding author

Correspondence to Duc M. Hoang .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Hoang, D.M., Pham, P.T., Bach, T.Q. et al. Stem cell-based therapy for human diseases. Sig Transduct Target Ther 7 , 272 (2022). https://doi.org/10.1038/s41392-022-01134-4

Download citation

Received : 15 March 2022

Revised : 19 July 2022

Accepted : 21 July 2022

Published : 06 August 2022

DOI : https://doi.org/10.1038/s41392-022-01134-4

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Adipose-derived stem cell-based optimization strategies for musculoskeletal regeneration: recent advances and perspectives.

  • Chenrui Yuan

Stem Cell Research & Therapy (2024)

CD317+ MSCs expanded with chemically defined media have enhanced immunological anti-inflammatory activities

  • Jianyong Xu

The remodeling of ovarian function: targeted delivery strategies for mesenchymal stem cells and their derived extracellular vesicles

  • Yinhua Song

Adipose-derived stem cells in immune-related skin disease: a review of current research and underlying mechanisms

Combined transplantation of hipsc-nsc and hmsc ameliorated neuroinflammation and promoted neuroregeneration in acute spinal cord injury.

  • Xiaofeng Du
  • Desheng Kong

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

stem cells research papers

Stem cells: past, present, and future

Affiliations.

  • 1 Department of Experimental Surgery and Biomaterials Research, Wroclaw Medical University, Bujwida 44, Wrocław, 50-345, Poland. [email protected].
  • 2 Department of Conservative Dentistry and Pedodontics, Krakowska 26, Wrocław, 50-425, Poland.
  • 3 Department of Experimental Surgery and Biomaterials Research, Wroclaw Medical University, Bujwida 44, Wrocław, 50-345, Poland.
  • PMID: 30808416
  • PMCID: PMC6390367
  • DOI: 10.1186/s13287-019-1165-5

In recent years, stem cell therapy has become a very promising and advanced scientific research topic. The development of treatment methods has evoked great expectations. This paper is a review focused on the discovery of different stem cells and the potential therapies based on these cells. The genesis of stem cells is followed by laboratory steps of controlled stem cell culturing and derivation. Quality control and teratoma formation assays are important procedures in assessing the properties of the stem cells tested. Derivation methods and the utilization of culturing media are crucial to set proper environmental conditions for controlled differentiation. Among many types of stem tissue applications, the use of graphene scaffolds and the potential of extracellular vesicle-based therapies require attention due to their versatility. The review is summarized by challenges that stem cell therapy must overcome to be accepted worldwide. A wide variety of possibilities makes this cutting edge therapy a turning point in modern medicine, providing hope for untreatable diseases.

Keywords: Differentiation; Growth media; Induced pluripotent stem cell (iPSC); Pluripotency; Stem cell derivation; Stem cells; Teratoma formation assay; Tissue banks; Tissue transplantation.

Publication types

  • Research Support, Non-U.S. Gov't
  • Cell Differentiation / genetics*
  • Cell- and Tissue-Based Therapy / trends*
  • Graphite / chemistry
  • Graphite / therapeutic use
  • Induced Pluripotent Stem Cells / transplantation*
  • Stem Cell Transplantation / classification
  • Stem Cells / classification
  • Stem Cells / cytology*
  • Tissue Scaffolds / chemistry

The Complex Terrain of Stem Cell Research

This essay about the challenges and limitations of stem cell research and therapy outlines the significant hurdles that must be overcome to fully realize the potential of stem cells in medicine. It discusses technical difficulties in controlling stem cell differentiation, the risk of immune rejection, the potential for stem cells to form tumors (tumorigenicity), and the ethical controversies surrounding the use of embryonic stem cells. Additionally, the essay examines regulatory barriers that slow the pace of research and the translation of discoveries into treatments, as well as the need for improved scalability, safety, and efficacy in stem cell-based therapies. Highlighting the complexity of transitioning from laboratory research to clinical applications, the essay underscores the multifaceted approach needed, including advances in stem cell biology, ethical dialogue, regulatory navigation, and innovation, to address these challenges and enhance the impact of stem cells in healthcare. PapersOwl offers a variety of free essay examples on the topic of Stem Cell.

How it works

Cellular genesis exploration, acclaimed as a preeminent domain in contemporary medicinal sciences, harbors the potential to redefine the therapeutic landscape for an extensive gamut of maladies and traumas. Nonetheless, the trajectory from laboratory conjecture to clinical fruition is replete with formidable impediments and constraints necessitating cautious navigation. This treatise delves into the multifaceted obstacles confronting cellular genesis research and therapeutic endeavors, encompassing technical intricacies, immunogenic repudiation quandaries, tumorigenic susceptibility, ethical disputations, regulatory encumbrances, and the overarching imperative for augmented expansibility, integrity, and efficacy in cell-derived remedial modalities.

A paramount technical quandary in cellular genesis research pertains to the arduous task of governing cell differentiation. The process of orchestrating cellular transformation into the requisite cell type for therapeutic utility is intricate and still veiled in obscurity. This indeterminacy engenders outcome variability and poses a substantive impediment to the establishment of dependable treatments. Furthermore, ensuring the sanctity and caliber of cellular colonies is imperative, as contamination or disparities can jeopardize the safety and efficacy of therapies engendered from these cellular cohorts.

Immunogenic repudiation presents another formidable hurdle. Analogous to organ grafts, cellular genesis therapies are susceptible to rejection by the recipient’s immune apparatus. Although stratagems such as autologous transplantation—employing the recipient’s own cells—can assuage this predicament, they are not universally tenable or efficacious for all pathologies. The conception of universal donor cells, less predisposed to rejection, is an arena of ongoing exploration albeit fraught with its own intricacies and ethical ramifications.

Tumorigenic susceptibility—manifesting as the proclivity of stem cells to engender neoplasms—emerges as a preoccupation, particularly concerning embryonic stem cells and induced pluripotent stem cells (iPSCs). The capability of these cells to perpetuate ad infinitum presents a dual-edged phenomenon, as unbridled proliferation can precipitate neoplastic formations. Meticulous examination and protracted studies are imperative to evaluate and mitigate this hazard, supplementing time and expense to the maturation of cellular genesis therapies.

Ethical contentions, notably pertaining to the utilization of embryonic stem cells, pose momentous predicaments to the domain. Disputations concerning the moral status of embryos and the ethical ramifications of their utilization for investigative purposes have engendered regulatory constraints in myriad nations, influencing funding accessibility and research latitude. These ethical deliberations underscore the necessity for sustained discourse and introspection within the scientific milieu and society at large.

Regulatory impediments further convolute the terrain of cellular genesis research. The regulatory gauntlet for procuring approval for clinical assays and therapies is protracted and convoluted, mandating substantive corroboration of safety and efficacy. While these regulations are indispensable for safeguarding patients, they can also impede the pace of research and the translation of breakthroughs into therapeutics.

Lastly, the expansibility of cellular production and the exigency for refined integrity and efficacy in treatments constitute overarching hurdles necessitating redress for cellular genesis therapies to attain ubiquity and cost-effectiveness. The formulation of methodologies for mass-producing cells sans compromise to quality, coupled with the assurance that cellular genesis-based therapeutics can be safely and efficaciously administered to a diverse patient cadre, constitutes pivotal aspirations for the domain.

In summation, notwithstanding the prodigious potential of cellular genesis research for propelling medical interventions, the odyssey from conception to fruition is convoluted and fraught with impediments. Navigating these obstacles mandates a multifaceted stratagem, encompassing the augmentation of our comprehension of cellular biology, innovation in technological spheres, ethical introspection and dialogue, circumnavigation of regulatory mazes, and the perpetual endeavor to ameliorate the safety, efficacy, and accessibility of cellular genesis therapies. Through collaborative and inventive endeavors, the latent potential of cellular genesis research can be fully harnessed, unveiling novel vistas in medicine and healthcare.

owl

Cite this page

The Complex Terrain of Stem Cell Research. (2024, Mar 01). Retrieved from https://papersowl.com/examples/the-complex-terrain-of-stem-cell-research/

"The Complex Terrain of Stem Cell Research." PapersOwl.com , 1 Mar 2024, https://papersowl.com/examples/the-complex-terrain-of-stem-cell-research/

PapersOwl.com. (2024). The Complex Terrain of Stem Cell Research . [Online]. Available at: https://papersowl.com/examples/the-complex-terrain-of-stem-cell-research/ [Accessed: 14 Apr. 2024]

"The Complex Terrain of Stem Cell Research." PapersOwl.com, Mar 01, 2024. Accessed April 14, 2024. https://papersowl.com/examples/the-complex-terrain-of-stem-cell-research/

"The Complex Terrain of Stem Cell Research," PapersOwl.com , 01-Mar-2024. [Online]. Available: https://papersowl.com/examples/the-complex-terrain-of-stem-cell-research/. [Accessed: 14-Apr-2024]

PapersOwl.com. (2024). The Complex Terrain of Stem Cell Research . [Online]. Available at: https://papersowl.com/examples/the-complex-terrain-of-stem-cell-research/ [Accessed: 14-Apr-2024]

Don't let plagiarism ruin your grade

Hire a writer to get a unique paper crafted to your needs.

owl

Our writers will help you fix any mistakes and get an A+!

Please check your inbox.

You can order an original essay written according to your instructions.

Trusted by over 1 million students worldwide

1. Tell Us Your Requirements

2. Pick your perfect writer

3. Get Your Paper and Pay

Hi! I'm Amy, your personal assistant!

Don't know where to start? Give me your paper requirements and I connect you to an academic expert.

short deadlines

100% Plagiarism-Free

Certified writers

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Int J Mol Sci

Logo of ijms

Stem Cell Therapy: From Idea to Clinical Practice

Regenerative medicine is a new and promising mode of therapy for patients who have limited or no other options for the treatment of their illness. Due to their pleotropic therapeutic potential through the inhibition of inflammation or apoptosis, cell recruitment, stimulation of angiogenesis, and differentiation, stem cells present a novel and effective approach to several challenging human diseases. In recent years, encouraging findings in preclinical studies have paved the way for many clinical trials using stem cells for the treatment of various diseases. The translation of these new therapeutic products from the laboratory to the market is conducted under highly defined regulations and directives provided by competent regulatory authorities. This review seeks to familiarize the reader with the process of translation from an idea to clinical practice, in the context of stem cell products. We address some required guidelines for clinical trial approval, including regulations and directives presented by the Food and Drug Administration (FDA) of the United States, as well as those of the European Medicine Agency (EMA). Moreover, we review, summarize, and discuss regenerative medicine clinical trial studies registered on the Clinicaltrials.gov website.

1. Introduction

Despite the progress in medical science, there still exist various diseases in the world for which there is no suitable treatment. People affected by incurable disorders typically use treatment methods intended to decrease the somatic and psychological symptoms and, in these situations, the physician offers treatment methods only to manage the disease, not treat it. Therefore, researchers are attempting to develop new treatment methods to not only control the symptoms of, but also to treat those diseases for which no cure is available at present.

Regenerative medicine is considered a promising new source of treatment for untreatable diseases in modern science [ 1 ]. Regenerative medicine is a multidisciplinary field including cell biology, genetic, biomechanics, material science, and computer science [ 2 , 3 ], the ultimate target of which is returning normal function to defective cells and tissues [ 4 ]. Since the discovery of stem cells and the spread of awareness regarding their unique properties, they have been defined as therapeutic agents for organ and tissue repair, and so are widely considered good candidates for regenerative medicine, due to their many potential applications [ 5 ]. Regenerative medicine is now regarded as an alternative to traditional drug-based treatments by researchers who study its potential applications in various diseases, including degenerative diseases, among others [ 6 , 7 , 8 , 9 , 10 ]. The main concept of regenerative medicine is implied tissue/organ regeneration using cells and, to reach this target, different kinds of cells have been used. However, various studies have indicated that cell therapy is restricted by a few limitations. In recent years, different alternatives have been introduced for cell therapy in order to resolve these limitations, including the improved application of stem cells for the restoration of tissue, such as the combination of cells with scaffolds, cell cultures with suitable biochemical properties, gene editing, and the immunomodulation of stem cells, as well as the use of stem cell derivatives [ 11 , 12 , 13 , 14 , 15 ]; however, the use of these alternatives clinically may be postponed, as more preclinical studies are required due to their status as newer technologies [ 16 ].

Stem cells are a group of immature cells that have the potential to build and recover every tissue/organ in the body due to their unique proliferative, differentiation, and self-renewal abilities [ 17 ]. Stem cells provide therapeutic effects which improve physical development by regenerating damaged cells to assist in organ recovery. Relying on the natural abilities of stem cells, researchers have used their biological mechanisms for stem-cell-based therapy. The mechanisms of action through which stem cells can promote the regeneration of tissue are diverse, including (1) inhibition of inflammation cascades [ 18 , 19 ], (2) reduction of apoptosis [ 20 , 21 ], (3) cell recruitment [ 22 , 23 ], (4) stimulation of angiogenesis [ 24 , 25 ], and (5) differentiation [ 26 ]. The cause of a disease is a vital consideration in selecting the proper stem cell mechanism and in the regeneration of tissue/organs using stem cells. Many examinations must be carried out to determine the main mechanisms involved in treatment when these cells are to be used in clinical practice, and the convergence of stem cell therapeutic mechanisms and disease mechanisms is expected to increase the chance of developing cures through stem cell applications.

From 1971 to 2021, 40,183 research papers were published regarding stem-cell-based therapies. All of these studies were conducted around discoveries and for the goal of “Stem Cell Therapy” based on the therapeutic efficacy of stem cells [ 27 ]. As basic stem cell research has soared over the past few years, “translation research”, a relatively new field of research, has recently greatly developed, making use of basic research results to develop new treatments. Although many articles on stem-cell-based therapies are published annually and their number increases every year, the number of clinical trial studies has not increased rapidly. Furthermore, among these studies, only a small portion of them can receive full regulatory approval for verification as treatment methods. Although one reason for this difference is due to the need for various prerequisite preclinical studies before carrying out a clinical trial study, the main reason is due to the sharply defined guidelines which prevent the translation of many preclinical studies to clinical trials.

In this review, we provide a general overview regarding the translation of stem cell therapies from idea to clinical service. Understanding the step-by-step knowledge underlying the translation of ideas to medical services is the first step in introducing a new treatment method. In this review, we divide this pathway into four levels, including idea evaluation, preclinical studies, clinical trial studies, and clinical practice. We focus not only on understanding each level’s requirements, but also discuss how an idea is assessed during the transition from one level to the next and, finally, move on to marketing.

2. From Idea to Preclinical Study

If a researcher has an idea regarding regenerative medicine using stem cells that inspires their use in a study, it must first be evaluated. During the evaluation step, it is important to select the target disease and make sure that the mechanism causing the disease is understood. Disease-related mechanisms refer to the cellular and molecular processes by which a particular disorder is caused [ 28 , 29 ], and stem-cell-based therapies are considered a treatment method intended to compensate for the disruption caused by such mechanisms in order to finally restore the defective tissue. Multiple mechanisms cause diseases [ 30 , 31 , 32 ]; however, stem cells, with their tremendous differentiation, self-renewal, angiogenesis, anti-inflammation, anti-apoptotic, and immunomodulatory potentials, as well as their capacity for induction of growth factor secretion and cell signaling, can affect these mechanisms [ 33 , 34 , 35 , 36 , 37 ].

After subject evaluation, preclinical studies should be carried out to determine whether the idea has any potential to treat the disease, and the safety of the final product should be assessed in an animal model of the target disease [ 38 , 39 , 40 ]. Preclinical studies are composed of in vitro and in vivo studies. In vitro experiments are performed with biological molecules and cells based on various hypotheses and, during the in vitro evaluation, a new treatment method is assayed in this controlled environment [ 39 ]. In contrast, during in vivo studies, as controlling all biological entities is impossible, the new product may be affected by various factors and thus present different effects. The general purpose of a preclinical study is to present scientific evidence supporting the performance of a clinical study, and the following are required for a decision to move forward to clinical study: (i) the feasibility and establishment of the rationale (e.g., validation, separation of active ingredients in vitro, and determination of its mechanism in vivo), (ii) establishment of a pharmacologically effective capacity (e.g., secure initial dose verification), (iii) optimization of administration route and usage (e.g., safe administration method, repeated administration, and interval verification), (iv) identification and verification of the potential activity and toxicity (e.g., toxicity analysis according to single and repetitive testing), (v) identification of the potential for special toxicity (e.g., genetic, carcinogenic, immunological, and neurotoxic analyses), and (vi) determination of whether to continue or discontinue development of the treatment [ 41 , 42 ].

3. From Preclinical Study to Clinical Trial

In principle, any idea regarding stem cell therapy should be assessed using comprehensive studies (i.e., in vitro and in vivo) before a clinical trial is considered, and the results of these studies should be proved by competent authorities. It can be easy during an in vitro study to create manipulative biological environments such as through the use of genetic mutation, drug testing, and pharmaceuticals, and it is easy to observe changes through the application of manipulated variables through living cells [ 43 , 44 , 45 ]. However, given the many associated variables, such as molecular transport through circulating blood and organ interactions, it is hard to say whether such a study can completely mimic the in vivo environment [ 43 , 44 , 45 ]. Before application in patients, in vivo experiments are conducted after in vitro experiments in order to overcome these weaknesses.

Many researchers use rodents for in vivo studies, due to their anatomical, physiological, and genetic similarities to humans, as well as their other unique advantages including small size, ease of maintenance, short life cycle, and abundant genetic resources [ 46 ]. The strength of in vivo studies is that they can supplement the limitations of in vitro studies, and the outcomes of their applications can be inferred in humans through the use of human-like biological environments. To establish in vivo experiments for stem cell therapies, the most correlated animal model should be selected depending on the specific safety aspects to be evaluated. Where possible, cell-derived drugs made for humans should be used for proof-of-concept and safety studies [ 47 ]. Homogeneous animal models can also be utilized as the most correlated systems in proof-of-concept studies [ 48 ].

Furthermore, in vivo studies require ethical responsibilities and obligations to be upheld according to experimental animal ethics. In other words, unnecessary and unethical experiments must be avoided. Summing up the above, we can see that both in vitro and in vivo approaches are used in preclinical studies, which should be carried out before clinical trial applications based on various interests.

Several factors must be considered in different in vitro and in vivo studies, including cell type determination, cell dose specification, route of administration, and safety and efficiency.

3.1. Stem Cell Source Determination

As expectations rise for regenerative treatment through the application of stem cell therapies, the number of applications of various types and stem cell sources has increased, and stem cell therapies have diversified from autologous to allogenic to iPSCs. These stem cell treatments can vary in risk, depending on the cell manufacturing process [ 49 ], among other factors, and in clinical experience, such that all types of stem cell treatments must be evaluated on the same basis [ 50 ]. Therefore, the strengths and weaknesses of each type of stem cell should be identified in order to determine the maximum therapeutic effect of stem cells in various diseases. This will enable us to build disease-targeted stem cells by applying the appropriate stem cells to the appropriate diseases. Below, we briefly discuss the characteristics of various stem cells.

3.1.1. Mesenchymal Stem Cells (MSCs)

MSCs are lineage-committed cells that divide into mesenchymal systems, primarily fatty cells, chondrocytes, and osteocytes [ 51 ]. It is well known that MSCs can be differentiated into dry cells, nerve cells, glioma cells, and skeletal muscle cells under proper in vitro culture conditions [ 52 , 53 , 54 , 55 , 56 , 57 ]. MSCs are primarily derived from myeloid and adipose tissues [ 58 , 59 ]. At present, MSCs are also isolated from many other tissues, such as the retina, liver, gastric mucosa, tendon, cartilage, placenta, cord blood, and blood [ 60 , 61 , 62 , 63 ]. The biggest characteristics of MSCs are their immunosuppressive functions, which prevent the proliferation of activated T cells through immunosuppressive cytokine secretion and suppression of programmed cell death signaling [ 64 , 65 ]. Due to this role, they have been spotlighted as a potential treatment for immune-related inflammation and disease. The initial clinical application of MSCs was in a case of patients with severe graft versus host disease (GVHD), and these cells have since been well applied in clinical practice, as evidenced through various studies [ 66 , 67 , 68 ].

MSCs have a variety of characteristics according to their organ of origin [ 69 ]. BM-MSCs, which are isolated from bone marrow, are useable in both autologous and allogenic contexts, and can perform stromal functions. However, the process of cell isolation from bone marrow is not only accompanied by the risk of pain and infection, but also has a lower efficiency of collection than other MSC sources. Furthermore, these cells have a longer doubling time (DT) in comparison to MSCs derived from other sources (approximately 60 h) [ 70 ]. Compared to BM-MSCs, AD-MSCs are not only easy to collect, but are also 100 to 500 times more efficient to harvest and have a shorter DT (approximately 20 h) [ 71 ]. However, these are adipose-derived stem cells that have a strong characteristic of adipogenic differentiation, such that they can be suggested as a valid alternative to BM-MSCs, but their nature must be considered regarding proper culture and body environment. Furthermore, there are concerns that these factors may affect the efficacy of treatment, as the amount of cytokines secreted is significantly lower when compared to BM-MSCs [ 72 ]. MSCs extracted from the umbilical cord (UC-MSCs) have come into the spotlight to compensate for these issues: UC-MSCs not only have the advantage of being easily collected compared to other stem cells, but also avoid ethical or donor age issues. They have superior proliferation and differentiation capabilities compared to BM-MSCs and AD-MSCs, and their DT has been reported as 24 h [ 69 , 73 ]. UC-MSCs are currently a subject of concern, as although they are easy to store frozen for a long time (e.g., in a cord blood bank), the cell survival rate and success rate during extraction are not high, due to exposure to cryogenic protectors during cryogenic storage [ 73 ]. Furthermore, as the cells are isolated from other organs, they have limited self-renewal capacity, and their senescence is faster than in other stem cells in long-term cultivation [ 66 , 74 ].

3.1.2. Hematopoietic Stem Cells (HSCs)

HSCs can be differentiated into cells from all hematopoietic systems present in the bone marrow and chest glands, namely myeloid cells and lymphocytes. HSCs can be obtained at good levels from adult bone marrow, the placenta, and cord blood. They can cause immunological problems such as transplant rejection. Nevertheless, they have been shown to be an effective treatment method in various diseases, including leukemia, malignant lymphoma, and regenerative anemia, as well as congenital metabolism, congenital immunodeficiency, nonresponsive autoimmune disease, and solid cancer to date. Furthermore, HSCs are the only stem cell type approved for stem cell treatment by the Food and Drug Administration (FDA) [ 75 , 76 ].

3.1.3. Embryonic Stem Cells (ESCs)

ESCs have established cell lines that can be maintained through in vitro culture. They are pluripotent cells that can be differentiated into almost any type of cell present in the body, and can be differentiated in vitro by adding external factors to the culture medium or by genetic modification. However, they may form teratomas, which are composed of various forms of cells derived from the endoderm, mesoderm, and exoderm, when transplanted into an acceptable host [ 77 ].

3.1.4. Induced Pluripotent Stem Cells (iPSCs)

iPSCs are artificially created stem cells. These cells are made by reprogramming adult somatic cells such as fibroblast cells. They share many of the characteristics of ESCs, including self-renewability, pluripotent differentiation, and malformed species performance. Unfortunately, these cells have little scientific evidence regarding changes in cell-specific regulatory pathways, gene expression, and epigenetic regulation. These characteristics pose a risk of tissue chimerism or cell dysfunction [ 78 ].

In summary, although the FDA-approved stem cell type is HSCs from healthy donors, a variety of issues have been raised, including a lack of donors and immune rejection. Therefore, we need to understand the characteristics of stem cells in order to handle them accordingly and overcome their disadvantages while maximizing their advantages. As stem cells derived from various sources have different characteristics, capabilities, potential, and efficiency, selecting the right source of stem cells that is appropriate for the target can be effective in assuring treatment efficiency.

3.2. Cell Dose Specification

The effective range of administration (i.e., dosage) of stem cells or stem-cell-derived products used in treatment should be determined through in vivo and in vitro studies. The safe and effective treatment capacity must be identified and, where possible, the minimum effective capacity must also be determined. When administered to vulnerable areas such as the central nervous system and myocardium, it has been reported that conducting normal dosage determination tests is unlikely. Thus, if the results of nonclinical studies can safety demonstrate treatment validity, it may be appropriate to conduct early human clinical trials with doses that may indicate therapeutic effects [ 79 ].

Will a high cell dose have better effects, considering only the effectiveness of stem cells? We answer this question below. An increasing dose of CD34 + cells (0.5 × 10 5 per mouse) has been shown to have positive effects, stimulating multilineage hematopoiesis at early stages and increasing the magnitude of reconstitution at post-transplant stages. Furthermore, improved T-cell reconstitution was correlated with higher cell doses of stem cells, compared to lower cell doses [ 80 ]. However, a few studies related to acute myeloblastic leukemia (AML) have reported that high doses of HSCs were correlated with restored function and rapid hematological and immunological recovery, but these results were not unconditional. In this study, a higher dose of HSCs (≥7 × 10 6 /kg) resulted in poorer outcomes and a higher relapse rate than the lower dose of HSCs (<1 × 10 6 /kg) [ 81 ]. In preclinical studies on heart disease, Golpanian et al. have demonstrated, through comparison of some preclinical studies for optimized cell dose, the therapeutic effects of stem cell types (i.e., allogenic and autologous MSCs), as well as the proper cell dose of stem cells and route of administration (direct epicardial and intravenous) in heart disease. Their results showed that the total number of cells used was different, but were inconsistent with the hypothesis that a higher number of cells would have higher therapeutic efficacy [ 82 ]. Therefore, these conclusions suggest that the currently reported data do not provide a decisive answer, such that sufficient and detailed early-stage studies may be needed before proceeding with clinical trials.

3.3. Route of Administration

Stem cells have been extensively studied under various disease conditions, depending on their type and characteristics. At this time, the route of administration should not be overlooked in favor of the number of stem cells transplanted. Several reports have shown that engraftment ability typically has a lower rate of reaching target organs relative to the number of transplanted cells, and does not have a temporary longer duration [ 83 , 84 ].

The methods of stem cell administration can largely be divided into local and systemic transmission. Local transmission involves specific injections through various manipulations and direct intra-organ injections, such as intraperitoneal (IP), intramuscular, and intracardiac injections. Systemic transmission uses vascular pathways, such as intravenous (IV) and intra-arterial (IA) methods. According to the publications in the literature, IV is the most common method, followed by intrasplenic and IP [ 85 , 86 , 87 ]. In a liver disease model, IV was shown to be not only suitable for targeting the liver, but also showed better liver regeneration effects than other routes of administration [ 85 , 88 ]. Intracardial injection showed better cell retention in heart disease, while intradermal injection showed better treatment in skin diseases [ 89 , 90 ]. Hence, we can determine that, in the context of these various diseases, the routes of administration should be different depending on the target organ. Many researchers have suggested that intravascular injection is a minimally invasive procedure, but it also poses a risk of clogged blood vessels, such that direct intravascular injection increases the risk of requiring open-air operations [ 91 ]. Clinical trials have reported that the number of cells and treatment efficacy under the same conditions, as in preclinical studies, are not significant, but also differ in significance depending on the route of administration [ 92 , 93 ]. Therefore, researchers should continue to study which cells are appropriate for a given route of administration—even within the same disease—based on many precedents [ 82 ]. In addition, researchers should explore the appropriate routes of administration for safer and more effective therapeutic effects.

3.4. Manipulation of Cell Transplantation for Safety and Efficiency Improvement of Administration

All medical treatments have benefits and risks. It is not particularly safe to apply these unproven stem cell treatments to patients. As expectations for regenerative treatment through stem cell therapies increase, the application of various administration pathways, including through the spinal cord, subcutaneous, and intramuscular, as well as the stem cell therapies themselves, have been diversifying, from autologous to homogenous to iPS. These stem cell treatments can vary in risk, depending on the cell type manufacturing process among other factors, and they differ in clinical experience, such that all types of stem cell treatments must be evaluate on the same basis. Furthermore, it should only be in limited and justified contexts that stem cells which can proliferate and have all-purpose differentiation remain in a final product.

Unfortunately, the only safe stem cells that have been employed in regenerative medicine so far are omnipotent stem cells, such as HSCs and MSCs, which are isolated from their self-origin [ 94 ]. Unfortunately, potential clinical applications using iPSCs and ESCs face many hurdles, as they present higher risk, including the possibility of rejection, teratoma formation, and genomic instability [ 95 ]. Hence, many researchers have attempted to overcome stem cell tracking for safety assessment. To check the engraftment and the remaining amount of stem cells, they have been labeled using BrdU, CM-Dil, and iron oxide nanoparticles, and visualized using Magnetic resonance imaging (MRI) [ 84 , 96 , 97 ].

A close analysis of the distribution patterns of administrative sites and target organs is required, as well as whether a distribution across the body is expected, and the organ that the cells are predicted to be distributed through should undergo a full-term analysis, including evaluation at administrative sites. To date, studies have reported assessments in the brain, lungs, heart, spleen, testicles, ovaries, kidneys, pancreas, bone marrow, blood, and lymph nodes, including areas of administration [ 98 ].

Some researchers have carried out the detection of transplanted UC-MSCs delivered by IV injection in the lung, heart, spleen, kidney, and liver. According to their results, the transplanted cells were not detected in other organs, except the lung and liver, for 7 days. In the lung and liver, the detected cells persisted at least 7 days after the transplant [ 99 ]. Furthermore, in a study comparing BM-MSCs and UC-MSCs in terms of cell tracking, they reported on the persistence of stem cells according to the route of administration used. In the results of the comparison of intracardiac and intravenous routes, the transplanted stem cells were detected in the lung for 10 days, but the signal disappeared after 21 days [ 100 ]. In other research, the stem cells were transplanted with using a biomaterial scaffold. The AD-MSCs were transplanted with hyaluronic acid/alginate hydrogel through intradermal injection, and could be detected by CM-Dil staining for 30 days [ 101 ]. These studies may show that the transplanted cells localized to the damaged organs through their homing ability, but the results of these previous studies seem to indicate that the residual volume and the residual date vary significantly depending on the target disease, organs, and type of stem cells. The cell residual means the survival of the cell, which represents the risk of formation of tumors. To overcome the problem of teratoma formation, the following results have been reported: According to one study, ESCs showed the following rates of teratoma formation: 100% under the kidney capsule, 60% intratesticular, 25–100% subcutaneous, and 12.5% intramuscular. To overcome this problem, the investigators performed a co-injection with Matrigel into an animal model. According to their results, subcutaneous implantation of ESCs in the presence of Matrigel appeared to be the most efficient, reproducible, and easiest approach for preventing teratoma formation, other than only using ESCs [ 102 ]. Moreover, cellular products derived from iPSCs have higher potential as potential cell sources in personalized medicine [ 103 ]. Their applicability is currently limited due to concerns regarding the potential risk of serious transplant-related side effects, such as tumor formation due to residual pluripotent cells [ 104 ]. Hence, a recent study reported the establishment of an optimized tool for therapeutic intervention that allows for controlled specific and selective ablation of iPSCs through the use of LVCAGs–transgenic iPSCs [ 104 ].

Unlike MSCs, which are generally considered immune-tolerant as an immunomodulator, transplantation of ESCs and HSCs requires close examination of the matching of histocompatibility antigen (HLA) between the donor and beneficiary [ 105 , 106 ]. Although homogeneous mesenchymal stem cells are known to have immunogenicity in immune-active rodent models and are quickly removed from the peripheral blood, studies have shown that a few MSCs remain for weeks to months. Therefore, it is recommended to conduct a study to assess the persistence of MSCs in the cell preparations administered, in order to assess the risk of stem cell removal. Therefore, for stem cell therapies that have undergone extensive in vitro manipulation such as long-term cell culture—including those derived from ESCs and iPSCs—both oncogenicity and genetic stability must be evaluated before clinical research begins. Furthermore, we must constantly review and study the latest research on safety, as well as the effects of regeneration using stem cells, and discuss and study the potential of regenerative medicine [ 107 , 108 , 109 , 110 , 111 ].

As discussed earlier, in vitro and in vivo preclinical studies are the direction of current research, and encompass the tasks that need to be completed. If we reinforce the current strengths and weaknesses based on the preceding content, we are already a step closer to developing stem cell treatments.

4. From Clinical Trial to Clinical Practice

Before a treatment is applied in humans (i.e., patients), preclinical study must involve checking whether the effect of treatment will be positive or negative and, if there are any negative effects, the researcher must check the safety possibilities at every step. Due to concerns relating to treatment using stem-cell-based products, deciding whether preclinical studies are sufficient for translating to clinical trials raises several issues that must be assessed by competent authorities. An application for a clinical trial should be submitted to the Food and Drug Administration (FDA), the European Medicine Agency (EMA), or another organization, based on the country [ 112 ].

The FDA is responsible for certifying clinical trial studies for stem-cell-based products in the United States [ 113 ]. If a new drug is introduced to a clinical investigator which has not been approved by the FDA, an Investigational New Drug (IND) application may need to be submitted [ 114 ]. The IND application includes data from animal pharmacology and toxicology studies, clinical protocols, and investigator information [ 115 ]. A lack of preclinical support (e.g., in vitro and in vivo studies) can lead to required modification or disapproval. If the FDA has announced that an IND requires modifications (meaning that the application is intended to secure approval but has not yet been approved), the results of the preclinical studies were deemed insufficient or inadequate for translation to clinical trial study, such that further study must be completed, after which an amended IND should be submitted.

The FDA has published guidelines for the submission of an IND in the Code of Federal Regulations (CFR). These regulations are presented in 21 CFR part 210, 211 (Current Good Manufacturing Practice (cGMP)), 21 CFR part 312 (Investigational New Drug Application), 21 CFR 610 (General Biological Product Standards), and 21 CFR 1271 (Human Cells, Tissues, and Cellular and Tissue-Based Products) [ 116 , 117 , 118 ]. These guidelines have been issued for the development of stem cell products with the highest standards of safety and potential effective translation to clinical trial studies.

The FDA issued 21 CFR parts 210 and 211to ensure the quality of the final products [ 119 ]. The 21 CFR part 210 contains the minimum current good manufacturing practice (cGMP) considered at the stages of manufacturing, processing, packing, or holding of a drug, while the 21 CFR part 211 contains the cGMP for producing final products. The 21 CFR 211 includes FDA guidelines for personnel, buildings and facilities, equipment, and control of components, process, packaging, labeling, holding, and so on, all of which are critical for pharmaceutical production [ 116 , 117 , 118 , 119 , 120 , 121 ]. The requirements for IND submission and conducting clinical trial studies, reviewed by the FDA in the 21 CFR part 312 (Investigational New Drug Applications), includes exemptions that are described in detail in 312.2 (general provisions). Such exemptions do not require an IND to be submitted, but other studies must present an IND based on 21 CFR part 312. The section, 21 CFR part 312, provides different information, including the requirements for an IND, its content and format, protocols, general principles of IND submission, and so on. In addition, the FDA describes the administrative actions of IND submission, the responsibilities of sponsors and investigators, and so on, in this section [ 116 , 117 , 122 ]. The 21 CFR part 610 contains general biological product standards for final product characterization. The master cell bank (MCB) or working cell bank (WCB) used as a source for stem-cell-based final products must be tested before the release or use of the product in humans. The MCB and WCB should be tested for sterility, mycoplasma, purity, identity, and potency, among other tests based on the final products (e.g., viability, stability, phenotypes), before use at the clinical level. The FDA provides all required information regarding general biological product standards in this section, including release requirements, testing requirements, labeling standards, and so on [ 116 , 117 , 123 , 124 ]. The 21 CFR part 1271 focuses on introducing the regulations for human cells, tissues, and cellular and tissue-based products (HCT/P’s), in order to ensure adequate control for preventing the transmission of communicable disease from cell/tissue products. Current Good Tissue Practice (GTP) is a part of 21 CFR part 1271, where the purpose of GTP is to present regulations for the establishment and maintenance of quality control for prevention of introduction, transmission, or spread of communicable diseases, including regulations for personnel, procedures, facilities, environmental control, equipment, and so on [ 125 , 126 , 127 , 128 ].

The EMA is an agency in the European Union (EU) which is responsible for evaluating any investigational medical products (IMPs) in order to make sure that the final product is safe and efficient for public use. When planning to introduce a new drug for a clinical trial in Europe, one may be required to submit clinical trial applications to the EMA for IMPs. Clinical trial applications for IMPs include summaries of chemical, pharmacological, and biological preclinical data (e.g., from in vivo and in vitro studies) [ 129 ]. The EMA has presented different regulations to support the development of safe and efficient products for public usage, including Regulation (EC) No. 1394/2007, Directive 2004/23/EC, Directive 2006/17/EC, Directive 2006/86/EC, Directive 2001/83/EC, Directive 2001/20/EC, and Directive 2003/94/EC.

Regulation (EC) No. 1394/2007 defines the criteria for regulation regarding ATMPs. Advanced therapy products (ATMPs) are focused on gene therapy medicinal products (GTMP), somatic cell therapy medicinal products (sCTMP), tissue-engineered products (TEP), and combined ATMPs, which refers to a combination of two different medical technologies. Regulation (EC) No. 1394/2007 includes the requirements to be used in development, manufacturing, or administration of ATMPS [ 130 , 131 , 132 ]. Directive 2004/23/EC, Directive 2006/17/EC, and Directive 2006/86/EC define standards for safety and quality, as well as technical requirements for donation, procurement, testing, preservation, storage, and distribution of tissue and cells intended for human applications [ 133 , 134 , 135 ]. Directive 2001/83/EC applies to medicinal products for human use [ 136 ]. Directive 2001/20/EC presents the regulations for the implantation of products in clinical trials in the EU [ 137 ]; however, this directive will be replaced by regulation (EU) No. 536/2014. Regulation (EU) No. 536/2014 was adapted by the European Parliament in 2014, and provides regulation for clinical trials on medical products intended for human use. The new EU regulation comes into effect on 31 January 2022 and aims to coordinate all clinical trials performed throughout the EU, using clinical trials submitted into CTIS (Clinical Trials Information System). The definition of regulation (EU) No. 536/2014 as a homogeneous regulation serves an important role in the EU, as all member states of the EU can be involved in multicenter clinical trials using international coordination, thus allowing larger patient populations [ 138 ]. Directive 2003/94/EC provides Good Manufacturing Practice (GMP) Guidelines in relation to medicinal products or IMPs intended for human use [ 139 ]. All process and application requirements for the IMP application are present in the regulations and directives of the EMA. After presenting an IND/IMP to the regulatory authority responsible for clinical trial oversight (FDA or EMA), the application will be reviewed in accordance with the FDA/EMA criteria and, if assured of the protection of humans enrolled in the clinical study, the application will be approved by the investigational review boards (IRBs) in the United States or Ethics Committees (ECs) in the European Union. Clinical trial studies are composed of different steps where, at each step, products are assessed using different quality and quantity measurements by the responsible agency. An efficient clinical trial study should address the safety and efficiency of new stem cell products in each of the different steps, and it is important to complete each step based on defined instructions and regulations, as the results of previous steps are needed to move forward.

Almost all clinical trial studies that have been approved for testing in humans have been registered online ( https://www.clinicaltrials.gov/ accessed 12 December 2021). Our search on this website revealed more than 6500 records for interventional studies registered using “Stem Cells” up to December 2021. The recorded clinical trials can be analyzed from different aspects.

Recruiting status: The recruiting status of these studies indicated that 18% of these studies were ongoing (recruitment) and 42% were completed ( Figure 1 ). Although completed, suspended, terminated, and withdrawn studies are all terms used for studies that have ended, each is used to describe a different status. Completed studies are those that have ended normally and the participants were completely enrolled in the study. Suspended, terminated, and withdrawn studies are studies that stopped early; however, the participant enrolment status differs between them. A suspended study may start again, but nobody can continue to participate in terminated or withdrawn studies [ 140 , 141 ].

An external file that holds a picture, illustration, etc.
Object name is ijms-23-02850-g001.jpg

Status of clinical trials using stem cells.

Type of disease: Stem-cell-based therapy is a new approach for the treatment of various diseases in different clinical trial studies. Blood and lymph diseases are the most common diseases that have benefited from this new approach ( Figure 2 ). Blood and lymph diseases refer to any type of disorder related to blood and lymph deficiency or abnormality, such as anemia, blood protein disorder, bone marrow disease, leukemia, hemophilia, thalassemia, thrombophilia, lymphatic disease, lymphoproliferative disease, thymoma, and so on. In addition, various clinical trial studies have been performed using stem cells to treat immune system disease; neoplasm, heart, and blood disease; and gland- and hormone-related disease ( Figure 2 ). However, this does not mean that all of these studies had great results, nor does it mean that all of these studies introduced a new treatment method; some of these clinical trial studies were only intended to increase treatment efficiency, compare different types of treatment methods, or analyze various parameters after the administration of stem cells into the body.

An external file that holds a picture, illustration, etc.
Object name is ijms-23-02850-g002.jpg

Diseases considered in clinical trials using stem cells.

Autologous vs. Allogenic: Stem-cell-based products for use in clinical trial studies can be divided into two categories: autologous and allogeneic stem cells. In autologous stem cell therapy, the stem cells are collected from the patient’s own body. Culture-expanded autologous stem cells are autologous stem cells that are expanded before transplantation, and can be divided into two groups: modified and unmodified expanded autologous stem cells. If autologous stem cells were transplanted to the donor immediately after collection, this is a nonexpanded autologous stem cell treatment. The use of these cells usually has fewer restrictions for receiving clinical trial authorization. The classification of allogenic stem cells is similar to that of autologous stem cells, except that allogeneic stem cells are collected from a healthy donor. The use of these cells requires more prerequisite tests, in order to check the donor’s health. Allogenic stem cells have been used more than autologous stem cells in the clinical trial studies (46.34% vs. 44.51%), as shown in Figure 3 .

An external file that holds a picture, illustration, etc.
Object name is ijms-23-02850-g003.jpg

Applied stem cell types in clinical trials using stem cells.

Phase: Clinical trial studies are conducted in different phases. In each phase, the purpose of study, the number of participants, and the follow-up duration may differ. A new phase of clinical trials should not be started unless the results of the completed phase(s) have been reviewed by competent authorities, in order to that certify the results of the completed phase(s) are valid for authorization of the start a new phase of the clinical trial. For this purpose, at the end of each phase of a clinical trial study, competent authorities evaluate whether the new drug is safe, efficient, and effective for the treatment of the target disease ( Figure 4 ).

An external file that holds a picture, illustration, etc.
Object name is ijms-23-02850-g004.jpg

Status of clinical phase within clinical trials using stem cells.

Early Phase I emphasizes the effects of the drug on the human body and how the drug is processed in the body.

Phase I of a clinical trial is carried out to ensure that a new treatment is safe and to determine how the new medicine works in humans. The FDA has estimated that about 70% of the studies pass this phase.

In Phase II, the accurate dose is determined and initial data on the efficiency and possible side effects are collected. The FDA has estimated that roughly 33% of the studies move to the next phase.

Phase III evaluates the safety and effectiveness of products. The result of this phase is submitted to the FDA/EMA for new product approval, which allows manufacturing and marketing of the drug. The FDA has estimated that 25%–30% of the drugs pass at this phase.

Phase IV take place after the approval of new products and is carried out to determine the public safety of the new product [ 142 , 143 , 144 ].

The number of participants and the duration: A new stem cell product is eligible for marketing after completing successful clinical trial phases. As the new product has been used on volunteers and the effects/side effects of the drug have also been followed for a long time throughout the different phases, it is now possible to make a decision regarding its introduction to the market for public use. The number of participants and the duration of long-term follow-up in each study and each phase differ ( Figure 5 and Figure 6 ). The number of volunteers that participate in each phase of a clinical trial study varies, as each phase has a different target. The FDA has recommended 20–80, 100–300, and several hundred to thousands of volunteers for Phase I, Phase II, and Phase III, respectively [ 144 , 145 ]. Although the FDA has defined a range for enrolments per phase, the number of participants can vary depending on the type of disease. The number of participants for clinical studies in rare diseases will be lower than when studying common diseases. Searching for stem cells in clinicaltrial.gov, studies can be found with only one participant (e.g., {"type":"clinical-trial","attrs":{"text":"NCT02235844","term_id":"NCT02235844"}} NCT02235844 , {"type":"clinical-trial","attrs":{"text":"NCT02383654","term_id":"NCT02383654"}} NCT02383654 , {"type":"clinical-trial","attrs":{"text":"NCT03979898","term_id":"NCT03979898"}} NCT03979898 , and {"type":"clinical-trial","attrs":{"text":"NCT01142856","term_id":"NCT01142856"}} NCT01142856 ). The sponsor/investigator must provide the FDA with strong documentation regarding the selection of such a number of volunteers. The volunteers for each clinical trial study, before attending, should be informed about the enrolment criteria of each study, possible side effects, and the advantages of the study.

An external file that holds a picture, illustration, etc.
Object name is ijms-23-02850-g005.jpg

Enrolment of clinical trials using stem cells.

An external file that holds a picture, illustration, etc.
Object name is ijms-23-02850-g006.jpg

The duration of each clinical trial study using stem cells.

Age of participants: Roughly 190,000 people participated in all the completed clinical trial studies using stem cells that had been registered. Each clinical study was performed in different age groups, which differed among the various studies depending on the type of drug, type of disease, and sponsor decision, as shown in Figure 7 .

An external file that holds a picture, illustration, etc.
Object name is ijms-23-02850-g007.jpg

The age of patients participating in clinical trials using stem cells.

Number of clinical trial studies: The number of clinical trial studies increased gradually from 2000 to 2014, although it fluctuated after 2014 but did not change significantly ( Figure 8 ). The reason for this increase in 2014 is not clear, but it may have been related to the introduction of the first advanced medicinal therapy product containing stem cells (Holoclar) by the EMA in 2014–2015 [ 146 ].

An external file that holds a picture, illustration, etc.
Object name is ijms-23-02850-g008.jpg

The proportion of clinical trials using stem cells by year: ( A ) the proportion of new clinical trial studies using stem cells by year (green bar) and the proportion of registration results accordingly (orange color line); ( B ) the proportion of completed registered clinical trial studies using stem cells by year (blue bar) and the updated results of completed clinical trial studies using stem cells by year (orange line).

Place of study: According to economic website reports, the cell therapy market has grown significantly in recent years, and it is expected to grow more in the coming years; therefore, many countries have begun research in this field. Our data from clinicaltrial.gov showed that the United States has conducted the most clinical trials using stem cells ( Figure 9 ). Government agencies, industry, individuals, universities, and private organizations have all invested in stem-cell-based therapy. The number of stem-cell-based companies has rapidly increased in recent years, and a brief overview of the submitted clinical trial studies indicated that the studies were mostly aimed at introducing therapeutic products for clinical applications. Therefore, we can expect the introduction of stem-cell-based products to the market.

An external file that holds a picture, illustration, etc.
Object name is ijms-23-02850-g009.jpg

The registered and completed clinical trial studies using stem cells according to participating countries: ( A ) top 10 participating countries with registered clinical trials using stem cells; and ( B ) top 10 countries based on the completion of registered clinical trials using stem cells.

As indicated above, translational research from the laboratory to clinical services has many layers which must be passed through, each with its own requirements and measurements. Therefore, the only way to introduce a new stem-cell-based product onto the market is for competent authorities to make sure that the discovery is safe and effective for its intended human use, and that the product has successfully passed all of the clinical trial stages.

5. Challenges and Future Directions

One of the most important issues regarding the introduction of a new product for use in humans through a clinical trial is evaluation of its safety. Although many clinical trials have been performed using stem cells for the treatment of various diseases, as stem-cell-based therapies are one of the newest groups of therapeutic products in medicine, it is very hard to introduce new products based on stem cells onto the market, as many different parameters must be evaluated. There are several concerns regarding stem-cell-based therapies, including genetic instability after long-term expansion, stem cell migration to inappropriate regions of the body, immunological reaction, and so on. However, all challenges depend on the type of stem cell (e.g., embryonic stem cell, adult stem cell, iPS), type of disease, route of administration, and many other factors. Almost all researchers in the field of stem cell therapy believe that despite stem cells having great potential to treat disease through their intrinsic potential, unproven stem-cell-based therapies that have not been shown to be safe or effective may be accompanied by very serious health risks. In order to receive clinical trial approval from a competent regulatory authority, different tests must be performed for each study phase, and the results of one study should not be generalized to another study. The FDA and EMA have defined different regulations to ensure that stem-cell-based products are consistently controlled through the use of different preclinical studies (in vitro and in vivo). Based on these preclinical data, the FDA and EMA have the authority to approve a clinical trial study, as discussed in this review.

Another challenge that researchers and companies face is the duration of a clinical trial study before a stem-cell-based product can be introduced onto the market. At present, hematopoietic progenitor cells are the only FDA-approved product for use in patients with defects in blood production, while other stem-cell-based products used in clinical trials have not yet been introduced to the market.

In the past few years, several clinical trials have been conducted using stem cells, most of which have indicated the safety and high efficiency of stem-cell-based therapies. An attractive future option for regenerative medicine is the use of cell derivatives, including exosomes, amniotic fluid, Wharton’s jelly, and so on, for the treatment of diseases. Recently, the safety and efficiency of these products have been evaluated and optimized in preclinical studies. In addition, regenerative medicine using modified stem cells and combinations of stem cells with scaffolds and chemicals to overcome stem cell therapy challenges and increase the associated efficiency are two important future directions of research. However, establishing a safe method for stem cell modification and moving this technology toward clinical trial studies requires many preclinical studies.

The regenerative medicine market is developing and, due to encouraging findings in preclinical studies and predictable economic benefits, competition has increased between companies focused on the development of cell products. Therefore, government agencies, industries, individuals, universities, and private organizations have invested heavily into the development of the regenerative medicine market in recent years, such that we can be more hopeful about the future of stem-cell-based therapies.

6. Conclusions

In recent years, regenerative medicine has become a promising treatment option for various diseases. Due to their therapeutic potential, including the inhibition of inflammation or apoptosis, cell recruitment, stimulation of angiogenesis, and differentiation, stem cells can been seen as good candidates for regenerative medicine. In the last 50 years, more than 40,000 research papers have focused on stem-cell-based therapies. In this review study, we present a general overview of the translation of stem cell therapy from scientific ideas to clinical applications. Multiple mechanisms causing disease could be reversed by stem cells, due to their tremendous therapeutic potential. However, preclinical studies including in vitro and in vivo experiments are necessary to evaluate the potential of stem-cell-based treatments. Through preclinical research, it is possible to present scientific evidence and optimal treatment options for subsequent clinical studies. Before starting a clinical trial based on preclinical data, the application must be approved by a relevant regulatory administration, such as the FDA, EMA, or another organization. If the application is for the use of a new drug (including stem cells) which has never been tested before, the submission of an IND is required for FDA approval. Approximately 50% of clinical trials using stem cells take 2 to 5 years to complete. To minimize possible side effects, every new stem cell product should be approved for clinical marketing only after completing Phase I–IV clinical trials successfully. Interestingly, the number of stem-cell-based companies aimed at introducing clinical applications has rapidly increased in recent years. Therefore, it may be possible to find stem-cell-based products on the clinical market in the near future. As described in this paper, there are several steps that should be carried out on the path from the laboratory to the clinical setting. To develop new stem-cell-based medicine for the clinical market, researchers should follow the guidelines suggested by the relevant authorities. Through these well-controlled development processes, researchers can achieve safe and effective stem-cell-based therapies, thus brings their research ideas into the clinical field.

Author Contributions

All authors have read and agreed to the published version of the manuscript.

This review funded by National Institutes of Health grant: R01HD087417-01A1, R01HD094378-01, R01HD094380-01, R01HD100367-01, R01HD100563, R01HD100563.

Conflicts of Interest

The author has no conflicts of interest to declare.

Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

IMAGES

  1. (PDF) What are Stem Cells?

    stem cells research papers

  2. 😊 Research papers on stem cells. Stem Cells Research Paper ⋆ Research

    stem cells research papers

  3. (PDF) Advances in stem cell research

    stem cells research papers

  4. Essay On Stem Cells

    stem cells research papers

  5. (PDF) Stem Cells Research and its Applications: A Review

    stem cells research papers

  6. (PDF) Stem Cell Research

    stem cells research papers

VIDEO

  1. The future of Sam cell research speech July 13 2023 

  2. Stem Cells Part 2: The Ethics

  3. Mini-organs grown from stem cells for the first time

  4. STEM CELLS: Long Term Live Imaging of Mouse Embryonic Stem Cells for 15 hours

  5. Stem Cells Research: Cures & Treatments

  6. Stem cells

COMMENTS

  1. Stem cells

    Stem cells are cells that have the capacity to self-renew by dividing and to develop into more mature, specialised cells. Stem cells can be unipotent, multipotent, pluripotent or totipotent ...

  2. Stem cells: a comprehensive review of origins and emerging clinical

    Embryonic stem cells have been utilized in the past for research, but ethical concerns have led to them being replaced largely by stem cells derived from other origins. 12 Common tissues from which adult oligopotent and unipotent stem cells are isolated include bone marrow, adipose tissue, and trabecular bone. 13 Bone marrow has traditionally ...

  3. Stem cells: past, present, and future

    In recent years, stem cell therapy has become a very promising and advanced scientific research topic. The development of treatment methods has evoked great expectations. This paper is a review focused on the discovery of different stem cells and the potential therapies based on these cells. The genesis of stem cells is followed by laboratory steps of controlled stem cell culturing and derivation.

  4. Stem Cell Research

    Stem Cell Research is dedicated to publishing high-quality manuscripts focusing on the biology and applications of stem cell research. Submissions to Stem Cell Research, may cover all aspects of stem cells, including embryonic stem cells, tissue-specific stem cells, cancer stem cells, developmental …. View full aims & scope.

  5. Stem-cell research

    Stem-cell research articles from across Nature Portfolio. Stem-cell research is the area of research that studies the properties of stem cells and their potential use in medicine. As stem cells ...

  6. Innovations in Stem Cell Biology 2023

    Innovations in Stem Cell Biology 2023. ... clinical and translational research and tissue stem cells. ... Two side-by-side papers report that H3K27me3 deposited by polycomb repressive complex 2 ...

  7. Articles

    Citation: Stem Cell Research & Therapy 2024 15:100 Content type: Research Published on: 8 April 2024. View Full Text ... 1.440 - SNIP (Source Normalized Impact per Paper) 1.498 - SJR (SCImago Journal Rank) 2023 Speed 33 days submission to first editorial decision for all manuscripts (Median) 211 days submission to accept (Median)

  8. Home page

    Stem Cell Research & Therapy is the major forum for translational research into stem cell therapies. An international peer-reviewed journal, it publishes high-quality open access research articles with a special emphasis on basic, translational and clinical research into stem cell therapeutics and regenerative therapies, including animal models and clinical trials.

  9. Stem cells

    Stem-Cell Aging and Pathways to Precancer Evolution. C.H. Jamieson and I.L. WeissmanN Engl J Med 2023;389:1310-1319. Stem cells can renew themselves without differentiating. Aging, inflammation ...

  10. Stem Cell Research and Health Education

    The fourth main controversy concerns whether adult stem cells are as beneficial as embryonic stem cells. A seminal paper from a group led by Catherine Verfaillie (see Jiang et al. 107) reported that adult stem cells from the bone marrow of rats, which they called "multipotent adult progenitor cells" (MAPCs), had the potential to ...

  11. About the journal

    About the journal. STEM CELLS, a peer-reviewed journal published monthly, provides a forum for prompt publication of original investigative papers and concise reviews.STEM CELLS is read and written by clinical and basic scientists whose expertise encompasses the rapidly expanding fields of stem and progenitor cell biology.. STEM CELLS welcomes original articles and concise reviews describing ...

  12. Advances in stem cell research and therapeutic development

    Stem Cell Research & Therapy (2023) Despite many reports of putative stem-cell-based treatments in genetic and degenerative disorders or severe injuries, the number of proven stem cell therapies ...

  13. Stem cells: past, present, and future

    This paper is a review focused on the discovery of different stem cells and the potential therapies based on these cells. The genesis of stem cells is followed by laboratory steps of controlled stem cell culturing and derivation. ... After this success, the method opened a new field in stem cell research with a generation of iPSC lines that can ...

  14. STEM CELLS

    STEM CELLS, a peer reviewed journal published monthly, provides a forum for prompt publication of original investigative papers and concise reviews. STEM CELLS is read and written by clinical and basic scientists whose expertise encompasses the rapidly expanding fields of stem and progenitor cell biology. STEM CELLS welcomes original articles and concise reviews describing basic laboratory ...

  15. Cells

    Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications. ... Adult stem cells on the other ...

  16. Stem cell-based therapy for human diseases

    The discovery of hPSCs, including human embryonic stem cells (hESCs) and human induced pluripotent stem cells (hiPSCs), has revolutionized stem cell research and cell-based therapy. 98 hESCs were ...

  17. Current state of stem cell-based therapies: an overview

    Stem cell-based therapies. Stem cell-based therapies are defined as any treatment for a disease or a medical condition that fundamentally involves the use of any type of viable human stem cells including embryonic stem cells (ESCs), iPSCs and adult stem cells for autologous and allogeneic therapies ().Stem cells offer the perfect solution when there is a need for tissue and organ ...

  18. Stem cells: past, present, and future

    In recent years, stem cell therapy has become a very promising and advanced scientific research topic. The development of treatment methods has evoked great expectations. This paper is a review focused on the discovery of different stem cells and the potential therapies based on these cells. The genesis of stem cells is followed by laboratory ...

  19. Frontiers in Genetics

    Telocytes in Tissue Organization in Health and Disease. Hanan H. Abd- ElHafeez. Alaa Sayed Abou-Elhamd. Soha Soliman. Sanda Maria Cretoiu. 337 views. Covers all aspects of stem cell research - from molecular and cell biology, to tissue regeneration and translational research.

  20. Cells

    Special Issue Information. Keywords. Published Papers. A special issue of Cells (ISSN 2073-4409). This special issue belongs to the section "Stem Cells". Deadline for manuscript submissions: closed (15 April 2022) | Viewed by 87220. Printed Edition Available! A printed edition of this Special Issue is available here.

  21. Stem Cell Research

    Stem Cell Research is dedicated to publishing high-quality manuscripts focusing on the biology and applications of stem cell research. ... Research papers: Research papers should not exceed 55,000 characters (including spaces, references, methods, figure legends, and an abstract of 200 words or less). We accept a maximum of 7 figures/tables for ...

  22. Stem cell-based regenerative medicine

    Despite the increasing interest in totipotent and pluripotent stem cells, unipotent stem cells have not received the most attention in research. A unipotent stem cell is a cell that can create cells with only one lineage differentiation. Muscle stem cells are one of the example of this type of cell . The word 'uni' is derivative from the ...

  23. The Complex Terrain of Stem Cell Research

    This essay about the challenges and limitations of stem cell research and therapy outlines the significant hurdles that must be overcome to fully realize the potential of stem cells in medicine. It discusses technical difficulties in controlling stem cell differentiation, the risk of immune rejection, the potential for stem cells to form tumors ...

  24. Stem Cell Therapy: From Idea to Clinical Practice

    From 1971 to 2021, 40,183 research papers were published regarding stem-cell-based therapies. All of these studies were conducted around discoveries and for the goal of "Stem Cell Therapy" based on the therapeutic efficacy of stem cells . As basic stem cell research has soared over the past few years, "translation research", a ...

  25. Role of exosomes in the communication and treatment between OSCC and

    Oral squamous cell carcinoma (OSCC) is a prevalent cancer that needs new therapeutic targets due to the poor postoperative prognosis in patients. Exosomes are currently one of important research areas owing to their unique properties. Exosomes are capable of acting as drug transporters, as well as facilitating interactions between OSCC and normal cells. Exosomes can be detected in body fluids ...