Skip to main content

Cord blood in regenerative medicine: do we need immune suppression?

Abstract

Cord blood is currently used as an alternative to bone marrow as a source of stem cells for hematopoietic reconstitution after ablation. It is also under intense preclinical investigation for a variety of indications ranging from stroke, to limb ischemia, to myocardial regeneration. A major drawback in the current use of cord blood is that substantial morbidity and mortality are associated with pre-transplant ablation of the recipient hematopoietic system. Here we raise the possibility that due to unique immunological properties of both the stem cell and non-stem cell components of cord blood, it may be possible to utilize allogeneic cells for regenerative applications without needing to fully compromise the recipient immune system. Issues raised will include: graft versus host potential, the immunogeneicity of the cord blood graft, and the parallels between cord blood transplantation and fetal to maternal trafficking. The previous use of unmatched cord blood in absence of any immune ablation, as well as potential steps for widespread clinical implementation of allogeneic cord blood grafts will also be discussed.

Background

The first widespread utilization of cord blood as a stem cell source was in the treatment of pediatric hematological malignancies after myeloablative conditioning. Since matching requirements for this type of transplant are not as strict as for hematopoietic stem cell sources, cord blood began gaining acceptance in adult patients lacking bone marrow donors [1–6]. Outside the area of oncology, the clinical use of cord blood has expanded into various areas that range from reconstituting a defective immune system [7], to correcting congenital hematological abnormalities [8], to inducing angiogenesis [9]. A sample of some of cord blood clinical studies addressing non-malignant disorders is presented in Table 1.

Table 1 Non-Malignant Disorders Treated With Cord Blood

In addition to current clinical use, cord blood is currently under intense experimental investigation in preclinical models of pathophysiologies that range from myocardial ischemia, to stroke, to muscle regeneration [10–13]. It is anticipated that in the next several years that widespread clinical entry of cord blood for non-hematopoietic tissue regeneration will occur. When this happens, the main question will be how to select patients that can be myeloablated so as to allow acceptance of the cord blood graft. According to current dogma in the discipline, it is believed that myeloablation, or at minimum non-ablative immune suppression of the recipient is strictly required. In situations of hematological malignancy it is desirable to myeloablate the recipient so as to eradicate the leukemic population while creating "space" for the donor cells to engraft. However, the question is, in patients that are not suffering from a disease that is associated with an aberrant bone marrow such as hematological malignancies or immunological dysfunctions, how is it justifiable to subject them to the high levels of morbidity and mortality associated with immune suppression? For conditions such as Krabbe disease where patients rarely survive beyond the age of 2 and cord blood transplant was demonstrated to induce 100% survival in a subgroup of patients treated [14], the justification for myeloablation can be made. However for conditions such as post-stroke regeneration or induction of angiogenesis in angina patients, in which the population already suffers from major comorbidities and the potential benefit of cord blood therapy is only speculative, the ability to justify myeloablative protocols rapidly diminishes.

The purpose of this paper is to put forth the notion that the immunology of cord blood transplants for regenerative applications has to be viewed differently from the perspective and the practice of cord blood transplants for hematopoietic reconstitution. Specifically, we will provide reasons and rational for why in some situations, administration of cord blood, or stem cells derived thereof, may be possible with no, or minimal immune suppression of the recipient. Evaluation of this possibility will lead to acceleration of clinical entry and wide-spread utilization of cord blood transplants for non-hematopoietic indications.

Regenerative cells in cord blood

Numerous publications have described the regenerative ability of cord blood cells in a myriad of preclinical disease models. Although the purpose of this paper is to discuss the immunology of cord blood transplants for regenerative uses, we will first overview some of the therapeutic stem cell populations found in cord blood so that we may discuss their immunogenic consequences later in the paper.

Hematopoietic stem cells

The original clinically attractive feature of cord blood was the high concentration of hematopoietic stem cells, which is similar to that found in bone marrow: approximately 0.1–0.8 CD34+ cells per 100 nucleated cells. However, in contrast to marrow, CD34+ cells from cord blood possess higher proliferative potential in vitro [15], superior numbers of long term culture initiating cells and SCID repopulating cells [16, 17], as well as higher telomerase activity [18]. The potent hematopoietic activity of cord blood derived CD34+ cells may be attributed to the fact that cord blood is a much more developmentally immature source of stem cells as opposed to stem cells derived from adult sources. Attesting to the robust hematopoietic activity of cord blood derived CD34+ cells in comparison to bone marrow cells is the fact that successful reconstitution, albeit delayed, of post-ablative hematopoiesis occurs in patients receiving approximately one tenth of the total nucleated cell number in a cord blood graft compared to a bone marrow graft.

Endothelial progenitors and angiogenesis stimulating cells

In addition to being a source of hematopoietic cells, cord blood contains potent angiogenesis stimulating cells. Several phenotypes have been ascribed to cord blood angiogenic stimulating cells. In one report, the CD34+, CD11b+ fraction, which is approximately less than half of the CD34+ fraction of cord blood was demonstrated to possess ability to differentiate into functional endothelial cells in vitro and in vivo [19]. In another report, VEGF-R3+, CD34+ cells were shown to possess not only the ability to differentiate into endothelial cells in vivo, but also to be able to expand approximately 40-fold in vitro and subsequently maintain angiogenic function in vivo. The same study demonstrated that the concentration of this endothelial progenitor fraction found in cord blood CD34+ cells is approximately tenfold higher as compared to bone marrow CD34+ cells [20]. Regardless of the phenotype of the cord blood cell with angiogenesis stimulating ability, unfractionated cord blood mononuclear cells have also been used in numerous animal models [21–23], as well as in the clinic [9], for successful stimulation of angiogenesis.

In addition to endothelial progenitors, mesenchymal stem cells (discussed below in more detail), which are found in cord blood, are known to secrete numerous cytokines and growth factors such as VEGF and FGF-2 [24, 25] which stimulate angiogenic processes. In fact, there are reports of mesenchymal stem cells contributing to angiogenesis through direct differentiation into endothelial cells [26].

Mesenchymal stem cells

Mesenchymal stem cells are a type of cell capable of differentiating into various non-hematopoietic tissues. Currently this cell population is second to bone marrow stem cells in terms of clinical entry in that Phase III clinical trials are already underway with these cells. Mesenchymal stem cells are classically defined as adhere to plastic and expressing a non-hematopoietic cell surface phenotype, consisting of CD34-, CD45-, HLA-DR-, while possessing markers such as STRO-1, VCAM, CD13, CD29, CD44, CD90, CD105, SH-3, and STRO-1 [27]. To date mesenchymal stem cells have been purified from bone marrow [28], adipose tissue [29], placenta [30, 31], scalp tissue [32] and cord blood [33]. Cord blood-derived mesenchymal stem cells have demonstrated ability to differentiate into a wide variety of tissues in vitro including neuronal [34–36], hepatic [37, 38], osteoblastic [39], and cardiac [33]. An important aspect of this cell population is their anti-inflammatory and immunomodulator activity. For example, they constitutively secrete immune inhibitory cytokines such as IL-10 and TGF-β while maintaining ability to present antigens to T cells, thus suggesting they may act as a tolerogenic antigen presenting cell [40, 41]. Conceptually, the mesenchymal content of umbilical cord blood grafts may explain the tolerogenic capabilities, which some have speculated to be donor specific.

Although the majority of published studies have examined bone marrow derived mesenchymal stem cells, and thus are outside the scope of the present review, it is important to note differences between mesenchymal stem cells derived from different sources. A recent study compared mesenchymal stem cells from bone marrow, cord blood and adipose. Cord blood mesenchymal stem cells which were capable of expansion to approximately 20 times, whereas adipose derived cells expanded an average of 8 times and bone marrow derived cells expanded 5 times [42]. This, and other studies support the important role of mesenchymal stem cell content in the biological activities of the cord blood graft.

Unrestricted somatic stem cells

Cells with markers and activities resembling embryonic stem cells have been found in cord blood. Zhao et al identified a population of CD34- cells expressing OCT-4, Nanog, SSEA-3 and SSEA-4, which could differentiate into cells of the mesoderm, ectoderm and endoderm lineage. In vivo administration of these cells into the streptozotocin-induced murine model of diabetes was able to significantly reduce hypoglycemia [43]. The existence of cells with such pluripotency in cord blood was also observed by Kogler et al who identified an Unrestricted Somatic Stem Cell (USSC) with capability of differentiation into functional osteoblasts, chondroblasts, adipocytes, hematopoietic and neural cells. USSC were demonstrated to be capable of > 40 population doublings in vitro without spontaneous differentiation or loss of telomere length. Interestingly, administration of these cells (derived from human cord blood) into fetal sheep resulted significant human hematopoiesis (up to 5%), hepatic chimerism with > 20% albumin-producing human parenchymal hepatic cells, as well as detection of human cardiomyocytes. The mechanism of differentiation was not associated with fusion [44]. Support for presence of such pluripotency in cord blood cells also comes from a similar experiment in which CD34+ Lineage- cells were transfected with GFP and administered in utero to goats. GFP+ cells were detected in blood, bone marrow, spleen, liver, kidney, muscle, lung, and heart of the recipient goats (1.2–36% of all cells examined) [45]. In other studies, McGuckin et al demonstrated that culture of cord blood cells that have been depleted of lineage committed cells in a TPO, c-kit, and flt-3 ligand culture express markers of embryonic stem cells including TRA-1–60, TRA-1–81, SSEA-4, SSEA-3 and Oct-4. Functionally, these cells also demonstrated pluripotent differentiation ability and in vivo hematopoietic activity [46–48].

Cord blood transplantation without host preconditioning: will there be GVHD?

The possibility of using cord blood in absence of host preconditioning would open up the door for a multitude of stem cell therapeutic applications. The currently dogma amongst cord blood transplanters is that administration of allogeneic cord blood, even if HLA-matched, would in the best case scenario lead to immunologically-mediated rejection or the graft, and in the worst case cause GVHD. Here we provide rationale for the preliminary clinical exploration of cord blood administration in a non-preconditioned host.

ABO-matched, HLA-mismatched transfusions

In the 1930s it was reported that cord blood could be safely used as a substitute for peripheral blood for performing transfusions [49]. Since HLA-matching was not available at that time and no adverse effects were noted, the feasibility of cord blood administration to a non-preconditioned host was suggested. A more recent Lancet publication described the use of cord blood as a source of blood donation for malaria infested regions in Africa. 128 pediatric patients with severe anemia needing transfusions were transplanted with an average of 85 ml of ABO matched cord blood with no HLA matching. No report of graft versus host was noted, and cord blood was proposed as a transfusion source when peripheral blood is not available due to economical or social reasons [50]. An extensive review of 129 patients transplanted with a total of 413 Units of cord blood (average 86 ml) with no preconditioning or HLA matching between 1999 to 2004 was published by Bhattacharya [51]. Of these patients, aged 2–86 years old and suffering from advanced cancer (56.58%) and other diseases (43.42%) such as ankylosing spondylitis, lupus erythematosus, rheumatoid arthritis, aplastic anemia, and thalassemia major, no immunological reactions were noted in patients followed for 1–4 years. The same author reported other patient cohorts that have been similarly treated and had no GVHD or other immune reactions [52–55]. Furthermore, transfusion of cord blood in non-HLA matched recipients was also associated with transient increases in peripheral CD34 counts, without evidence of GVHD in patients with cancer and HIV [56, 57]. Unfortunately in these studies did not perform long-term molecular analysis for chimerism. Despite this drawback, it is evident from the initial work in the 1930s, to the numerous cases reported by Bhattacharya, to the publication in Lancet, that administration of cord blood is a safe procedure not associated with immunological consequences. Thus based on the current data, the worse a cord blood transplant will do is do nothing.

Administration of adult lymphocytes does not elicit GVHD

The cells that are "dangerous" from the cord blood from a GVHD perspective are lymphocytes that may have alloreactive potential. Lymphocytes from cord blood, in contrast to adult blood, are generally immature and usually do not secrete as many inflammatory cytokines. Therefore administration of allogeneic lymphocytes purified from adult blood would be a much more dangerous procedure, at least as far as GVHD is concerned, in contrast to administration of lymphocytes from cord blood. The fact is administration of lymphocytes from paternal sources has been performed in numerous reports in the clinical practice of using "paternal lymphocyte immunotherapy" for treatment of spontaneous abortions. Numerous trials have been conducted administering doses of up to 2 × 109 paternal lymphocytes into pregnant mothers who have had recurrent miscarriages [58, 59]. These doses are higher than the 1.5–3 × 107 nucleated cells/kg administered during a cord blood transplant [60]. Interestingly, in pregnant women administered these high doses of completely allogeneic cells, no GVHD has ever been observed, although Th2 immune deviation has been reported by some groups [61, 62]. Thus according to the current evidence, there is no fear of GVHD being induced after cord blood transplant. Bhattacharya even administered as many as 32 units of cord blood to an individual without seeing GVHD [57].

Homeostatic proliferation in lymphopenic environment causes GVHD

The reader will ask, in response to the above arguments regarding GVHD-inducing ability of cord blood, "why is GVHD, a clinical reality in patients receiving cord blood for hematological malignancies?" The answer is that current day cord blood transplants take place following ablation of host T cells. This creation of an "empty compartment" allows for homeostatic expansion of the newly introduced T cells, which primes them for aggressive immune reactions and alleviates their requirement for costimulation [63]. It is known from the transplantation literature that T cells reconstituting a host that has been lymphoablated are resistant to costimulatory blockade and tolerance induction [64]. Furthermore, the pioneering experiments of Rosenberg's group demonstrated that infusion of tumor specific lymphocytes following ablation of the recipient T cells, using conditions similar to those used in cord blood transplant preconditioning allows for highly aggressive anti-tumor responses that otherwise would not be observed [65]. Further supporting the concept that reconstitution of a lymphocyte deficient immune system can cause immune hyperreactivity comes from clinical observations of "autologous GVHD" in patients administered drugs associated with induction of lymphopenia [66, 67]. We therefore propose that GVHD is not an intrinsic property of the allogeneic cells introduced into the host, but a result of the lymphoablation induced in the recipient prior to cellular administration.

Will the graft be cleared?

If cord blood can be administered into a non-preconditioned patient without fear of GVHD, then the next question arises as to whether the infused cells will actually endow some type of benefit or be rapidly cleared by the immune system. As previously mentioned, biological effects of mismatched cells, even if they are cleared by the immune system may have a beneficial role in inflammatory pathologies through exertion of a Th2 phenotype as seen in mothers being administered their mate's lymphocytes. Furthermore, even if transplanted cells are cleared by the immune system, it is known that apoptotic cells can mediate various therapeutic anti-inflammatory effects that are clinically relevant [68]. Animal models suggest that human cord blood cells may be administered for therapeutic benefit into other species, in absence of immune suppression. For example, Vendrame et al reported that administration of human cord blood into a non-immunosuppressed rat model of stroke resulted in increased neuronal survival, and interestingly decreased inflammatory infiltrates as compared to controls [69].

Fetal cells do not get cleared by maternal immune system

However, we do not believe that complete immune mediated clearing of regeneratively-important cord blood constituents occurs in the allogeneic setting. One reason for this notion comes from an interesting phenomena observed in pregnancy. It is well established that during pregnancy fetal cells enter maternal circulation [70]. While circulating CD34+ cells of fetal origin are found a percentage of women who have had children [71], in the bone marrow 100% of women who have had children were found to contain offspring-derived mesenchymal cells in their bone marrow [72]. Although some studies have correlated autoimmunity with residual lymphocytes causing a GVHD-like reaction in the mother, more careful analysis of these studies show that immune cells of fetal origin are largely outnumbered by cells of maternal origin. This is the basis for the proposition of Khosrotehrani et al that the fetal cells are actually "pregnancy associated progenitor cells" that act as allogeneic "repair cells" [73]. The authors of this hypothesis believe that these repair cells are actually migrating to the site of autoimmune damage in order to control injury and cause regeneration. The authors cite numerous examples in support of their idea, more notably, a case report of a hepatitis C patient who stopped treatment but disease relapse was not observed. Biopsy analysis demonstrated the liver parenchyma was heavily populated with cells of male origin that based on DNA polymorphism analysis were derived from a previous pregnancy more than a decade earlier [74]. Additionally, they cite reports of fetal cells differentiating into thyroid, cervix, gallbladder and intestinal epithelial cells [75–78]. Data from animal models, although scarce, supports the notion that fetal cells trafficking into the mother may play some reparative function. For example, it was reported that EGFP expressing fetal cells would selectively home into damaged maternal renal and hepatic tissues after gentamycin and ethanol induced injury [79]. Furthermore, another study demonstrated that subsequent to excitotoxic injury in the maternal brain, fetal-derived EFGP positive cells can be identified which express morphology and markers of neurons, astrocytes, and oligodendrocytes [80]. The authors of this paper are not stating that the fetal transfer of mesenchymal cells to the maternal host is an exact duplicate of an allogeneic cord blood transplant in absence of immune suppression. Rather, we are proposing fetal to maternal trafficking as a possible example of a natural biological situation in which stem cells may persist in an allogeneic environment without induction of complete tolerance.

Mesenchymal stem cells do not need myeloablation for efficacy

Currently there are several ongoing clinical trials in Phase I-III using "universal donor" mesenchymal stem cells in non-conditioned recipients for treatment of Crohn's disease [81], GVHD [82], and myocardial infarction [83]. Although these cells are bone marrow expanded mesenchymal cells, the superior proliferative potential of cord blood mesenchymal cells may allow them to not only escape immune destruction, but also expand in vivo and mediate therapeutic effects superior to those derived from the bone marrow. The fact that regulatory agencies have allowed advancement of "off-the-shelf" universal donor mesenchymal stem cells supports the numerous reports of clinical efficacy in an allogeneic setting.

Clinical evidence of cord blood efficacy in absence of myeloablation

To the knowledge of the authors, there have only been 3 published reports of non-conditioned recipients receiving cord blood cells for regenerative purposes. The first report is of 4 patients with Buerger's disease who were administered 4/6 HLA matched allogeneic cord blood cells locally in the area of limb ischemia. In all patients rest pain disappeared and necrotic lesions healed approximately 4 weeks subsequent to cell administration. Significant angiographically evidenced neoangiogenesis proximal to area of administration was observed [84]. No graft versus host, or inflammation was observed at the site of injection. The second published report is of a matrix delivered allogeneic cord blood dose to a patient who suffered from spinal cord injury. Improved sensory perception and movement was observed, as well as CT and MRI observation of tissue regeneration at site of injury was reported [85]. The third report described 2 patients with non-healing wounds who were treated with autologous fibrin glue containing matched (2 mismatches allowed) allogeneic cord blood isolated CD34+ cells. Significant wound healing was observed with no indication of GVHD at 3 and 7 months subsequent to treatment [86]. These reports give some suggestion that administration of mismatched cord blood stem cells may endow potentially therapeutic benefit without rejection, at least immediate rejection, by the host versus graft process. More importantly, these studies demonstrate that allogeneic cord blood cells may be used clinically without immune-mediated clearing before therapeutic properties are exerted.

Strategies for clinical implementation

Given the unique regenerative capabilities of cord blood, the easy accessibility of HLA matched donors, and relative inexpensiveness as compared to other cellular therapies; it is of great interest therapeutically to expand its use into non-conditioned recipients. Below we will provide some initial thoughts on how this may be performed.

Cord blood together with stem cell activators

One simple method of stem cell therapy would be administration of cord blood units in patients with degenerative diseases in the form of direct transfusions has described by Bhattacharya [57]. This approach, however has not demonstrated significant regenerative activity. A more promising approach may be administration of cord blood cells in combination with activators of endogenous stem cells. For example, clinically used agents such as thalidomide [87], valproic acid [88], or 5-azacytidine [89, 90] all have demonstrated ability to induce proliferation of CD34+ stem cells in vitro and/or in vivo. In the majority of cases, stem cell activating compounds are administered for therapeutic benefit in absence of addition of exogenous stem cell sources. One example is current clinical trials of hCG and EPO for treatment of acute stroke [91]. The addition of exogeneous allogeneic stem cells to clinically accepted stem cell expanding agents may be a relatively easy initial clinical trial of regenerative cord blood therapy in absence of recipient conditioning. An interesting addition to administration of stem cell activators is the localized introduction of stem cell specific chemoattractant agents at the site in need of repair, followed by systemic administration of cord blood stem cells. Chemoattractant agents could include stromal derived growth factor-1 [92], other various agonists of CXCR-4 [93], or hepatocyte growth factor [94].

Cord blood plus natural chemoattractant

An alternative approach to administration of exogenous stem cell chemoattractants would be to administer stem cells at the narrow window period after tissue injury when endogenous chemoattractants are secreted by the injured tissue. For example, following myocardial infarction, as well as stroke, there is a period of time which concentration of local stem cell chemoattractants are so high that bone marrow derived progenitors are actually mobilized into systemic circulation [95]. Activators of endogenous stem cells may be administered to allow localized tissue repair, while exogenous stem cells are administered to provide support to the activated endogenous cells.

De-immunogenized cord blood graft

Clinical entry of a cord blood therapeutic in patients who are not preconditioned would require a high margin of safety to be met. Accordingly, one approach for beginning cord blood clinical trials may be administration of grafts that are depleted of T cells, B cells, and dendritic cells. In this manner, even the remote possibility of GVHD would be negated, and the there would be no B cell or dendritic cell induced antigenicity. Essentially, the percentage stem cell content of the graft would increase while the possible immunogenic components would decrease. A method of accomplishing this would be the pretreatment of cord blood units with the clinically used anti-CD52 monoclonal antibody CAMPATH. It has been previously demonstrated that this agent can be used in substantially "cleaning" grafts of T cells without affecting hematopoietic activity both in vitro [96] and in the clinic [97]. Furthermore, CAMPATH has been shown to deplete B cells [97], as well as circulating blood dendritic cells [98, 99]. An ideal indication for utilization of such a treatment would be in limb ischemia in which a variety of stem cell sources have demonstrated angiogenic effects, however clinical implementation of stem cells for limb ischemia is currently limited by the difficulties of obtaining and purifying autologous stem cells.

Conclusion

In summary, the authors propose that expanding the use of cord blood to non-preconditioned adult recipients for regenerative purposes would be a great step for the practical advancement of stem cell therapeutics. By overcoming allogeneic barriers in regenerative medicine, the fundamental limitations of autologous cell therapy may result in effective standardized "off-the-shelf" cellular products for regenerative therapeutics. This major step can only be performed by understanding the unique immunology of cord blood grafts, leveraging the graft's regenerative capability for specific indications, and identifying methods of amplifying cellular effects through administration of various drugs.

References

  1. Cornetta K, Laughlin M, Carter S, Wall D, Weinthal J, Delaney C, Wagner J, Sweetman R, McCarthy P, Chao N: Umbilical cord blood transplantation in adults: results of the prospective Cord Blood Transplantation (COBLT). Biol Blood Marrow Transplant. 2005, 11: 149-160. 10.1016/j.bbmt.2004.11.020.

    PubMed  Google Scholar 

  2. Schonberger S, Niehues T, Meisel R, Bernbeck B, Laws HJ, Kogler G, Enzmann J, Wernet P, Gobel U, Dilloo D: Transplantation of haematopoietic stem cells derived from cord blood, bone marrow or peripheral blood: a single centre matched-pair analysis in a heterogeneous risk population. Klin Padiatr. 2004, 216: 356-363. 10.1055/s-2004-832357.

    CAS  PubMed  Google Scholar 

  3. Lekakis L, Giralt S, Couriel D, Shpall EJ, Hosing C, Khouri IF, Anderlini P, Korbling M, Martin T, Champlin RE, de Lima M: Phase II study of unrelated cord blood transplantation for adults with high-risk hematologic malignancies. Bone Marrow Transplant. 2006, 38: 421-426. 10.1038/sj.bmt.1705467.

    CAS  PubMed  Google Scholar 

  4. Tomonari A, Takahashi S, Ooi J, Nakaoka T, Takasugi K, Uchiyama M, Tsukada N, Konuma T, Iseki T, Tojo A, Asano S: Cord blood transplantation for acute myelogenous leukemia using a conditioning regimen consisting of granulocyte colony-stimulating factor-combined high-dose cytarabine, fludarabine, and total body irradiation. Eur J Haematol. 2006, 77: 46-50. 10.1111/j.0902-4441.2006.t01-1-EJH2608.x.

    CAS  PubMed  Google Scholar 

  5. Laporte JP, Lesage S, Portnoi MF, Landman J, Rubinstein P, Najman A, Gorin NC: Unrelated mismatched cord blood transplantation in patients with hematological malignancies: a single institution experience. Bone Marrow Transplant. 1998, 22 Suppl 1: S76-7.

    CAS  PubMed  Google Scholar 

  6. Sanz GF, Saavedra S, Jimenez C, Senent L, Cervera J, Planelles D, Bolufer P, Larrea L, Martin G, Martinez J, Jarque I, Moscardo F, Plume G, Andreu R, de la Rubia J, Barragan E, Solves P, Soler MA, Sanz MA: Unrelated donor cord blood transplantation in adults with chronic myelogenous leukemia: results in nine patients from a single institution. Bone Marrow Transplant. 2001, 27: 693-701. 10.1038/sj.bmt.1702878.

    CAS  PubMed  Google Scholar 

  7. Knutsen AP, Wall DA: Umbilical cord blood transplantation in severe T-cell immunodeficiency disorders: two-year experience. J Clin Immunol. 2000, 20: 466-476. 10.1023/A:1026463900925.

    CAS  PubMed  Google Scholar 

  8. Jaing TH, Hung IJ, Yang CP, Chen SH, Sun CF, Chow R: Rapid and complete donor chimerism after unrelated mismatched cord blood transplantation in 5 children with beta-thalassemia major. Biol Blood Marrow Transplant. 2005, 11: 349-353. 10.1016/j.bbmt.2005.02.003.

    PubMed  Google Scholar 

  9. Tomonari A, Tojo A, Takahashi T, Iseki T, Ooi J, Takahashi S, Nagamura F, Uchimaru K, Asano S: Resolution of Behcet's disease after HLA-mismatched unrelated cord blood transplantation for myelodysplastic syndrome. Ann Hematol. 2004, 83: 464-466. 10.1007/s00277-003-0819-6.

    PubMed  Google Scholar 

  10. Brzoska E, Grabowska I, Hoser G, Streminska W, Wasilewska D, Machaj EK, Pojda Z, Moraczewski J, Kawiak J: Participation of stem cells from human cord blood in skeletal muscle regeneration of SCID mice. Exp Hematol. 2006, 34: 1262-1270. 10.1016/j.exphem.2006.05.009.

    CAS  PubMed  Google Scholar 

  11. Hu CH, Wu GF, Wang XQ, Yang YH, Du ZM, He XH, Xiang P: Transplanted human umbilical cord blood mononuclear cells improve left ventricular function through angiogenesis in myocardial infarction. Chin Med J (Engl). 2006, 119: 1499-1506.

    CAS  Google Scholar 

  12. Leor J, Guetta E, Feinberg MS, Galski H, Bar I, Holbova R, Miller L, Zarin P, Castel D, Barbash IM, Nagler A: Human umbilical cord blood-derived CD133+ cells enhance function and repair of the infarcted myocardium. Stem Cells. 2006, 24: 772-780. 10.1634/stemcells.2005-0212.

    PubMed  Google Scholar 

  13. Newcomb JD, Ajmo CT, Sanberg CD, Sanberg PR, Pennypacker KR, Willing AE: Timing of cord blood treatment after experimental stroke determines therapeutic efficacy. Cell Transplant. 2006, 15: 213-223.

    PubMed  Google Scholar 

  14. Escolar ML, Poe MD, Provenzale JM, Richards KC, Allison J, Wood S, Wenger DA, Pietryga D, Wall D, Champagne M, Morse R, Krivit W, Kurtzberg J: Transplantation of umbilical-cord blood in babies with infantile Krabbe's disease. N Engl J Med. 2005, 352: 2069-2081. 10.1056/NEJMoa042604.

    CAS  PubMed  Google Scholar 

  15. Theunissen K, Verfaillie CM: A multifactorial analysis of umbilical cord blood, adult bone marrow and mobilized peripheral blood progenitors using the improved ML-IC assay. Exp Hematol. 2005, 33: 165-172. 10.1016/j.exphem.2004.10.016.

    CAS  PubMed  Google Scholar 

  16. Ng YY, van Kessel B, Lokhorst HM, Baert MR, van den Burg CM, Bloem AC, Staal FJ: Gene-expression profiling of CD34+ cells from various hematopoietic stem-cell sources reveals functional differences in stem-cell activity. J Leukoc Biol. 2004, 75: 314-323. 10.1189/jlb.0603287.

    CAS  PubMed  Google Scholar 

  17. Hogan CJ, Shpall EJ, McNulty O, McNiece I, Dick JE, Shultz LD, Keller G: Engraftment and development of human CD34(+)-enriched cells from umbilical cord blood in NOD/LtSz-scid/scid mice. Blood. 1997, 90: 85-96.

    CAS  PubMed  Google Scholar 

  18. Sakabe H, Yahata N, Kimura T, Zeng ZZ, Minamiguchi H, Kaneko H, Mori KJ, Ohyashiki K, Ohyashiki JH, Toyama K, Abe T, Sonoda Y: Human cord blood-derived primitive progenitors are enriched in CD34+c-kit- cells: correlation between long-term culture-initiating cells and telomerase expression. Leukemia. 1998, 12: 728-734. 10.1038/sj.leu.2401001.

    CAS  PubMed  Google Scholar 

  19. Hildbrand P, Cirulli V, Prinsen RC, Smith KA, Torbett BE, Salomon DR, Crisa L: The role of angiopoietins in the development of endothelial cells from cord blood CD34+ progenitors. Blood. 2004, 104: 2010-2019. 10.1182/blood-2003-12-4219.

    CAS  PubMed  Google Scholar 

  20. Salven P, Mustjoki S, Alitalo R, Alitalo K, Rafii S: VEGFR-3 and CD133 identify a population of CD34+ lymphatic/vascular endothelial precursor cells. Blood. 2003, 101: 168-172. 10.1182/blood-2002-03-0755.

    CAS  PubMed  Google Scholar 

  21. Cho SW, Gwak SJ, Kang SW, Bhang SH, Song KW, Yang YS, Choi CY, Kim BS: Enhancement of Angiogenic Efficacy of Human Cord Blood Cell Transplantation. Tissue Eng. 2006

    Google Scholar 

  22. Botta R, Gao E, Stassi G, Bonci D, Pelosi E, Zwas D, Patti M, Colonna L, Baiocchi M, Coppola S, Ma X, Condorelli G, Peschle C: Heart infarct in NOD-SCID mice: therapeutic vasculogenesis by transplantation of human CD34+ cells and low dose CD34+KDR+ cells. Faseb J. 2004, 18: 1392-1394.

    CAS  PubMed  Google Scholar 

  23. Le Ricousse-Roussanne S, Barateau V, Contreres JO, Boval B, Kraus-Berthier L, Tobelem G: Ex vivo differentiated endothelial and smooth muscle cells from human cord blood progenitors home to the angiogenic tumor vasculature. Cardiovasc Res. 2004, 62: 176-184. 10.1016/j.cardiores.2004.01.017.

    CAS  PubMed  Google Scholar 

  24. Mayer H, Bertram H, Lindenmaier W, Korff T, Weber H, Weich H: Vascular endothelial growth factor (VEGF-A) expression in human mesenchymal stem cells: autocrine and paracrine role on osteoblastic and endothelial differentiation. J Cell Biochem. 2005, 95: 827-839. 10.1002/jcb.20462.

    CAS  PubMed  Google Scholar 

  25. Liu CH, Hwang SM: Cytokine interactions in mesenchymal stem cells from cord blood. Cytokine. 2005, 32: 270-279. 10.1016/j.cyto.2005.11.003.

    CAS  PubMed  Google Scholar 

  26. Gang EJ, Jeong JA, Han S, Yan Q, Jeon CJ, Kim H: In vitro endothelial potential of human UC blood-derived mesenchymal stem cells. Cytotherapy. 2006, 8: 215-227. 10.1080/14653240600735933.

    CAS  PubMed  Google Scholar 

  27. De Ugarte DA, Alfonso Z, Zuk PA, Elbarbary A, Zhu M, Ashjian P, Benhaim P, Hedrick MH, Fraser JK: Differential expression of stem cell mobilization-associated molecules on multi-lineage cells from adipose tissue and bone marrow. Immunol Lett. 2003, 89: 267-270. 10.1016/S0165-2478(03)00108-1.

    CAS  PubMed  Google Scholar 

  28. Vaananen HK: Mesenchymal stem cells. Ann Med. 2005, 37: 469-479. 10.1080/07853890500371957.

    PubMed  Google Scholar 

  29. Knippenberg M, Helder MN, Doulabi BZ, Semeins CM, Wuisman PI, Klein-Nulend J: Adipose tissue-derived mesenchymal stem cells acquire bone cell-like responsiveness to fluid shear stress on osteogenic stimulation. Tissue Eng. 2005, 11: 1780-1788. 10.1089/ten.2005.11.1780.

    CAS  PubMed  Google Scholar 

  30. Portmann-Lanz CB, Schoeberlein A, Huber A, Sager R, Malek A, Holzgreve W, Surbek DV: Placental mesenchymal stem cells as potential autologous graft for pre- and perinatal neuroregeneration. Am J Obstet Gynecol. 2006, 194: 664-673. 10.1016/j.ajog.2006.01.101.

    CAS  PubMed  Google Scholar 

  31. Zhang X, Mitsuru A, Igura K, Takahashi K, Ichinose S, Yamaguchi S, Takahashi TA: Mesenchymal progenitor cells derived from chorionic villi of human placenta for cartilage tissue engineering. Biochem Biophys Res Commun. 2006, 340: 944-952. 10.1016/j.bbrc.2005.12.091.

    CAS  PubMed  Google Scholar 

  32. Shih DT, Lee DC, Chen SC, Tsai RY, Huang CT, Tsai CC, Shen EY, Chiu WT: Isolation and characterization of neurogenic mesenchymal stem cells in human scalp tissue. Stem Cells. 2005, 23: 1012-1020. 10.1634/stemcells.2004-0125.

    CAS  PubMed  Google Scholar 

  33. Kadivar M, Khatami S, Mortazavi Y, Shokrgozar MA, Taghikhani M, Soleimani M: In vitro cardiomyogenic potential of human umbilical vein-derived mesenchymal stem cells. Biochem Biophys Res Commun. 2006, 340: 639-647. 10.1016/j.bbrc.2005.12.047.

    CAS  PubMed  Google Scholar 

  34. Fu YS, Cheng YC, Lin MY, Cheng H, Chu PM, Chou SC, Shih YH, Ko MH, Sung MS: Conversion of human umbilical cord mesenchymal stem cells in Wharton's jelly to dopaminergic neurons in vitro: potential therapeutic application for Parkinsonism. Stem Cells. 2006, 24: 115-124. 10.1634/stemcells.2005-0053.

    PubMed  Google Scholar 

  35. Tondreau T, Meuleman N, Delforge A, Dejeneffe M, Leroy R, Massy M, Mortier C, Bron D, Lagneaux L: Mesenchymal stem cells derived from CD133-positive cells in mobilized peripheral blood and cord blood: proliferation, Oct4 expression, and plasticity. Stem Cells. 2005, 23: 1105-1112. 10.1634/stemcells.2004-0330.

    CAS  PubMed  Google Scholar 

  36. Jeong JA, Gang EJ, Hong SH, Hwang SH, Kim SW, Yang IH, Ahn C, Han H, Kim H: Rapid neural differentiation of human cord blood-derived mesenchymal stem cells. Neuroreport. 2004, 15: 1731-1734. 10.1097/01.wnr.0000134846.79002.5c.

    CAS  PubMed  Google Scholar 

  37. Kang XQ, Zang WJ, Bao LJ, Li DL, Song TS, Xu XL, Yu XJ: Fibroblast growth factor-4 and hepatocyte growth factor induce differentiation of human umbilical cord blood-derived mesenchymal stem cells into hepatocytes. World J Gastroenterol. 2005, 11: 7461-7465.

    CAS  PubMed  Google Scholar 

  38. Hong SH, Gang EJ, Jeong JA, Ahn C, Hwang SH, Yang IH, Park HK, Han H, Kim H: In vitro differentiation of human umbilical cord blood-derived mesenchymal stem cells into hepatocyte-like cells. Biochem Biophys Res Commun. 2005, 330: 1153-1161. 10.1016/j.bbrc.2005.03.086.

    CAS  PubMed  Google Scholar 

  39. Hutson EL, Boyer S, Genever PG: Rapid isolation, expansion, and differentiation of osteoprogenitors from full-term umbilical cord blood. Tissue Eng. 2005, 11: 1407-1420. 10.1089/ten.2005.11.1407.

    CAS  PubMed  Google Scholar 

  40. Liu J, Lu XF, Wan L, Li YP, Li SF, Zeng LY, Zeng YZ, Cheng LH, Lu YR, Cheng JQ: Suppression of human peripheral blood lymphocyte proliferation by immortalized mesenchymal stem cells derived from bone marrow of Banna Minipig inbred-line. Transplant Proc. 2004, 36: 3272-3275. 10.1016/j.transproceed.2004.11.090.

    CAS  PubMed  Google Scholar 

  41. Togel F, Hu Z, Weiss K, Isaac J, Lange C, Westenfelder C: Administered mesenchymal stem cells protect against ischemic acute renal failure through differentiation-independent mechanisms. Am J Physiol Renal Physiol. 2005, 289: F31-42. 10.1152/ajprenal.00007.2005.

    PubMed  Google Scholar 

  42. Kern S, Eichler H, Stoeve J, Kluter H, Bieback K: Comparative Analysis of Mesenchymal Stem Cells from Bone Marrow, Umbilical Cord Blood or Adipose Tissue. Stem Cells. 2006

    Google Scholar 

  43. Zhao Y, Wang H, Mazzone T: Identification of stem cells from human umbilical cord blood with embryonic and hematopoietic characteristics. Exp Cell Res. 2006, 312: 2454-2464.

    CAS  PubMed  Google Scholar 

  44. Kogler G, Sensken S, Airey JA, Trapp T, Muschen M, Feldhahn N, Liedtke S, Sorg RV, Fischer J, Rosenbaum C, Greschat S, Knipper A, Bender J, Degistirici O, Gao J, Caplan AI, Colletti EJ, Almeida-Porada G, Muller HW, Zanjani E, Wernet P: A new human somatic stem cell from placental cord blood with intrinsic pluripotent differentiation potential. J Exp Med. 2004, 200: 123-135. 10.1084/jem.20040440.

    PubMed Central  PubMed  Google Scholar 

  45. Zeng F, Chen MJ, Baldwin DA, Gong ZJ, Yan JB, Qian H, Wang J, Jiang X, Ren ZR, Sun D, Huang SZ: Multiorgan engraftment and differentiation of human cord blood CD34+ Lin- cells in goats assessed by gene expression profiling. Proc Natl Acad Sci U S A. 2006, 103: 7801-7806. 10.1073/pnas.0602646103.

    PubMed Central  CAS  PubMed  Google Scholar 

  46. McGuckin CP, Forraz N, Baradez MO, Navran S, Zhao J, Urban R, Tilton R, Denner L: Production of stem cells with embryonic characteristics from human umbilical cord blood. Cell Prolif. 2005, 38: 245-255. 10.1111/j.1365-2184.2005.00346.x.

    CAS  PubMed  Google Scholar 

  47. McGuckin CP, Forraz N, Allouard Q, Pettengell R: Umbilical cord blood stem cells can expand hematopoietic and neuroglial progenitors in vitro. Exp Cell Res. 2004, 295: 350-359. 10.1016/j.yexcr.2003.12.028.

    CAS  PubMed  Google Scholar 

  48. McGuckin CP, Forraz N, Pettengell R, Thompson A: Thrombopoietin, flt3-ligand and c-kit-ligand modulate HOX gene expression in expanding cord blood CD133 cells. Cell Prolif. 2004, 37: 295-306. 10.1111/j.1365-2184.2004.00313.x.

    CAS  PubMed  Google Scholar 

  49. Halbrecht J: Fresh and stored placental blood. Lancet. 1939, 2: 1263-10.1016/S0140-6736(00)74023-2.

    Google Scholar 

  50. Hassall O, Bedu-Addo G, Adarkwa M, Danso K, Bates I: Umbilical-cord blood for transfusion in children with severe anaemia in under-resourced countries. Lancet. 2003, 361: 678-679. 10.1016/S0140-6736(03)12565-2.

    CAS  PubMed  Google Scholar 

  51. Bhattacharya N: Placental umbilical cord whole blood transfusion: a safe and genuine blood substitute for patients of the under-resourced world at emergency. J Am Coll Surg. 2005, 200: 557-563. 10.1016/j.jamcollsurg.2004.12.007.

    PubMed  Google Scholar 

  52. Bhattacharya N: Placental umbilical cord whole blood transfusion to combat anemia in the background of tuberculosis and emaciation and its potential role as an immuno-adjuvant therapy for the under-resourced people of the world. Clin Exp Obstet Gynecol. 2006, 33: 99-104.

    CAS  PubMed  Google Scholar 

  53. Bhattacharya N: Placental umbilical cord blood transfusion: A novel method of treatment of patients with malaria in the background of anemia. Clin Exp Obstet Gynecol. 2006, 33: 39-43.

    CAS  PubMed  Google Scholar 

  54. Bhattacharya N: Placental umbilical cord whole blood transfusion to combat anemia in the background of advanced rheumatoid arthritis and emaciation and its potential role as immunoadjuvant therapy. Clin Exp Obstet Gynecol. 2006, 33: 28-33.

    CAS  PubMed  Google Scholar 

  55. Bhattacharya N: A preliminary study of placental umbilical cord whole blood transfusion in under resourced patients with malaria in the background of anaemia. Malar J. 2006, 5: 20-10.1186/1475-2875-5-20.

    PubMed Central  PubMed  Google Scholar 

  56. Bhattacharya N: A preliminary report of 123 units of placental umbilical cord whole blood transfusion in HIV-positive patients with anemia and emaciation. Clin Exp Obstet Gynecol. 2006, 33: 117-121.

    CAS  PubMed  Google Scholar 

  57. Bhattacharya N: Spontaneous transient rise of CD34 cells in peripheral blood after 72 hours in patients suffering from advanced malignancy with anemia: effect and prognostic implications of treatment with placental umbilical cord whole blood transfusion. Eur J Gynaecol Oncol. 2006, 27: 286-290.

    CAS  PubMed  Google Scholar 

  58. Ito K, Tanaka T, Tsutsumi N, Obata F, Kashiwagi N: Possible mechanisms of immunotherapy for maintaining pregnancy in recurrent spontaneous aborters: analysis of anti-idiotypic antibodies directed against autologous T-cell receptors. Hum Reprod. 1999, 14: 650-655. 10.1093/humrep/14.3.650.

    CAS  PubMed  Google Scholar 

  59. Smith JB, Cowchock FS, Lata JA, Hankinson BT: The number of cells used for immunotherapy of repeated spontaneous abortion influences pregnancy outcome. J Reprod Immunol. 1992, 22: 217-224. 10.1016/0165-0378(92)90044-5.

    CAS  PubMed  Google Scholar 

  60. Porter D, Levine JE: Graft-versus-host disease and graft-versus-leukemia after donor leukocyte infusion. Semin Hematol. 2006, 43: 53-61. 10.1053/j.seminhematol.2005.09.005.

    PubMed  Google Scholar 

  61. Szpakowski A, Malinowski A, Glowacka E, Wilczynski JR, Kolasa D, Dynski M, Tchorzewski H, Zeman K, Szpakowski M: [The influence of paternal lymphocyte immunization on the balance of Th1/Th2 type reactivity in women with unexplained recurrent spontaneous abortion]. Ginekol Pol. 2000, 71: 586-592.

    CAS  PubMed  Google Scholar 

  62. Hayakawa S, Karasaki-Suzuki M, Itoh T, Ishii M, Kanaeda T, Nagai N, Takahashi-Yamamoto N, Tochigi M, Chishima F, Fujii TK, Oyama J, Kitanaka S, Satoh K: Effects of paternal lymphocyte immunization on peripheral Th1/Th2 balance and TCR V beta and V gamma repertoire usage of patients with recurrent spontaneous abortions. Am J Reprod Immunol. 2000, 43: 107-115. 10.1111/j.8755-8920.2000.430207.x.

    CAS  PubMed  Google Scholar 

  63. Marleau AM, Sarvetnick N: T cell homeostasis in tolerance and immunity. J Leukoc Biol. 2005, 78: 575-584. 10.1189/jlb.0105050.

    CAS  PubMed  Google Scholar 

  64. Hickman SP, Turka LA: Homeostatic T cell proliferation as a barrier to T cell tolerance. Philos Trans R Soc Lond B Biol Sci. 2005, 360: 1713-1721. 10.1098/rstb.2005.1699.

    PubMed Central  CAS  PubMed  Google Scholar 

  65. Rosenberg SA, Dudley ME: Cancer regression in patients with metastatic melanoma after the transfer of autologous antitumor lymphocytes. Proc Natl Acad Sci U S A. 2004, 101 Suppl 2: 14639-14645. 10.1073/pnas.0405730101.

    PubMed  Google Scholar 

  66. Hess AD, Bright EC, Thoburn C, Vogelsang GB, Jones RJ, Kennedy MJ: Specificity of effector T lymphocytes in autologous graft-versus-host disease: role of the major histocompatibility complex class II invariant chain peptide. Blood. 1997, 89: 2203-2209.

    CAS  PubMed  Google Scholar 

  67. Miura Y, Thoburn CJ, Bright EC, Sommer M, Lefell S, Ueda M, Nakao S, Hess AD: Characterization of the T-cell repertoire in autologous graft-versus-host disease (GVHD): evidence for the involvement of antigen-driven T-cell response in the development of autologous GVHD. Blood. 2001, 98: 868-876. 10.1182/blood.V98.3.868.

    CAS  PubMed  Google Scholar 

  68. Maeda A, Schwarz A, Kernebeck K, Gross N, Aragane Y, Peritt D, Schwarz T: Intravenous infusion of syngeneic apoptotic cells by photopheresis induces antigen-specific regulatory T cells. J Immunol. 2005, 174: 5968-5976.

    CAS  PubMed  Google Scholar 

  69. Vendrame M, Gemma C, de Mesquita D, Collier L, Bickford PC, Sanberg CD, Sanberg PR, Pennypacker KR, Willing AE: Anti-inflammatory effects of human cord blood cells in a rat model of stroke. Stem Cells Dev. 2005, 14: 595-604. 10.1089/scd.2005.14.595.

    CAS  PubMed  Google Scholar 

  70. Lo YM, Lo ES, Watson N, Noakes L, Sargent IL, Thilaganathan B, Wainscoat JS: Two-way cell traffic between mother and fetus: biologic and clinical implications. Blood. 1996, 88: 4390-4395.

    CAS  PubMed  Google Scholar 

  71. Bianchi DW, Zickwolf GK, Weil GJ, Sylvester S, DeMaria MA: Male fetal progenitor cells persist in maternal blood for as long as 27 years postpartum. Proc Natl Acad Sci U S A. 1996, 93: 705-708. 10.1073/pnas.93.2.705.

    PubMed Central  CAS  PubMed  Google Scholar 

  72. O'Donoghue K, Chan J, de la Fuente J, Kennea N, Sandison A, Anderson JR, Roberts IA, Fisk NM: Microchimerism in female bone marrow and bone decades after fetal mesenchymal stem-cell trafficking in pregnancy. Lancet. 2004, 364: 179-182. 10.1016/S0140-6736(04)16631-2.

    PubMed  Google Scholar 

  73. Khosrotehrani K, Bianchi DW: Multi-lineage potential of fetal cells in maternal tissue: a legacy in reverse. J Cell Sci. 2005, 118: 1559-1563. 10.1242/jcs.02332.

    CAS  PubMed  Google Scholar 

  74. Johnson KL, Samura O, Nelson JL, McDonnell MWM, Bianchi DW: Significant fetal cell microchimerism in a nontransfused woman with hepatitis C: Evidence of long-term survival and expansion. Hepatology. 2002, 36: 1295-1297. 10.1053/jhep.2002.35622.

    PubMed  Google Scholar 

  75. Srivatsa B, Srivatsa S, Johnson KL, Samura O, Lee SL, Bianchi DW: Microchimerism of presumed fetal origin in thyroid specimens from women: a case-control study. Lancet. 2001, 358: 2034-2038. 10.1016/S0140-6736(01)07099-4.

    CAS  PubMed  Google Scholar 

  76. Khosrotehrani K, Johnson KL, Cha DH, Salomon RN, Bianchi DW: Transfer of fetal cells with multilineage potential to maternal tissue. Jama. 2004, 292: 75-80. 10.1001/jama.292.1.75.

    CAS  PubMed  Google Scholar 

  77. Khosrotehrani K, Stroh H, Bianchi DW, Johnson KL: Combined FISH and immunolabeling on paraffin-embedded tissue sections for the study of microchimerism. Biotechniques. 2003, 34: 242-244.

    CAS  PubMed  Google Scholar 

  78. Khosrotehrani K, Bianchi DW: Fetal cell microchimerism: helpful or harmful to the parous woman?. Curr Opin Obstet Gynecol. 2003, 15: 195-199. 10.1097/00001703-200304000-00014.

    PubMed  Google Scholar 

  79. Wang Y, Iwatani H, Ito T, Horimoto N, Yamato M, Matsui I, Imai E, Hori M: Fetal cells in mother rats contribute to the remodeling of liver and kidney after injury. Biochem Biophys Res Commun. 2004, 325: 961-967. 10.1016/j.bbrc.2004.10.105.

    CAS  PubMed  Google Scholar 

  80. Tan XW, Liao H, Sun L, Okabe M, Xiao ZC, Dawe GS: Fetal microchimerism in the maternal mouse brain: a novel population of fetal progenitor or stem cells able to cross the blood-brain barrier?. Stem Cells. 2005, 23: 1443-1452. 10.1634/stemcells.2004-0169.

    CAS  PubMed  Google Scholar 

  81. Taupin P: OTI-010 Osiris Therapeutics/JCR Pharmaceuticals. Curr Opin Investig Drugs. 2006, 7: 473-481.

    CAS  PubMed  Google Scholar 

  82. http://www.clinicaltrials.gov/ct/show/NCT00366145?order=7: .

  83. Kim SW, Han H, Chae GT, Lee SH, Bo S, Yoon JH, Lee YS, Lee KS, Park HK, Kang KS: Successful stem cell therapy using umbilical cord blood-derived multipotent stem cells for Buerger's disease and ischemic limb disease animal model. Stem Cells. 2006, 24: 1620-1626. 10.1634/stemcells.2005-0365.

    PubMed  Google Scholar 

  84. Kang KS, Kim SW, Oh YH, Yu JW, Kim KY, Park HK, Song CH, Han H: A 37-year-old spinal cord-injured female patient, transplanted of multipotent stem cells from human UC blood, with improved sensory perception and mobility, both functionally and morphologically: a case study. Cytotherapy. 2005, 7: 368-373. 10.1080/14653240500238160.

    PubMed  Google Scholar 

  85. Valbonesi M, Giannini G, Migliori F, Dalla Costa R, Dejana AM: Cord blood (CB) stem cells for wound repair. Preliminary report of 2 cases. Transfus Apher Sci. 2004, 30: 153-156. 10.1016/j.transci.2003.11.006.

    CAS  PubMed  Google Scholar 

  86. Corso A, Lorenzi A, Zappasodi P, Invernizzi R, Vanelli L, Lazzarino M: Early changes in bone marrow morphology induced by thalidomide in refractory myeloma patients. Haematologica. 2003, 88: 958-960.

    CAS  PubMed  Google Scholar 

  87. De Felice L, Tatarelli C, Mascolo MG, Gregorj C, Agostini F, Fiorini R, Gelmetti V, Pascale S, Padula F, Petrucci MT, Arcese W, Nervi C: Histone deacetylase inhibitor valproic acid enhances the cytokine-induced expansion of human hematopoietic stem cells. Cancer Res. 2005, 65: 1505-1513. 10.1158/0008-5472.CAN-04-3063.

    CAS  PubMed  Google Scholar 

  88. Araki H, Mahmud N, Milhem M, Nunez R, Xu M, Beam CA, Hoffman R: Expansion of human umbilical cord blood SCID-repopulating cells using chromatin-modifying agents. Exp Hematol. 2006, 34: 140-149. 10.1016/j.exphem.2005.10.002.

    CAS  PubMed  Google Scholar 

  89. Suzuki M, Harashima A, Okochi A, Yamamoto M, Nakamura S, Motoda R, Yamasaki F, Orita K: 5-Azacytidine supports the long-term repopulating activity of cord blood CD34(+) cells. Am J Hematol. 2004, 77: 313-315. 10.1002/ajh.20178.

    CAS  PubMed  Google Scholar 

  90. Shyu WC, Lee YJ, Liu DD, Lin SZ, Li H: Homing genes, cell therapy and stroke. Front Biosci. 2006, 11: 899-907. 10.2741/1846.

    CAS  PubMed  Google Scholar 

  91. Fricker SP, Anastassov V, Cox J, Darkes MC, Grujic O, Idzan SR, Labrecque J, Lau G, Mosi RM, Nelson KL, Qin L, Santucci Z, Wong RS: Characterization of the molecular pharmacology of AMD3100: a specific antagonist of the G-protein coupled chemokine receptor, CXCR4. Biochem Pharmacol. 2006, 72: 588-596. 10.1016/j.bcp.2006.05.010.

    CAS  PubMed  Google Scholar 

  92. Son BR, Marquez-Curtis LA, Kucia M, Wysoczynski M, Turner AR, Ratajczak J, Ratajczak MZ, Janowska-Wieczorek A: Migration of bone marrow and cord blood mesenchymal stem cells in vitro is regulated by stromal-derived factor-1-CXCR4 and hepatocyte growth factor-c-met axes and involves matrix metalloproteinases. Stem Cells. 2006, 24: 1254-1264. 10.1634/stemcells.2005-0271.

    CAS  PubMed  Google Scholar 

  93. Paczkowska E, Larysz B, Rzeuski R, Karbicka A, Jalowinski R, Kornacewicz-Jach Z, Ratajczak MZ, Machalinski B: Human hematopoietic stem/progenitor-enriched CD34(+) cells are mobilized into peripheral blood during stress related to ischemic stroke or acute myocardial infarction. Eur J Haematol. 2005, 75: 461-467. 10.1111/j.1600-0609.2005.00536.x.

    CAS  PubMed  Google Scholar 

  94. Williams RJ, Clarke E, Blair A, Evely R, Hale G, Waldmann H, Brookes S, Pamphilon DH: Impact on T-cell depletion and CD34+ cell recovery using humanised CD52 monoclonal antibody (CAMPATH-1H) in BM and PSBC collections; comparison with CAMPATH-1M and CAMPATH-1G. Cytotherapy. 2000, 2: 5-14. 10.1080/146532400539008.

    CAS  PubMed  Google Scholar 

  95. Gennery AR, Dickinson AM, Brigham K, Barge D, Spickett GP, Curtis A, Spencer V, Jackson A, Cavanagh G, Carter V, Palmer P, Flood TJ, Cant AJ, Abinun M: CAMPATH-1M T-cell depleted BMT for SCID: long-term follow-up of 19 children treated 1987-98 in a single center. Cytotherapy. 2001, 3: 221-232. 10.1080/146532401753174052.

    CAS  PubMed  Google Scholar 

  96. Klangsinsirikul P, Carter GI, Byrne JL, Hale G, Russell NH: Campath-1G causes rapid depletion of circulating host dendritic cells (DCs) before allogeneic transplantation but does not delay donor DC reconstitution. Blood. 2002, 99: 2586-2591. 10.1182/blood.V99.7.2586.

    CAS  PubMed  Google Scholar 

  97. Ratzinger G, Reagan JL, Heller G, Busam KJ, Young JW: Differential CD52 expression by distinct myeloid dendritic cell subsets: implications for alemtuzumab activity at the level of antigen presentation in allogeneic graft-host interactions in transplantation. Blood. 2003, 101: 1422-1429. 10.1182/blood-2002-04-1093.

    CAS  PubMed  Google Scholar 

  98. Staba SL, Escolar ML, Poe M, Kim Y, Martin PL, Szabolcs P, Allison-Thacker J, Wood S, Wenger DA, Rubinstein P, Hopwood JJ, Krivit W, Kurtzberg J: Cord-blood transplants from unrelated donors in patients with Hurler's syndrome. N Engl J Med. 2004, 350: 1960-1969. 10.1056/NEJMoa032613.

    CAS  PubMed  Google Scholar 

  99. Zhang C, Chen W, Xiao LL, Tan EX, Luo SK, Zheng D, Ye X, Li Z, Lu XL, Liu Y: [Allogeneic umbilical cord blood stem cell transplantation in Duchenne muscular dystrophy]. Zhonghua Yi Xue Za Zhi. 2005, 85: 522-525.

    PubMed  Google Scholar 

  100. Jaing TH, Hou JW, Chen SH, Huang IA, Wang CJ, Lee WI: Successful unrelated mismatched cord blood transplantation in a child with malignant infantile osteopetrosis. Pediatr Transplant. 2006, 10: 629-631. 10.1111/j.1399-3046.2006.00537.x.

    PubMed  Google Scholar 

  101. Tomizawa D, Aoki Y, Nagasawa M, Morio T, Kajiwara M, Sekine T, Shimizu N, Kato M, Yachie A, Mizutani S: Novel adopted immunotherapy for mixed chimerism after unrelated cord blood transplantation in Omenn syndrome. Eur J Haematol. 2005, 75: 441-444. 10.1111/j.1600-0609.2005.00535.x.

    PubMed  Google Scholar 

  102. Ooi J, Iseki T, Takahashi S, Tomonari A, Takasugi K, Uchiyama M, Konuma T, Fukuno K, Soda Y, Ohno N, Nagamura F, Uchimaru K, Tojo A, Asano S: Unrelated cord blood transplantation after myeloablative conditioning for adult patients with refractory anemia. Int J Hematol. 2005, 81: 424-427. 10.1532/IJH97.05014.

    PubMed  Google Scholar 

  103. Azuma E, Hirayama M, Bonno M, Iwamoto S, Kumamoto T, Kobayashi M, Komada Y, Taniguchi K, Nakano T, Kamiya H: Successful immunization following cord blood transplantation in a child with Diamond-Blackfan anemia. Pediatr Hematol Oncol. 2001, 18: 193-197. 10.1080/08880010151114822.

    CAS  PubMed  Google Scholar 

  104. Ishimura M, Ohga S, Nomura A, Toubo T, Morihana E, Saito Y, Nishio H, Ide M, Takada H, Hara T: Successful umbilical cord blood transplantation for severe chronic active Epstein-Barr virus infection after the double failure of hematopoietic stem cell transplantation. Am J Hematol. 2005, 80: 207-212. 10.1002/ajh.20430.

    PubMed  Google Scholar 

  105. Mullen CA, Thompson JN, Richard LA, Chan KW: Unrelated umbilical cord blood transplantation in infancy for mucopolysaccharidosis type IIB (Hunter syndrome) complicated by autoimmune hemolytic anemia. Bone Marrow Transplant. 2000, 25: 1093-1097. 10.1038/sj.bmt.1702397.

    CAS  PubMed  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas E Ichim.

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Riordan, N.H., Chan, K., Marleau, A.M. et al. Cord blood in regenerative medicine: do we need immune suppression?. J Transl Med 5, 8 (2007). https://doi.org/10.1186/1479-5876-5-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1479-5876-5-8

Keywords