Allogeneic hematopoietic stem cell transplantation (HSCT) has advanced to a common procedure for treating also older patients with malignancies and immunodeficiency disorders by redirecting the immune system. Unfortunately, cure is often hampered by relapse of the underlying disease, graft-versus-host disease, or severe opportunistic infections, which account for the majority of deaths after HSCT. Enhancing immune reconstitution is therefore an area of intensive research. An increasing variety of approaches has been explored preclinically and clinically: the application of cytokines, keratinocyte growth factor, growth hormone, cytotoxic lymphocytes, and mesenchymal stem cells or the blockade of sex hormones. New developments of allogeneic HSCT, for example, umbilical cord blood or haploidentical graft preparations leading to prolonged immunodeficiency, have further increased the need to improve immune reconstitution. Although a slow T-cell reconstitution is regarded as primarily responsible for deleterious infections with viruses and fungi, graft-versus-host disease, and relapse, the importance of innate immune cells for disease and infection control is currently being reevaluated. The groundwork has been prepared for the creation of individualized therapy partially based on genetic features of the underlying disease. We provide an update on selected issues of development in this fast evolving field; however, we do not claim completeness.

Allogeneic hematopoietic stem cell transplantation (HSCT) has become a common procedure for the therapy of hematologic malignancies and immune disorders. Donor bone marrow and mobilized peripheral blood stem cells are routinely used for the reconstitution of immune function in leukemia and lymphoma patients after radiation and/or chemotherapy. Recently, there have also been increasing attempts to treat solid tumors by immunotherapeutic approaches to induce a graft-versus-tumor effect, reviewed by Demirer et al.1 

Tempo of immune reconstitution

The reconstitution of different immune cell subsets after an allogeneic HSCT occurs at different tempos. After the conditioning regimen there is the “aplastic phase” (neutropenia) until neutrophils recover at approximately 14 days after peripheral blood stem cell transplantation, approximately 21 days after bone marrow transplantation, and approximately 30 days after umbilical cord blood transplantation (UCBT). The infections encountered during the aplastic phase do not differ from those found in other neutropenic patients and consist primarily of bacterial infections. The first 100 days after HSCT are characterized by cellular immune deficiencies with a reduced number of cytotoxic lymphocytes, natural killer (NK) cells of the innate immune system, and T cells of the specific immune system (Figure 1). This renders the patient especially susceptible to viral and fungal infections. Rapid recovery of NK cells after HSCT is based on an expansion of the cytokine-producing CD56bright NK-cell subset. This early expansion of CD56bright NK cells gradually declines but may persist for 1 year. Initial recovery of the T-cell compartment relies on peripheral expansion of memory T cells, driven by cytokines and the presence of alloreactive antigens, before the production of naive T cells in the thymus starts. This is especially true for CD4+ T cells that reconstitute later than CD8+ T cells and rely more on thymic production of naive T cells after HSCT, leading to an inversion of the CD4/CD8 ratio. T-cell receptor rearrangement excision DNA circles (TRECs) are used as markers for naive T-cell reconstitution occurring in the thymus. TREC levels remain low until 3 to 6 months after allogeneic HSCT. Increasing patient age is associated with thymic atrophy leading to decreased thymic output of naive T cells resulting in low TREC levels. The lack of naive T cells with a broad T-cell receptor (TCR) repertoire leads to an increased risk for opportunistic infections and leukemia relapse and is aggravated by graft-versus-host disease (GVHD; Figure 1).

Figure 1

Immunosuppressive agents hamper expansion/production of immune cells in lymphoid organs and periphery.

Figure 1

Immunosuppressive agents hamper expansion/production of immune cells in lymphoid organs and periphery.

Close modal

The reconstitution of the B-cell compartment representing the humoral immunity may take up to 2 years after HSCT. First, transitional CD19+CD24++CD38++ B cells occur in the peripheral blood, before their percentage progressively decreases, whereas the mature B-cell proportion rises.2  The lack of memory B cells, decreased levels of circulating immunoglobulins, impaired immunoglobulin class switching, and a loss of complexity in immunoglobulin gene rearrangement patterns leave allogeneic HSCT patients vulnerable to encapsulated bacteria such as Streptococcus pneumoniae and Haemophilus influenzae. The patients may obtain protective titers of pathogen-specific antibodies through vaccination. The optimal timing of immunization depends on the time elapsed since HSCT, the type of the graft, and the presence or absence of GVHD, as recently reviewed in extenso elsewhere.3 

Aspergillosis and CMV reactivation as major obstacles after HSCT

Notable for recent developments to improve T-cell immune reconstitution, we reviewed invasive aspergillosis and cytomegalovirus (CMV) reactivation, as it poses major obstacles after HSCT despite promising new approaches. A retrospective European Group for Blood and Marrow Transplantation (EBMT) study has determined the following risk factors for invasive aspergillosis: length of neutropenia, the status of the underlying disease, myeloablative conditioning, CMV disease, GVHD grades II-IV, and bone marrow or cord blood as a source of stem cells compared with peripheral blood stem cells.4  A risk model (low, intermediate, high) for progression of aspergillosis was developed based on this retrospective EBMT survey and needs to be validated in future trials on invasive aspergillosis. As the transfer of naive T cells is associated with delayed immune reconstitution, adoptive transfer of functionally active Aspergillus fumigatus–specific T cells might be a therapeutic option to restore immune effector mechanisms. Recently, a method for the rapid clinical scale generation of functionally active A fumigatus–specific T cells has been developed.5  Validation in clinical trials is pending. CMV reactivation may cause different diseases affecting lung, liver, gut, or central nervous system after allogeneic HSCT. CMV seropositivity remains a major risk factor for transplant-related mortality (TRM). Seropositive recipients obtaining a graft from a seropositive donor have improved survival and reduced TRM compared with those who received a transplant from a seronegative donor because of lack of transfer of CMV-specific T cells in the graft. Lack of CMV-specific T cells after the first episode of CMV reactivation is associated with multiple subsequent reactivations. This topic has recently been reviewed in depth.6  T cell–depleted (TCD) grafts have been introduced into allogeneic HSCT to reduce the risk of T cell–driven GVHD,7  at the expense of a higher incidence of fungal or viral infections after HSCT. However, a recent study demonstrated that TCD HSCT may be as successful as regular HSCT if antithymocyte globulin is excluded from the conditioning regimen, leading to durable engraftment with a low incidence of GVHD.8  TCD grafts supplemented with committed progenitor cells may enhance immune reconstitution of different immune cell subsets in the absence of GVHD.9  In addition, in TCD allogeneic HSCT, NK cells appear to play a major role in the control of viral reactivation. Chen et al10  have determined in a retrospective single-center study of 131 HLA-identical sibling donor/recipient pairs that additional activating killer immunoglobulin-like receptor (KIR) genes in the donor compared with the recipient were associated with lower TRM, increased survival, and lower incidence of CMV reactivation. Validation in a larger prospective clinical trial is warranted.

RIC regimens and increasing age of allogeneic hematopoietic stem cell transplant recipients

With the development of reduced-intensity conditioning (RIC) regimens in allogeneic transplantation, it has become possible to include disabled and older patients.11  The age limits are steadily rising, and some transplant physicians even question whether age limits still apply. The damage done to the thymus as production site for T cells is less severe in RIC compared with myeloablative conditioning.11  However, performing HSCT in older patients introduces new problems, as the lymphoid structures of the immune system also age.12  The immune system takes longer to rebuild and is more likely to be insufficient after transplantation. The older the patient, the more the T-cell recovery stems from expansion of memory T cells in the graft (Figure 1). The de novo production of naive T cells in the thymus with a broad T-cell receptor repertoire necessary for a powerful control of pathogens and of malignancies is limited. Therefore, the susceptibility to viral and fungal infections is elevated. The prediction of immune reconstitution in older patients is difficult, as the speed of immune system aging is very variable. Furthermore, the age not only of the recipient, but also of the donor is decisive as the graft of an older donor has decreased immune-reconstituting ability in the recipient.13  Rejuvenating the aging immune system is an area of rising importance considering the predicted increased percentage of aged patients in the population of developed countries.

Immune reconstitution after transplantation is also influenced by the occurrence of acute or chronic GVHD and by the immunosuppressive treatment selected (Figures 1 and 2). In general, potentially life-threatening GVHD leads to delays in immune reconstitution. Even so, the risk of GVHD is reduced if a TCD graft has been infused and/or antithymocyte globulin conditioning chosen; both approaches create, on the other hand, a deficit in T-cell immunity, leaving the recipient more susceptible to infections and/or leukemia relapse.14 

Figure 2

Strategies to enhance immune reconstitution.

Figure 2

Strategies to enhance immune reconstitution.

Close modal

There is consensus that the ideal donor for an allogeneic HSCT is a matched related donor. Unfortunately, a matched related donor is found in only approximately 25% of all cases; as an alternative, a matched unrelated donor (MUD) is found in only 30% (range, 10%-60% depending on the ethnic background). Luckily, the options of hematopoietic stem cell (HSC) transplant grafting have broadened. The numbers of haploidentical transplants from a sibling donor15  or the application of one or more umbilical cord blood (UCB) units for adult patients is steadily increasing. The advantage of a family donor (haploidentical or other) and, to an even greater extent, a cord blood donor is that they are, in general, immediately available, which is essential in high-risk leukemia with no time to search for an MUD. In addition, a second graft of the original family donor or from another family member is easily obtainable in the case of graft rejection, which is not true for a UCBT. However, this development poses new problems, which have not yet been sufficiently solved: for example, 40% of nonrelapse causes for posttransplantation mortality in high-risk leukemia patients after haploidentical transplantation are from CMV or Aspergillus infections.16 

Umbilical cord blood transplantation in adults

The number of adult patients treated with UCB is increasing as it is generally abundant, and no serious ethical dilemmas exist in terms of donor collection. UCB has become a valuable alternative to other hematopoietic stem cell sources. However, UCB often bears the disadvantage of insufficient cell numbers for immune reconstitution in adult patients because of the recipient's body weight, thymopoietic failure, or lack of antigen-experienced T cells, which leads to an impaired response toward cognate antigens and deficits in T-cell signal transduction mechanisms.17  In addition, UCB contains T regulatory (Treg) cells with more potent suppressor function than adult Treg cells. Many infections that afflict transplant patients are particularly frequent and more severe in the context of UCB transplantation. Therefore, current research attempts focus on transplanting more than a single UCB unit or on expanding selected subpopulations from UCB in vitro. UCB transplantation and its immune reconstitution features that apply to adult recipients has been extensively reviewed.18 

Haploidentical HSCT in adults

Because of HLA disparity, the risk of fatal GVHD in haploidentical transplantation is high. Different approaches to minimize this risk have been developed. The graft is depleted of T cells either by a positive CD34+ cell selection aiming for high numbers of hematopoietic progenitor cells including so-called “veto cells” to overcome major genetic barriers and enable rapid and durable engraftment (mega-dose concept)15  or by a positive CD3+/CD19+ cell depletion to ensure that engraftment-facilitating cells such as dendritic cells or NK cells are transplanted.19,20  Attempts in haploidentical transplantation to infuse T cell–repleted grafts after an intensive myeloablative conditioning and myelosuppressive treatment have led to incidences of GVHD grades II-IV of 78%.21,22  In T cell–depleted haploidentical transplantation, the graft-versus-leukemia (GVL) effect is provided primarily by NK cells if killer immunoglobulin ligand (KIR) incompatibility in the GVH direction is present. KIR ligand incompatibility led to improved outcomes in human haploidentical HSCT as shown by the Velardi group (Ruggeri et al).23  Other groups could not show the beneficial effects of NK cells,24  which are probably explained by the lack of T-cell depletion in that setting as the potency of the NK-cell effect appears to be diminished by the presence of T cells. KIRs are specific for allotypic determinants that are shared by different HLA class I alleles (referred to as KIR ligands). They may be activating or inhibitory. NK-cell infusion has the advantage of inducing no GVHD while maintaining GVL effects.23,25  Thus far, the features of NK-cell alloreactivity are not completely understood, so that still more research is necessary to ensure the selection of the optimal donor. Likewise, γδ T cells are thought to promote antileukemic effects and are capable of recognizing malignant cells through mechanisms that require no prior antigen exposure or priming.26  Current data on immune reconstitution of this T-cell subset are limited and it remains unclear whether the observed antileukemic effects are related to graft processing or donor factors.

The thymus involutes with age, leading to a gradual loss of newly produced naive T cells resulting in a restricted T-cell receptor repertoire skewed toward memory T cells. Thymic tissue has been considered a requirement for the generation of T cells with a broad antigen repertoire. The need for the thymus to repopulate the peripheral T-cell pool in adults has been acknowledged only since the advent of myeloablative therapy in HSCT and the treatment of AIDS patients (Figure 1). Different attempts to strengthen the endogenous immune system have been made by protecting the thymus either with keratinocyte growth factor or through sex hormone blockade. In addition, growth hormone might have a positive effect on the immune system, which would be useful in HSCT.

Application of keratinocyte growth factor, growth hormone, or androgen blockade

Keratinocyte growth factor (KGF; palifermin, Kepivance; Biovitrum, formerly Amgen) is a member of the fibroblast growth factor family that mediates epithelial cell proliferation and differentiation in a variety of tissues such as gut, skin, and thymus. Because of its protective effects on mucosa, KGF has entered the clinic for the prophylaxis of severe mucositis. The administration of KGF before conditioning and after transplantation resulted in enhanced thymopoiesis and peripheral T-cell numbers in different animal models. The KGF receptor is expressed on thymic epithelial cells (TECs) producing interleukin-7 (IL-7), suggesting preservation of IL-7 production is a possible mechanism.27  KGF's function as a protective and tropic factor for TECs ensures thymocyte proliferation and maturation. KGF-treated mice showed preservation of the thymic microenvironment, allowing normal thymopoiesis during acute GVHD after allogeneic HSCT.28  These findings are important as TECs are targets of the GVHD reaction. Enhancement of T-cell reconstitution after HSCT has also been demonstrated in an autologous nonhuman primate transplantation model for clinically relevant specific immune responses toward a T cell–dependent neoantigen.29  Unfortunately, in patients who underwent an allogeneic HSCT, the incidence of acute GVHD was not reduced and engraftment and early survival remained unaffected.30  In a long-term follow-up of these patients, no significant differences between the KGF-treated versus the untreated patients were observed regarding CMV or invasive fungal infection, chronic GVHD, or long-term survival.31 

The age-related atrophy of the thymus is accompanied by decreased thymic output. This coincides with increased amounts of circulating sex steroids from puberty. Surgical or pharmacologic sex hormone blockade has been demonstrated in murine models and in prostate cancer patients to induce a regeneration of the thymus leading to a restoration of peripheral naive T-cell phenotype (Figure 2). It also reversed a decline in B-cell production by increasing bone marrow cellularity. The Boyd group (Sutherland et al)32  reported enhanced immune system regeneration in humans after allogeneic or autologous transplantation. Although the differences in TREC production, T-cell repertoire regeneration, and amount of naive CD4+ and CD8+ T cells were minor in the early posttransplantation period, the authors found a significant survival advantage in the patients treated with luteinizing hormone–releasing hormone agonist (goserelin, Zoladex; AstraZeneca) compared with the control group in the autologous but not in the allogeneic setting. Despite enhanced T-cell responses in vitro, no increase in GVHD rates was observed in the luteinizing hormone–releasing hormone agonist–treated group. Overall, the patient numbers per group were small, and the diseases and conditioning regimens rather heterogeneous, so that no final conclusions should be drawn. Larger clinical trials are necessary. However, the combined use of KGF with sex hormone blockade in patients undergoing an allogeneic transplantation after myeloablative conditioning also led to a supranormal thymopoiesis and thymic output, a broad Vbeta repertoire, and decreased homeostatic T-cell proliferation.33  The right approach may thus be the combination of both attempts.

Neuroendocrine hormones have been shown to affect numerous immunologic responses after in vivo administration. The importance of endogenous growth hormone for the immune system has been demonstrated in growth hormone–deficient mice with a defective cellular immunity and thymus atrophy reversible when growth hormone was substituted.34  Recombinant human growth hormone (rhGH) led to an overall increase in thymocytes in a murine allogeneic transplantation model, but had no effect on the proportion of thymocyte subsets or TREC levels.35  Its beneficial effect on thymopoiesis might be because of the promotion of pluripotent HSCs or common lymphoid precursors homing to the thymus. This hypothesis is supported by observations in a murine fetal thymic organ culture model, in which an increase in thymocyte progenitors derived from bone marrow was reported.36  Furthermore, rhGH has been shown in HIV-infected patients to enhance thymic function and peripheral immune responses in humans (Figure 2).37  RhGH (Genotropin; Pfizer) is being evaluated in patients undergoing a UCBT to accelerate the immune reconstitution in a phase 1 trial (http://ClinicalTrials.gov identifier NCT00737113, “RhGH for accelerating immune reconstitution post unrelated cord blood transplant”).

Tyrosine kinase inhibition to facilitate thymic engraftment

The Mackall group (Krauss et al)38  demonstrated in a murine transplant model that tyrosine kinase inhibition by sunitinib facilitated thymic engraftment by modulating thymic niche accessibility (Figure 2). Sunitinib inhibits the important thymocyte growth factors c-kit and FLT-3.39  This effect was much more distinct compared with treatment with a monoclonal anti-CD25 antibody to open the CD4CD8 double-negative 3 (DN3) niche for thymocytes. Synergistic effects applying sunitinib and anti-CD25 antibody to open DN1-DN3 niches occurred. Currently, studies are ongoing in murine models to determine the effects of the combination on stem cell engraftment, durability of engraftment, and activity in minor mismatched models. The option to open up niches for naive T cells might be of particular interest to older patients as an increasing proportion of niches in peripheral immune tissues become occupied by terminally differentiated cells.40  Thus, the few naive T cells produced by the deteriorating thymus might not be able to take residence in the larger number of available niches.

Notch-based culture systems to promote T-cell reconstitution

Notch signaling is required for cellular differentiation processes. Four different Notch receptors (Notch 1-4) and 5 ligands (Jagged 1 and 2 and Delta-like 1, 3, and 4) have been identified. The inhibition of Notch 1 results in a partial inhibition of thymocyte differentiation41  and accumulation of precursors in the thymus. T-cell lineage committed precursor cells for adoptive cellular therapies can be obtained in large quantities applying Notch-based culture systems. They have been proven to be very effective in enhancing T-cell reconstitution and antitumor activity after allogeneic TCD HSCT in murine models (Figure 2).42,43  Currently, there is an observational trial ongoing in the United States to evaluate Notch-induced NK-cell activity in blood and bone marrow samples of cancer patients versus healthy controls (http://ClinicalTrials.gov identifier NCT00918658, “Study of NK cells in BM and blood samples from patients with hematologic cancer and from patients who do not have cancer”). Preliminary results of a phase 1 study evaluating the effect of applying Notch-based culture system for infusion of engraftment facilitating cells in human allogeneic HSCT were presented at the American Society of Hematology 2008 annual meeting.44 

Improved immune reconstitution after transplantation with exogenous cytokines

Application of IL-2.

IL-2 (Proleukin; Chiron Corporation) is a pleiotropic cytokine with a central role in immune responses. Its application has been evaluated as safe in small trials with a short follow-up. Lower doses of IL-2 appeared to enhance NK-cell numbers without significant effects on T cells.45  Recombinant IL-2 administered as consolidating immunotherapeutic agent early after HSCT at a time of minimal residual disease might reduce the relapse rate and increase the immunocompetence of these patients. This could be because of a lymphoid orientation of primitive CD34+CD105+ cells, which express high-affinity IL-2 receptors. Exogenous IL-2 might thus lead to an enhancement of the autologous GVL effect.46  The best evidence for IL-2 enhancement of donor T-cell function after allogeneic HSCT comes from reports of patients who failed to respond to donor lymphocyte infusion (DLI) for relapsed disease but achieved complete remission after treatment with IL-2.47,48  The role of IL-2, when added to DLI, needs to be further elucidated. At present, there is no general agreement on the optimum dosage or route of administration, and clinical trials in HSCT have led to conflicting results. The determination of optimum dosage schedules and methods of administration should enable a better assessment of the role of IL-2 in the treatment of these patients. Currently, there are several clinical trials just completed or still ongoing in the allogeneic HSCT setting, which it is hoped will define the use of IL-2 (eg, http://ClinicalTrials.gov identifiers NCT00003962, “IL-2 following BMT in treating patients with hematologic cancer,” and NCT00941928, “Haploidentical NK cells with epratuzumab [IL-2] for relapsed ALL”).

Application of IL-7.

IL-7 (CYT 99007; Cythris) plays a key role in the development of T cells and promotes thymic-dependent and -independent pathways.49,50  It is the most potent cytokine identified so far, promoting thymopoiesis by enhancing proliferation of immature thymic progenitors.51  Its impact on T-cell immune reconstitution after murine HSCT is still a controversial issue, depending on the characteristics of the transplantation, the model used, and the dose and length of IL-7 treatment. In some murine models, the application of IL-7 for 1 to 2 weeks early after HSCT led to an enhanced proliferation of immature thymic progenitors,51  however appeared to be transient.52  Alpdogan et al demonstrated in an allogeneic murine model that IL-7 increased the homeostatic proliferation of nonalloreactive T cells, but had no effect on alloreactive T cells and the development of GVHD.53  Others have found the opposite results in mice confirmed by IL-7Rα blockade experiments.54,55  In an autologous nonhuman primate model, treatment with IL-7 led to an increase of the CD4+ cell counts through peripheral expansion rather than de novo generation. In line with this observation were increased volumes of spleen and lymph nodes but not of the thymus in the IL-7–treated animals compared with the controls.56  Similarly, in a nontransplantation trial in human cancer patients, IL-7 application led to an increase of naive T cells with a broad TCR repertoire (Figure 2).57  Importantly, in the autologous nonhuman primate model with application of IL-7 for 6 to 10 weeks after transplantation, a GVHD-like gut infiltration by T cells was observed. Thus, this agent needs to be applied with caution in patients to avoid deleterious effects. Currently a phase 1 trial in allogeneic HSCT is ongoing (http://ClinicalTrials.gov identifier NCT00684008, “Safety Study of IL-7 in Recipients of a Hemopoietic Stem Cell Transplant Peripheral Blood Stem Cell Transplant”).

Application of IL-15.

IL-15 as member of the IL-2 cytokine family stimulates proliferation of T cells, NK cells, and B cells through a receptor consisting of IL-15Rα, IL-2Rβ, and the common γ chain. It has been found to be less toxic than IL-2 in murine transplantation models58  and shows potential as adjuvant for immunotherapy and tumor vaccination strategies.59,60  It has also been observed to improve T-cell engraftment in murine models.61  As the primary survival and growth factor for NK cells, it may promote GVL but not GVHD effects.62  IL-15 has been applied as adjuvant in vaccination, but thus far there have not been any trials with IL-15 in the HSCT setting.

Adoptive transfer of ex vivo–expanded immunomodulatory cells such as Treg cells, NK/Treg cells, donor-derived NK cells, and mesenchymal stem cells (MSCs) and adoptive transfer of allogeneic T cells specific for viral63,64  or tumor65,66  antigens appears promising to improve immune reconstitution after transplantation. In our review, we focus primarily on adoptive transfer of T cells67,68  and the cotransplantation of MSCs.

Adoptive transfer of viral and leukemic-specific T cells and use of regulatory T cells after HSCT

Increasing research activity is ongoing to enhance GVL-promoting cytotoxic T cells while suppressing the unwanted GVHD effects based on differential requirements for functionality, for example, tumor necrosis factor–related apoptosis-inducing ligand expression in murine models.69,70  Adoptive transfer of tumor-specific cytotoxic T cells is associated with different hurdles, as induction of unwanted GVHD renders DLI a potential toxic and insufficient controllable treatment regimen. To complicate issues further, the expansion of high-avidity tumor or leukemia-specific T cells in vitro is difficult and leads to a low efficacy in vivo based on, for example, exhaustion and loss of function in vitro.64  One promising approach to decrease the risk of GVHD is the genetic manipulation of T cells with herpes simplex virus thymidine kinase suicide gene as developed in Milan, Italy.71,72  If GVHD develops, transduced cells can be eliminated by ganciclovir treatment. However, difficulties remain because of low and transient transgene expression, unpredictable pairing of the exogenous and endogenous TCR chains, and poor survival and expansion potential of gene-modified effector T cells. A novel approach to increase functionality of the suicide gene–modified T cells with a central memory phenotype is therefore in vitro pretreatment with IL-7 and IL-15.73  Viral infection and reactivation contribute significantly to mortality and morbidity after allogeneic HSCT. There have been many attempts to selectively improve immune reconstitution of virus-specific T cells after HSCT. Recent approaches are the infusion of CMV, Epstein-Barr virus, or adenovirus-specific T cells selected from the donor and transferred to the recipient.74-78  In addition, CMV-specific T cells can be activated and expanded in vitro by stimulation with antigen-presenting cells loaded with specific proteins or peptides,79  as recently reviewed by our group.80  Essential for the impact of CMV reactivation on the HSCT outcome regarding triggering of GVHD and/or promoting GVL effects appears to be the quality of the T-cell response. Furthermore, there have been multiple approaches to selectively improve immune recovery by attenuating GVHD, for example, by expansion of T regulatory T cells (Treg cells), T-cell anergy induction, or selective reduction of alloreactivity as reviewed elsewhere.81-84 

GVL promotion by minor histocompatibility antigen–specific T cells

Increasing evidence suggests that CD8+ T cells recognizing minor histocompatibility antigens (mHags) on leukemic cells also play an essential role promoting GVL effects after allogeneic HSCT.85,86  These mHag-specific T cells may be generated applying a HLA/mHag multimer-guided approach, for example, coating autologous dendritic cells with HLA-A2/mHag complexes for in vitro generation of mHag-specific T cells.

Adoptive transfer of NK cells

The adoptive transfer of NK cells that do not promote GVHD but do promote GVL reactions is particularly useful in haploidentical HSCT where the risk of fatal GVHD is increased.87  In the case of a KIR mismatch, NK cells kill allogeneic cells that lack a class I MHC ligand for clonally distributed KIR as outlined in “Haploidentical HSCT in adults.”23,88 

Cotransplanting mesenchymal stem cells

The bone marrow microenvironment is increasingly being recognized as a potential target for immunotherapeutic attempts during or after an allogeneic HSCT. The results of applying mesenchymal stem cells (MSCs) are so far contradictory. Although there is consensus that they are pluripotent, have immunomodulatory abilities, and are a tool for managing or preventing GVHD as well as promoting clinical transplantation tolerance, it is not yet clear under which circumstances they support hematopoietic progenitor cell engraftment and show immunosuppressive properties in vivo.

In small case series and phase 1/2 clinical trials, hematopoietic engraftment has been accelerated when MSCs were cotransplanted in the haploidentical as well as in the UCB setting.89,90  Whether MSCs lead to a proliferation induction or inhibition of T cells in vitro appears to be dose dependent.87,91  When MSCs interact with NK cells, they inhibit IL-2–induced proliferation of resting NK cells, whereas they only partially affect the proliferation of activated NK cells in vitro.92  Several studies either as clinical case series or in animal models have advocated that MSCs may be transplanted across allogeneic barriers without eliciting an immune response. However, the Moretta group (Spaggiari et al92 ) demonstrated in vitro that IL-2–activated NK cells (but not freshly isolated NK cells) efficiently lyse autologous and allogeneic MSCs. A recent evaluation in a murine model questioned the immune privilege of MSCs but supported the hypothesis that they induce rejection in the allogeneic setting, which is followed by an immune memory potentially challenging long-term survival of allogeneic MSCs.93  Larger clinical trials with sufficient statistical power and an adequate scientific program are needed to further clarify the role of MSCs in allogeneic HSCT.

Unfortunately, many patients develop a relapse of their underlying disease after allogeneic HSCT. Although the majority of relapses occur in the periphery, isolated relapses at extramedullary sites as chloromas or plasmacytomas or at immunologically privileged sites including central nervous system and gonads are observed.9  Leukemia and other tumor cells escape from the immunologic attack by changes in antigen presentation and production of inhibitory cytokines.94  Recently, Vago et al95  demonstrated that the genomic loss of the mismatched HLA locus in leukemic cells is another major mechanism of in vivo escape from T-cell immune surveillance after haploidentical HSCT. They demonstrated that in 5 of 17 patients relapsing with AML, patient-specific HLA alleles could not be detected in bone marrow samples harvested at disease relapse. They showed that donor T cells were able to recognize the original HLA-heterozygous leukemia but not the mutant variant of the leukemia harvested at the time of relapse. The genomic rearrangement granted the disease an in vivo selective advantage in escaping from an established donor T-cell response (Figure 1). These findings support the hypothesis that major HLA antigens are principally responsible for the GVL effect. T-cell effector functions may be influenced by leukemic cells. Disappearance of high-avidity T cells in patients with active CML disease has been described as well as the reappearance of these cells in interferon-α responders.96 

Recently, monoclonal antibodies such as the anti-CD20 antibody rituximab (Mabthera; Roche) have been shown to be useful in several clinical trials for the treatment of GVHD or Epstein-Barr virus–associated lymphoproliferative diseases, and thus have a positive influence on immune reconstitution after HSCT.97  Furthermore, proteasome inhibitors such as bortezomib (Velcade; Ortho Biotech) and immunomodulatory drugs (so-called IMIDs) such as lenalidomide (Revlimid; Celgene), which are well established in cancer treatment because of their direct antitumor effects, also modulate the tumor microenvironment. They appear to suppress GVHD and potentially enhance GVL effects when used alone or in combination with DLI.98-100 

HSCT offers the opportunity for successfully treating malignancies and immune disorders. New therapies and better supportive care continue to improve outcome after HSCT. Nevertheless, delayed immune reconstitution and disease relapse remain major limitations to the widespread application of allogeneic transplantation. Understanding the causes of immune escape and hampered immune reconstitution provides the rationale to develop powerful immunotherapies. It is important to develop scores to identify those patients who would benefit most from immunomodulatory treatments during and after allogeneic transplantation.

The authors thank Torsten Steinbrunn for critical reading of the paper. The authors apologize to all authors whose work could not be cited because of length restrictions of this update on immune reconstitution after allogeneic HSCT.

Contribution: R.S. and H.E. wrote the paper.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Hermann Einsele, Department of Internal Medicine II, Julius-Maximilians-University, Oberdürrbacher Strasse 6, Bldg A3, Rm A3.-1.918, 97080 Würzburg, Germany; e-mail: einsele_h@klinik.uni-wuerzburg.de.

1
Demirer
 
T
Barkholt
 
L
Blaise
 
D
, et al. 
Transplantation of allogeneic hematopoietic stem cells: an emerging treatment modality for solid tumors.
Nat Clin Pract Oncol
2008
, vol. 
5
 
5
(pg. 
256
-
267
)
2
Marie-Cardine
 
A
Divay
 
F
Dutot
 
I
, et al. 
Transitional B cells in humans: characterization and insight from B lymphocyte reconstitution after hematopoietic stem cell transplantation.
Clin Immunol
2008
, vol. 
127
 
1
(pg. 
14
-
25
)
3
Tomblyn
 
M
Chiller
 
T
Einsele
 
H
, et al. 
Guidelines for preventing infectious complications among hematopoietic cell transplantation recipients: a global perspective.
Biol Blood Marrow Transplant
2009
, vol. 
15
 
10
(pg. 
1143
-
1238
)
4
Martino
 
R
Parody
 
R
Fukuda
 
T
, et al. 
Impact of the intensity of the pretransplantation conditioning regimen in patients with prior invasive aspergillosis undergoing allogeneic hematopoietic stem cell transplantation: a retrospective survey of the Infectious Diseases Working Party of the European Group for Blood and Marrow Transplantation.
Blood
2006
, vol. 
108
 
9
(pg. 
2928
-
2936
)
5
Tramsen
 
L
Koehl
 
U
Tonn
 
T
, et al. 
Clinical-scale generation of human anti-Aspergillus T cells for adoptive immunotherapy.
Bone Marrow Transplant
2009
, vol. 
43
 
1
(pg. 
13
-
19
)
6
Ljungman
 
P
CMV infections after hematopoietic stem cell transplantation.
Bone Marrow Transplant
2008
, vol. 
42
 
suppl 1
(pg. 
S70
-
S72
)
7
Gratama
 
JW
van Esser
 
JW
Lamers
 
CH
, et al. 
Tetramer-based quantification of cytomegalovirus (CMV)-specific CD8+ T lymphocytes in T-cell-depleted stem cell grafts and after transplantation may identify patients at risk for progressive CMV infection.
Blood
2001
, vol. 
98
 
5
(pg. 
1358
-
1364
)
8
Jakubowski
 
AA
Small
 
TN
Young
 
JW
, et al. 
T cell depleted stem-cell transplantation for adults with hematologic malignancies: sustained engraftment of HLA-matched related donor grafts without the use of antithymocyte globulin.
Blood
2007
, vol. 
110
 
13
(pg. 
4552
-
4559
)
9
Zakrzewski
 
JL
Holland
 
AM
van den Brink
 
MR
Adoptive precursor cell therapy to enhance immune reconstitution after hematopoietic stem cell transplantation.
J Mol Med
2007
, vol. 
85
 
8
(pg. 
837
-
843
)
10
Chen
 
C
Busson
 
M
Rocha
 
V
, et al. 
Activating KIR genes are associated with CMV reactivation and survival after non-T-cell depleted HLA-identical sibling bone marrow transplantation for malignant disorders.
Bone Marrow Transplant
2006
, vol. 
38
 
6
(pg. 
437
-
444
)
11
Jiménez
 
M
Ercilla
 
G
Martinez
 
C
Immune reconstitution after allogeneic stem cell transplantation with reduced-intensity conditioning regimens.
Leukemia
2007
, vol. 
21
 
8
(pg. 
1628
-
1637
)
12
Dorshkind
 
K
Montecino-Rodriguez
 
E
Signer
 
RA
The ageing immune system: is it ever too old to become young again?
Nat Rev Immunol
2009
, vol. 
9
 
1
(pg. 
57
-
62
)
13
Kollman
 
C
Howe
 
CW
Anasetti
 
C
, et al. 
Donor characteristics as risk factors in recipients after transplantation of bone marrow from unrelated donors: the effect of donor age.
Blood
2001
, vol. 
98
 
7
(pg. 
2043
-
2051
)
14
Fujimaki
 
K
Maruta
 
A
Yoshida
 
M
, et al. 
Immune reconstitution assessed during five years after allogeneic bone marrow transplantation.
Bone Marrow Transplant
2001
, vol. 
27
 
12
(pg. 
1275
-
1281
)
15
Aversa
 
F
Haploidentical haematopoietic stem cell transplantation for acute leukaemia in adults: experience in Europe and the United States.
Bone Marrow Transplant
2008
, vol. 
41
 
5
(pg. 
473
-
481
)
16
Aversa
 
F
Terenzi
 
A
Tabilio
 
A
, et al. 
Full haplotype-mismatched hematopoietic stem-cell transplantation: a phase II study in patients with acute leukemia at high risk of relapse.
J Clin Oncol
2005
, vol. 
23
 
15
(pg. 
3447
-
3454
)
17
Komanduri
 
KV
St John
 
LS
de Lima
 
M
, et al. 
Delayed immune reconstitution after cord blood transplantation is characterized by impaired thymopoiesis and late memory T-cell skewing.
Blood
2007
, vol. 
110
 
13
(pg. 
4543
-
4551
)
18
Brown
 
JA
Boussiotis
 
VA
Umbilical cord blood transplantation: basic biology and clinical challenges to immune reconstitution.
Clin Immunol
2008
, vol. 
127
 
3
(pg. 
286
-
297
)
19
Chen
 
X
Hale
 
GA
Barfield
 
R
, et al. 
Rapid immune reconstitution after a reduced-intensity conditioning regimen and a CD3-depleted haploidentical stem cell graft for paediatric refractory haematological malignancies.
Br J Haematol
2006
, vol. 
135
 
4
(pg. 
524
-
532
)
20
Bethge
 
WA
Haegele
 
M
Faul
 
C
, et al. 
Haploidentical allogeneic hematopoietic cell transplantation in adults with reduced-intensity conditioning and CD3/CD19 depletion: fast engraftment and low toxicity.
Exp Hematol
2006
, vol. 
34
 
12
(pg. 
1746
-
1752
)
21
Huang
 
XJ
Liu
 
DH
Liu
 
KY
, et al. 
Haploidentical hematopoietic stem cell transplantation without in vitro T-cell depletion for the treatment of hematological malignancies.
Bone Marrow Transplant
2006
, vol. 
38
 
4
(pg. 
291
-
297
)
22
Lu
 
DP
Dong
 
L
Wu
 
T
, et al. 
Conditioning including antithymocyte globulin followed by unmanipulated HLA-mismatched/haploidentical blood and marrow transplantation can achieve comparable outcomes with HLA-identical sibling transplantation.
Blood
2006
, vol. 
107
 
8
(pg. 
3065
-
3073
)
23
Ruggeri
 
L
Capanni
 
M
Urbani
 
E
, et al. 
Effectiveness of donor natural killer cell alloreactivity in mismatched hematopoietic transplants.
Science
2002
, vol. 
295
 
5562
(pg. 
2097
-
2100
)
24
Davies
 
SM
Ruggieri
 
L
DeFor
 
T
, et al. 
Evaluation of KIR ligand incompatibility in mismatched unrelated donor hematopoietic transplants: killer immunoglobulin-like receptor.
Blood
2002
, vol. 
100
 
10
(pg. 
3825
-
3827
)
25
Asai
 
O
Longo
 
DL
Tian
 
ZG
, et al. 
Suppression of graft-versus-host disease and amplification of graft-versus-tumor effects by activated natural killer cells after allogeneic bone marrow transplantation.
J Clin Invest
1998
, vol. 
101
 
9
(pg. 
1835
-
1842
)
26
Wilhelm
 
M
Kunzmann
 
V
Eckstein
 
S
, et al. 
Gammadelta T cells for immune therapy of patients with lymphoid malignancies.
Blood
2003
, vol. 
102
 
1
(pg. 
200
-
206
)
27
Min
 
D
Taylor
 
PA
Panoskaltsis-Mortari
 
A
, et al. 
Protection from thymic epithelial cell injury by keratinocyte growth factor: a new approach to improve thymic and peripheral T-cell reconstitution after bone marrow transplantation.
Blood
2002
, vol. 
99
 
12
(pg. 
4592
-
4600
)
28
Rossi
 
S
Blazar
 
BR
Farrell
 
CL
, et al. 
Keratinocyte growth factor preserves normal thymopoiesis and thymic microenvironment during experimental graft-versus-host disease.
Blood
2002
, vol. 
100
 
2
(pg. 
682
-
691
)
29
Seggewiss
 
R
Lore
 
K
Guenaga
 
FJ
, et al. 
Keratinocyte growth factor augments immune reconstitution after autologous hematopoietic progenitor cell transplantation in rhesus macaques.
Blood
2007
, vol. 
110
 
1
(pg. 
441
-
449
)
30
Langner
 
S
Staber
 
P
Schub
 
N
, et al. 
Palifermin reduces incidence and severity of oral mucositis in allogeneic stem-cell transplant recipients.
Bone Marrow Transplant
2008
, vol. 
42
 
4
(pg. 
275
-
279
)
31
Levine
 
JE
Blazar
 
BR
DeFor
 
T
Ferrara
 
JL
Weisdorf
 
DJ
Long-term follow-up of a phase I/II randomized, placebo-controlled trial of palifermin to prevent graft-versus-host disease (GVHD) after related donor allogeneic hematopoietic cell transplantation (HCT).
Biol Blood Marrow Transplant
2008
, vol. 
14
 
9
(pg. 
1017
-
1021
)
32
Sutherland
 
JS
Spyroglou
 
L
Muirhead
 
JL
, et al. 
Enhanced immune system regeneration in humans following allogeneic or autologous hemopoietic stem cell transplantation by temporary sex steroid blockade.
Clin Cancer Res
2008
, vol. 
14
 
4
(pg. 
1138
-
1149
)
33
Kelly
 
RM
Highfill
 
SL
Panoskaltsis-Mortari
 
A
, et al. 
Keratinocyte growth factor and androgen blockade work in concert to protect against conditioning regimen-induced thymic epithelial damage and enhance T-cell reconstitution after murine bone marrow transplantation.
Blood
2008
, vol. 
111
 
12
(pg. 
5734
-
5744
)
34
Murphy
 
WJ
Durum
 
SK
Anver
 
MR
Longo
 
DL
Immunologic and hematologic effects of neuroendocrine hormones: studies on DW/J dwarf mice.
J Immunol
1992
, vol. 
148
 
12
(pg. 
3799
-
3805
)
35
Chen
 
BJ
Cui
 
X
Sempowski
 
GD
Chao
 
NJ
Growth hormone accelerates immune recovery following allogeneic T-cell-depleted bone marrow transplantation in mice.
Exp Hematol
2003
, vol. 
31
 
10
(pg. 
953
-
958
)
36
Knyszynski
 
A
Adler-Kunin
 
S
Globerson
 
A
Effects of growth hormone on thymocyte development from progenitor cells in the bone marrow.
Brain Behav Immun
1992
, vol. 
6
 
4
(pg. 
327
-
340
)
37
Napolitano
 
LA
Schmidt
 
D
Gotway
 
MB
, et al. 
Growth hormone enhances thymic function in HIV-1-infected adults.
J Clin Invest
2008
, vol. 
118
 
3
(pg. 
1085
-
1098
)
38
Krauss
 
AC
Guimond
 
M
Dobre
 
S
Mackall
 
CL
Tyrosine kinase inhibition with sunitinib facilitates thymic engraftment by modulating thymic niche accessibility [abstract].
Blood
2008
, vol. 
112
  
Abstract 72
39
Faivre
 
S
Demetri
 
G
Sargent
 
W
Raymond
 
E
Molecular basis for sunitinib efficacy and future clinical development.
Nat Rev Drug Discov
2007
, vol. 
6
 
9
(pg. 
734
-
745
)
40
Nikolich-Zugich
 
J
Ageing and life-long maintenance of T-cell subsets in the face of latent persistent infections.
Nat Rev Immunol
2008
, vol. 
8
 
7
(pg. 
512
-
522
)
41
Radtke
 
F
Wilson
 
A
Mancini
 
SJ
MacDonald
 
HR
Notch regulation of lymphocyte development and function.
Nat Immunol
2004
, vol. 
5
 
3
(pg. 
247
-
253
)
42
Schmitt
 
TM
Zuniga-Pflucker
 
JC
Induction of T cell development from hematopoietic progenitor cells by delta-like-1 in vitro.
Immunity
2002
, vol. 
17
 
6
(pg. 
749
-
756
)
43
Zakrzewski
 
JL
Kochman
 
AA
Lu
 
SX
, et al. 
Adoptive transfer of T-cell precursors enhances T-cell reconstitution after allogeneic hematopoietic stem cell transplantation.
Nat Med
2006
, vol. 
12
 
9
(pg. 
1039
-
1047
)
44
Delaney
 
C
Brashem-Stein
 
C
Voorhies
 
H
, et al. 
Notch-mediated expansion of human cord blood progenitor cells results in rapid myeloid reconstitution in vivo following myeloablative cord blood transplantation [abstract].
Blood
2008
, vol. 
112
  
Abstract 212
45
Soiffer
 
RJ
Murray
 
C
Cochran
 
K
, et al. 
Clinical and immunologic effects of prolonged infusion of low-dose recombinant interleukin-2 after autologous and T-cell-depleted allogeneic bone marrow transplantation.
Blood
1992
, vol. 
79
 
2
(pg. 
517
-
526
)
46
Perillo
 
A
Pierelli
 
L
Battaglia
 
A
, et al. 
Administration of low-dose interleukin-2 plus G-CSF/EPO early after autologous PBSC transplantation: effects on immune recovery and NK activity in a prospective study in women with breast and ovarian cancer.
Bone Marrow Transplant
2002
, vol. 
30
 
9
(pg. 
571
-
578
)
47
Slavin
 
S
Naparstek
 
E
Nagler
 
A
, et al. 
Allogeneic cell therapy with donor peripheral blood cells and recombinant human interleukin-2 to treat leukemia relapse after allogeneic bone marrow transplantation.
Blood
1996
, vol. 
87
 
6
(pg. 
2195
-
2204
)
48
Nadal
 
E
Fowler
 
A
Kanfer
 
E
Apperley
 
J
Goldman
 
J
Dazzi
 
F
Adjuvant interleukin-2 therapy for patients refractory to donor lymphocyte infusions.
Exp Hematol
2004
, vol. 
32
 
2
(pg. 
218
-
223
)
49
Mackall
 
CL
Fry
 
TJ
Bare
 
C
Morgan
 
P
Galbraith
 
A
Gress
 
RE
IL-7 increases both thymic-dependent and thymic-independent T-cell regeneration after bone marrow transplantation.
Blood
2001
, vol. 
97
 
5
(pg. 
1491
-
1497
)
50
Fry
 
TJ
Connick
 
E
Falloon
 
J
, et al. 
A potential role for interleukin-7 in T-cell homeostasis.
Blood
2001
, vol. 
97
 
10
(pg. 
2983
-
2990
)
51
Bolotin
 
E
Smogorzewska
 
M
Smith
 
S
Widmer
 
M
Weinberg
 
K
Enhancement of thymopoiesis after bone marrow transplant by in vivo interleukin-7.
Blood
1996
, vol. 
88
 
5
(pg. 
1887
-
1894
)
52
Alpdogan
 
O
Schmaltz
 
C
Muriglan
 
SJ
, et al. 
Administration of interleukin-7 after allogeneic bone marrow transplantation improves immune reconstitution without aggravating graft-versus-host disease.
Blood
2001
, vol. 
98
 
7
(pg. 
2256
-
2265
)
53
Alpdogan
 
O
Muriglan
 
SJ
Eng
 
JM
, et al. 
IL-7 enhances peripheral T cell reconstitution after allogeneic hematopoietic stem cell transplantation.
J Clin Invest
2003
, vol. 
112
 
7
(pg. 
1095
-
1107
)
54
Chung
 
B
Dudl
 
EP
Min
 
D
Barsky
 
L
Smiley
 
N
Weinberg
 
KI
Prevention of graft-versus-host disease by anti IL-7Ralpha antibody.
Blood
2007
, vol. 
110
 
8
(pg. 
2803
-
2810
)
55
Sinha
 
ML
Fry
 
TJ
Fowler
 
DH
Miller
 
G
Mackall
 
CL
Interleukin 7 worsens graft-versus-host disease.
Blood
2002
, vol. 
100
 
7
(pg. 
2642
-
2649
)
56
Storek
 
J
Gillespy
 
T
Lu
 
H
, et al. 
Interleukin-7 improves CD4 T-cell reconstitution after autologous CD34 cell transplantation in monkeys.
Blood
2003
, vol. 
101
 
10
(pg. 
4209
-
4218
)
57
Sportès
 
C
Hakim
 
FT
Memon
 
SA
, et al. 
Administration of rhIL-7 in humans increases in vivo TCR repertoire diversity by preferential expansion of naive T cell subsets.
J Exp Med
2008
, vol. 
205
 
7
(pg. 
1701
-
1714
)
58
Katsanis
 
E
Xu
 
Z
Panoskaltsis-Mortari
 
A
Weisdorf
 
DJ
Widmer
 
MB
Blazar
 
BR
IL-15 administration following syngeneic bone marrow transplantation prolongs survival of lymphoma bearing mice.
Transplantation
1996
, vol. 
62
 
6
(pg. 
872
-
875
)
59
Klebanoff
 
CA
Finkelstein
 
SE
Surman
 
DR
, et al. 
IL-15 enhances the in vivo antitumor activity of tumor-reactive CD8+ T cells.
Proc Natl Acad Sci U S A
2004
, vol. 
101
 
7
(pg. 
1969
-
1974
)
60
Rubinstein
 
MP
Kadima
 
AN
Salem
 
ML
Nguyen
 
CL
Gillanders
 
WE
Cole
 
DJ
Systemic administration of IL-15 augments the antigen-specific primary CD8+ T cell response following vaccination with peptide-pulsed dendritic cells.
J Immunol
2002
, vol. 
169
 
9
(pg. 
4928
-
4935
)
61
Sun
 
A
Wei
 
H
Sun
 
R
Xiao
 
W
Yang
 
Y
Tian
 
Z
Human interleukin-15 improves engraftment of human T cells in NOD-SCID mice.
Clin Vaccine Immunol
2006
, vol. 
13
 
2
(pg. 
227
-
234
)
62
Lin
 
SJ
Cheng
 
PJ
Yan
 
DC
Lee
 
PT
Hsaio
 
HS
Effect of interleukin-15 on alloreactivity in umbilical cord blood.
Transpl Immunol
2006
, vol. 
16
 
2
(pg. 
112
-
116
)
63
Peggs
 
KS
Verfuerth
 
S
Pizzey
 
A
, et al. 
Adoptive cellular therapy for early cytomegalovirus infection after allogeneic stem-cell transplantation with virus-specific T-cell lines.
Lancet
2003
, vol. 
362
 
9393
(pg. 
1375
-
1377
)
64
Comoli
 
P
Basso
 
S
Zecca
 
M
, et al. 
Preemptive therapy of EBV-related lymphoproliferative disease after pediatric haploidentical stem cell transplantation.
Am J Transplant
2007
, vol. 
7
 
6
(pg. 
1648
-
1655
)
65
Thorne
 
SH
Negrin
 
RS
Contag
 
CH
Synergistic antitumor effects of immune cell-viral biotherapy.
Science
2006
, vol. 
311
 
5768
(pg. 
1780
-
1784
)
66
Einsele
 
H
Hamprecht
 
K
Immunotherapy of cytomegalovirus infection after stem-cell transplantation: a new option?
Lancet
2003
, vol. 
362
 
9393
(pg. 
1343
-
1344
)
67
Riddell
 
SR
Bleakley
 
M
Nishida
 
T
Berger
 
C
Warren
 
EH
Adoptive transfer of allogeneic antigen-specific T cells.
Biol Blood Marrow Transplant
2006
, vol. 
12
 
1 suppl 1
(pg. 
9
-
12
)
68
Kapp
 
M
Rasche
 
L
Einsele
 
H
Grigoleit
 
GU
Cellular therapy to control tumor progression.
Curr Opin Hematol
2009
, vol. 
16
 
6
(pg. 
437
-
443
)
69
Schmaltz
 
C
Alpdogan
 
O
Kappel
 
BJ
, et al. 
T cells require TRAIL for optimal graft-versus-tumor activity.
Nat Med
2002
, vol. 
8
 
12
(pg. 
1433
-
1437
)
70
Zheng
 
H
Matte-Martone
 
C
Li
 
H
, et al. 
Effector memory CD4+ T cells mediate graft-versus-leukemia without inducing graft-versus-host disease.
Blood
2008
, vol. 
111
 
4
(pg. 
2476
-
2484
)
71
Marktel
 
S
Magnani
 
Z
Ciceri
 
F
, et al. 
Immunologic potential of donor lymphocytes expressing a suicide gene for early immune reconstitution after hematopoietic T-cell-depleted stem cell transplantation.
Blood
2003
, vol. 
101
 
4
(pg. 
1290
-
1298
)
72
Ciceri
 
F
Bonini
 
C
Gallo-Stampino
 
C
Bordignon
 
C
Modulation of GvHD by suicide-gene transduced donor T lymphocytes: clinical applications in mismatched transplantation.
Cytotherapy
2005
, vol. 
7
 
2
(pg. 
144
-
149
)
73
Kaneko
 
S
Mastaglio
 
S
Bondanza
 
A
, et al. 
IL-7 and IL-15 allow the generation of suicide gene-modified alloreactive self-renewing central memory human T lymphocytes.
Blood
2009
, vol. 
113
 
5
(pg. 
1006
-
1015
)
74
Walter
 
EA
Greenberg
 
PD
Gilbert
 
MJ
, et al. 
Reconstitution of cellular immunity against cytomegalovirus in recipients of allogeneic bone marrow by transfer of T-cell clones from the donor.
N Engl J Med
1995
, vol. 
333
 
16
(pg. 
1038
-
1044
)
75
Rauser
 
G
Einsele
 
H
Sinzger
 
C
, et al. 
Rapid generation of combined CMV-specific CD4+ and CD8+ T-cell lines for adoptive transfer into recipients of allogeneic stem cell transplants.
Blood
2004
, vol. 
103
 
9
(pg. 
3565
-
3572
)
76
Cobbold
 
M
Khan
 
N
Pourgheysari
 
B
, et al. 
Adoptive transfer of cytomegalovirus-specific CTL to stem cell transplant patients after selection by HLA-peptide tetramers.
J Exp Med
2005
, vol. 
202
 
3
(pg. 
379
-
386
)
77
Leen
 
AM
Christin
 
A
Myers
 
GD
, et al. 
Cytotoxic T lymphocyte therapy with donor T cells prevents and treats adenovirus and Epstein-Barr virus infections after haploidentical and matched unrelated stem cell transplantation.
Blood
2009
, vol. 
114
 
19
(pg. 
4283
-
4292
)
78
Berger
 
C
Turtle
 
CJ
Jensen
 
MC
Riddell
 
SR
Adoptive transfer of virus-specific and tumor-specific T cell immunity.
Curr Opin Immunol
2009
, vol. 
21
 
2
(pg. 
224
-
232
)
79
Kleihauer
 
A
Grigoleit
 
U
Hebart
 
H
, et al. 
Ex vivo generation of human cytomegalovirus-specific cytotoxic T cells by peptide-pulsed dendritic cells.
Br J Haematol
2001
, vol. 
113
 
1
(pg. 
231
-
239
)
80
Einsele
 
H
Kapp
 
M
Grigoleit
 
GU
CMV-specific T cell therapy.
Blood Cells Mol Dis
2008
, vol. 
40
 
1
(pg. 
71
-
75
)
81
Nguyen
 
VH
Shashidhar
 
S
Chang
 
DS
, et al. 
The impact of regulatory T cells on T-cell immunity following hematopoietic cell transplantation.
Blood
2008
, vol. 
111
 
2
(pg. 
945
-
953
)
82
Riley
 
JL
June
 
CH
Blazar
 
BR
Human T regulatory cell therapy: take a billion or so and call me in the morning.
Immunity
2009
, vol. 
30
 
5
(pg. 
656
-
665
)
83
Davies
 
JK
Yuk
 
D
Nadler
 
LM
Guinan
 
EC
Induction of alloanergy in human donor T cells without loss of pathogen or tumor immunity.
Transplantation
2008
, vol. 
86
 
6
(pg. 
854
-
864
)
84
Amrolia
 
PJ
Mucioli-Casadei
 
G
Huls
 
H
, et al. 
Add-back of allodepleted donor T cells to improve immune reconstitution after haplo-identical stem cell transplantation.
Cytotherapy
2005
, vol. 
7
 
2
(pg. 
116
-
125
)
85
Norde
 
WJ
Overes
 
IM
Maas
 
F
, et al. 
Myeloid leukemic progenitor cells can be specifically targeted by minor histocompatibility antigen LRH-1-reactive cytotoxic T cells.
Blood
2009
, vol. 
113
 
10
(pg. 
2312
-
2323
)
86
van Bergen
 
CA
Kester
 
MG
Jedema
 
I
, et al. 
Multiple myeloma-reactive T cells recognize an activation-induced minor histocompatibility antigen encoded by the ATP-dependent interferon-responsive (ADIR) gene.
Blood
2007
, vol. 
109
 
9
(pg. 
4089
-
4096
)
87
Passweg
 
JR
Tichelli
 
A
Meyer-Monard
 
S
, et al. 
Purified donor NK-lymphocyte infusion to consolidate engraftment after haploidentical stem cell transplantation.
Leukemia
2004
, vol. 
18
 
11
(pg. 
1835
-
1838
)
88
Ruggeri
 
L
Capanni
 
M
Casucci
 
M
, et al. 
Role of natural killer cell alloreactivity in HLA-mismatched hematopoietic stem cell transplantation.
Blood
1999
, vol. 
94
 
1
(pg. 
333
-
339
)
89
Le Blanc
 
K
Samuelsson
 
H
Gustafsson
 
B
, et al. 
Transplantation of mesenchymal stem cells to enhance engraftment of hematopoietic stem cells.
Leukemia
2007
, vol. 
21
 
8
(pg. 
1733
-
1738
)
90
Macmillan
 
ML
Blazar
 
BR
DeFor
 
TE
Wagner
 
JE
Transplantation of ex-vivo culture-expanded parental haploidentical mesenchymal stem cells to promote engraftment in pediatric recipients of unrelated donor umbilical cord blood: results of a phase I-II clinical trial.
Bone Marrow Transplant
2009
, vol. 
43
 
6
(pg. 
447
-
454
)
91
Sudres
 
M
Norol
 
F
Trenado
 
A
, et al. 
Bone marrow mesenchymal stem cells suppress lymphocyte proliferation in vitro but fail to prevent graft-versus-host disease in mice.
J Immunol
2006
, vol. 
176
 
12
(pg. 
7761
-
7767
)
92
Spaggiari
 
GM
Capobianco
 
A
Becchetti
 
S
Mingari
 
MC
Moretta
 
L
Mesenchymal stem cell-natural killer cell interactions: evidence that activated NK cells are capable of killing MSCs, whereas MSCs can inhibit IL-2-induced NK-cell proliferation.
Blood
2006
, vol. 
107
 
4
(pg. 
1484
-
1490
)
93
Zangi
 
L
Beilhack
 
A
Negrin
 
R
Margalit
 
R
Reisner
 
Y
Mesenchymal stem cells (MSC) are not intrinsically immune privileged: direct demonstration in TCR transgenic mice and by imaging of allogeneic luciferase+ MSC in immune competent vs immune deficient mice [abstract].
Blood
2008
, vol. 
112
  
Abstract 549
94
Woiciechowsky
 
A
Regn
 
S
Kolb
 
HJ
Roskrow
 
M
Leukemic dendritic cells generated in the presence of FLT3 ligand have the capacity to stimulate an autologous leukemia-specific cytotoxic T cell response from patients with acute myeloid leukemia.
Leukemia
2001
, vol. 
15
 
2
(pg. 
246
-
255
)
95
Vago
 
L
Perna
 
SK
Zanussi
 
M
, et al. 
Loss of mismatched HLA in leukemia after stem-cell transplantation.
N Engl J Med
2009
, vol. 
361
 
5
(pg. 
478
-
488
)
96
Molldrem
 
JJ
Lee
 
PP
Kant
 
S
, et al. 
Chronic myelogenous leukemia shapes host immunity by selective deletion of high-avidity leukemia-specific T cells.
J Clin Invest
2003
, vol. 
111
 
5
(pg. 
639
-
647
)
97
Ratanatharathorn
 
V
Pavletic
 
S
Uberti
 
JP
Clinical applications of rituximab in allogeneic stem cell transplantation: anti-tumor and immunomodulatory effects.
Cancer Treat Rev
2009
, vol. 
35
 
8
(pg. 
653
-
661
)
98
Koreth
 
J
Alyea
 
EP
Murphy
 
WJ
Welniak
 
LA
Proteasome inhibition and allogeneic hematopoietic stem cell transplantation: a review.
Biol Blood Marrow Transplant
2009
, vol. 
15
 
12
(pg. 
1502
-
1512
)
99
Yong
 
AS
Keyvanfar
 
K
Hensel
 
N
, et al. 
Primitive quiescent CD34+ cells in chronic myeloid leukemia are targeted by in vitro expanded natural killer cells, which are functionally enhanced by bortezomib.
Blood
2009
, vol. 
113
 
4
(pg. 
875
-
882
)
100
Kröger
 
N
Badbaran
 
A
Lioznov
 
M
, et al. 
Post-transplant immunotherapy with donor-lymphocyte infusion and novel agents to upgrade partial into complete and molecular remission in allografted patients with multiple myeloma.
Exp Hematol
2009
, vol. 
37
 
7
(pg. 
791
-
798
)
Sign in via your Institution