A characteristic feature of leukemia cells is a blockade of differentiation at a distinct stage in cellular maturation. In the 1970s and 1980s, studies demonstrating the capabilities of certain chemicals to induce differentiation of hematopoietic cell lines fostered the concept of treating leukemia by forcing malignant cells to undergo terminal differentiation instead of killing them through cytotoxicity. The first promising reports on this notion prompted a review article on this subject by us 25 years ago. In this review, we revisit this interesting field of study and report the progress achieved in the course of nearly 3 decades. The best proof of principle for differentiation therapy has been the treatment of acute promyelocytic leukemia with all-trans retinoic acid. Attempts to emulate this success with other nuclear hormone ligands such as vitamin D compounds and PPARγ agonists or different classes of substances such as hematopoietic cytokines or compounds affecting the epigenetic landscape have not been successful on a broad scale. However, a multitude of studies demonstrating partial progress and improvements and, finally, the new powerful possibilities of forward and reverse engineering of differentiation pathways by manipulation of transcription factors support the continued enthusiasm for differentiation therapy of leukemia in the future.

A characteristic abnormality of leukemia cells is that they are blocked at an early stage of their development and fail to differentiate into functional mature cells. During the 1970s and 1980s, several scientific achievements popularized the strategy of inducing malignant cells to overcome their block of differentiation and enter the apoptotic pathways as an elegant alternative to killing cancer cells by cytotoxic therapies. This intervention could theoretically limit exposure to unwanted side effects of cytotoxic chemotherapy, and more importantly, improve complete remission and cure rates. Pioneering reports included studies demonstrating the differentiating capability of dimethylsulfoxide (DMSO) on erythropoiesis,1  efforts to elucidate substances to control the differentiation of myeloid leukemia,2  and the first evidence of the differentiating properties of retinoic acid.3,4  The initial promising preclinical results of this approach prompted a review article by us 25 years ago concerning the possibilities and therapeutic implications of differentiation induction in leukemia.5  At that stage, cell line models for in vitro differentiation experiments were described. Substances such as phorbol diesters, teleocidins, polar planar drugs, cytokines, retinoids, and vitamin D metabolites showed dramatic potential to differentiate cell lines such as HL-60, KG-1, ML-3, or K562 in vitro and fueled the hope of developing a new approach to treat cancers by overcoming their blocked differentiation. Today, 25 years later, we discuss the progress and the clinical achievements of this therapeutic approach

Poster child of success: complete remissions of APL by differentiation therapy

The potential for differentiating therapy to improve cure rates in leukemia is exemplified by the development of all-trans retinoic acid (ATRA) for the targeted treatment of acute promyelocytic leukemia (APL). One of the most remarkable results of initial in vitro experiments was achieved in differentiating HL-60 cells with ATRA, which produced terminal differentiation in 90% of cells with 10−6 M retinoic acid.3  Investigators soon realized that ATRA was specifically effective in APL cells carrying a typical chromosomal translocation between chromosomes 15 and 17 [t(15;17)(q22;q21)]6  but not in other leukemias.4  The first APL patients treated with ATRA in the early 1980s achieved encouraging remissions by the new therapy.7-10 

The first clinical trial of ATRA reported 16 newly diagnosed, and 8 anthracycline refractory patients who were induced with single-agent ATRA. Complete hematologic remission was induced in all patients; more than 90% of samples from these individuals demonstrated in vitro evidence of blast differentiation.10  This seminal trial changed the management and prognosis of APL, and introduced a paradigm for success of cell differentiation therapy. Subsequently, APL therapy was improved stepwise through elucidation of the most effective combination regimen of ATRA with cytotoxic chemotherapy.11-16 

Shortly after ATRA therapy became standard for induction of newly diagnosed patients with APL, strategies for overcoming ATRA resistance became necessary. Arsenic trioxide (ATO) emerged as an option for relapsed patients capable of producing high rates of molecular remissions and resensitizing patients to the differentiating effect of ATRA.17-22  Today, the wealth of clinical information and improvements of targeted differentiation therapy with ATRA have led to the development of guidelines including highly effective induction and consolidation regimen including ATRA/ATO and anthracycline-based chemotherapy, achieving complete remission rates of up to 90% to 95%.23 

The breakthrough in clinical oncology achieved by differentiation therapy with ATRA also sparked intensive research into mechanisms underlying the observed successes. ATRA is a ligand to retinoic acid receptors (RARs), which comprise a family of transcription factors that bind to retinoic acid response elements (RAREs) and regulate granulocytic differentiation24  (Figure 1A). The APL cells of 95% of patients have a characteristic chromosomal translocation between chromosomes 15 and 17 [t(15;17)(q22;q21)]6  leading to a fusion of the genes promyelocytic leukemia (PML) and the retinoic acid receptor alpha (RARα).25,26  The classical model describes that this fusion product acts as a dominant negative of RARα by forming homodimers, recruiting corepressors, and inhibiting expression of target genes necessary for granulocytic differentiation by binding to RAREs.27  However, because this repression of RARα target genes by PML-RARα is associated with the recruitment of DNA and histone-modifying enzymes such as histone deacetylases (HDACs), histone methyltransferases,28,29  and DNA methyltransferases,30  the leukemogenic activity of this fusion product is mediated by mechanisms beyond the simple repression of RARα-regulated genes. It leads to a highly repressive chromatin environment, which affects multiple pathways. Furthermore, the PML-RARα product may also inhibit the normal function of the PML protein as a tumor suppressor and therefore acts as a dominant negative against both proteins.31 

Figure 1

Ligands of nuclear hormone receptors. Depicts schematically the molecular mechanisms of differentiation induction of nuclear hormone receptor agonists. (A) All-trans retinoic acid (ATRA) for acute promyelocytic leukemia (APL). The characteristic chromosomal translocation t(15;17)(q22;q21) in APL leads to the production of a fusion protein between the PML protein and the retinoic acid receptor alpha (RARα). This fusion product is able to form homodimers and disrupt normal RARα signaling. It binds to retinoic acid response elements (RAREs) of target genes and recruits corepressors (Co-Rs) such as histone deacetylases (HDACs) and DNA methyltransferases (DNMTs), and sequesters retinoic X receptor (RXR) and the wild-type PML protein (PML), which finally leads to repression of genes necessary for granulocytic differentiation. Treatment with pharmacological concentrations of ATRA causes a conformational change of the PML-RARα fusion product leading to the release of the corepressors, recruitment of histone acetyl transferases (HATs), and relief of transcriptional repression. This causes the treated APL cells to undergo terminal granulocytic differentiation and finally apoptosis. (B) Biologically active vitamin D [1,25(OH)2D3] binds to the nuclear vitamin D receptor (VDR), which heterodimerizes with retinoic X receptor (RXR). This activated complex binds to vitamin D response elements (VDREs) in the promoter regions of genes inducing cell-cycle arrest, apoptosis, and differentiation in cancer cells. Furthermore, it leads to an up-regulation of the antimicrobial peptide cathelicidin (CAMP) in myeloid cells. (C) Thiazolidinediones (TZDs) bind to peroxisome proliferator activated receptor gamma (PPARγ). This activated complex acts as a transcription factor by heterodimerizing with (RXR) and binding to PPARγ-responsive elements in the promoter regions of target genes involved in cell-cycle arrest, apoptosis, growth inhibition, and differentiation of cancer cells.

Figure 1

Ligands of nuclear hormone receptors. Depicts schematically the molecular mechanisms of differentiation induction of nuclear hormone receptor agonists. (A) All-trans retinoic acid (ATRA) for acute promyelocytic leukemia (APL). The characteristic chromosomal translocation t(15;17)(q22;q21) in APL leads to the production of a fusion protein between the PML protein and the retinoic acid receptor alpha (RARα). This fusion product is able to form homodimers and disrupt normal RARα signaling. It binds to retinoic acid response elements (RAREs) of target genes and recruits corepressors (Co-Rs) such as histone deacetylases (HDACs) and DNA methyltransferases (DNMTs), and sequesters retinoic X receptor (RXR) and the wild-type PML protein (PML), which finally leads to repression of genes necessary for granulocytic differentiation. Treatment with pharmacological concentrations of ATRA causes a conformational change of the PML-RARα fusion product leading to the release of the corepressors, recruitment of histone acetyl transferases (HATs), and relief of transcriptional repression. This causes the treated APL cells to undergo terminal granulocytic differentiation and finally apoptosis. (B) Biologically active vitamin D [1,25(OH)2D3] binds to the nuclear vitamin D receptor (VDR), which heterodimerizes with retinoic X receptor (RXR). This activated complex binds to vitamin D response elements (VDREs) in the promoter regions of genes inducing cell-cycle arrest, apoptosis, and differentiation in cancer cells. Furthermore, it leads to an up-regulation of the antimicrobial peptide cathelicidin (CAMP) in myeloid cells. (C) Thiazolidinediones (TZDs) bind to peroxisome proliferator activated receptor gamma (PPARγ). This activated complex acts as a transcription factor by heterodimerizing with (RXR) and binding to PPARγ-responsive elements in the promoter regions of target genes involved in cell-cycle arrest, apoptosis, growth inhibition, and differentiation of cancer cells.

Close modal

Pharmacological concentrations of ATRA lead to a conformation change of the multifunctional molecule complex around PML-RARα. Corepessors are released, normal regulation of RARα-responsive genes is restored, and hence terminal differentiation of APL cells is induced.32  This is supported by several recent microarray and proteomic studies, which have identified hundreds of genes that are differentially regulated in the ATRA-induced differentiation of APL cells33,34  including down-regulation of c-myc35  and up-regulation of C/EBPϵ,36  as well as PU.1.37  Furthermore, genes governing increased protein synthesis such as PDCD4 or RTP801 are down-regulated in APL cells during ATRA exposure,38,39  whereas genes associated with protein degradation are up-regulated by ATRA, leading to a degradation of the PML-RARα fusion product.40,41 

The degradation of PML-RARα may also represent an intersection, where the mechanisms of action of ATO converge on those of ATRA. Initial endeavors to elucidate the molecular activities of ATO showed a dual mode of action. At low concentrations, ATO induced partial morphologic differentiation in APL cells, whereas at high concentrations, apoptosis induction predominated. Both effects were associated with a degradation of PML-RARα.42  ATO induced PML-RARα and PML degradation is associated with enhanced sumoylation of these proteins,43  indicating that effects achieved by ATO in APL may be attributed to an increased targeting of the PML moiety versus RARα targeting of ATRA. A comparison of ATO- and ATRA-induced gene expression and proteome profiles showed that both compounds regulate similar cellular factors. However, ATO's emphasis was on altering the proteome and inducing apoptosis as opposed to predominant regulation of gene expression and differentiation by ATRA.33  Interestingly, differentiation of APL cells by either ATRA or ATO is highly dependent on stimulation by myeloid growth factors as evidenced by experiments using growth factor–neutralizing antibodies.44  In addition, both substances can induce a side effect known as APL differentiation syndrome, suggesting some overlapping mechanisms of action.23  Sharing similar pathways but exhibiting different foci of action, the 2 compounds complement each other.

Taken together, therapy of APL with ATRA and ATO is to date the most successful example of differentiation therapy and its scientific history serves as a template for subsequent development of similar treatments in other leukemias and cancers.

Vitamin D compounds

Concurrent with the first observations of the differentiating action of retinoids on selected myeloid cell lines, similar promising effects were also demonstrated for the physiologically active form of vitamin D, 1,25 dihydroxy vitamin D3 [1,25(OH)2D3]. This seco-steroid potently differentiated cells of the myeloid lineage in vitro and ex vivo,5,45  which led to early clinical trials to test the ability of 1,25(OH)2D3 to treat myelodysplastic syndromes (MDSs) or acute myeloid leukemia (AML).46,47  Although 1,25(OH)2D3 induced partial differentiation of hematopoietic blast cells in some of these patients, clinical improvements such as blood counts or survival were modest. The same was true for studies that assessed the effect of 1,25(OH)2D3 in combination with other agents used to treat either MDS or AML.48-50  Initial studies also suggested that 1,25(OH)2D3 had cancer-preventive properties in prostate and colon cancers51-54  and exerted positive antitumor effects by regulation of proliferation, apoptosis, and angiogenesis.55-58 

As for the mechanism of action, 1,25(OH)2D3 binds and activates the vitamin D receptor (VDR), which heterodimerizes with the retinoic X receptor and binds to vitamin D–responsive elements (VDREs) in the promoter regions of target genes59-61  (Figure 1B). One of the main mechanisms of antiproliferative and differentiating action of 1,25(OH)2D3 is the induction of cell-cycle arrest by regulation of genes such as p21 and p27, which harbor VDREs.62-64  However, despite greater insight into the fundamental signaling of vitamin D compounds, the responses in individual tumor types are very heterogeneous so that a common understanding of how it mediates its antiproliferative activity is yet to be established.51 

A limiting factor in the clinical application of 1,25(OH)2D3 is hypercalcemia.65  Pharmacokinetic and pharmacodynamic optimization has shown that this can be mitigated by intermittent, high doses of 1,25(OH)2D3 and therapeutic support with glucosteroids.66  In addition, a large number of vitamin D analogs and vitamin D receptor modulators have been synthesized in the hope of gaining greater antitumor effects while deceasing their hypercalcemic activity.51  Paricalcitol or doxercalciferol have partly achieved this goal. However, summarizing clinical trials testing them against hematologic malignancies such as MDS or AML, the results are still rather disappointing.67,68 

Another important potential activity of 1,25(OH)2D3 is the ability to transcriptionally induce the expression of the antimicrobial peptide cathelicidin (CAMP).69,70  The ability of vitamin D compounds to enhance the expression of antimicrobial peptides has created an exciting new field of research to elucidate the role of vitamin D compounds as boosters of the immune system to fight infectious diseases.71  However, regarding hematologic malignancies, vitamin D compounds have not had a major impact on their clinical management at this stage.

PPARγ receptor ligands

Peroxisome proliferator activated receptor gamma (PPARγ) is also a member of the superfamily of nuclear hormone receptors (NHRs) with an important role in the regulation of fatty acid metabolism and a variety of endogenous ligands such as 15-deoxy-delta 12, 14-prostaglandin J2, and polyunsaturated fatty acids.72  In 1995, investigators discovered that PPARγ was the molecular target of thiazolidinediones (TZDs),73,74  a group of synthetic substances widely used in the treatment of type 2 diabetes. Reports soon noted that PPARγ agonists had the ability to prevent either the development or growth of tumors and to induce differentiation using various model systems.75-78  In the wake of this, PPARγ agonists were tested in vitro and in animal tumor models against a plethora of tumors including colon, breast, and prostate cancers and acute leukemias. The PPARγ agonists displayed various antiproliferative and differentiating potency in many cancer types,75  especially in models of myeloid, lymphoid, and chronic myelogenous leukemia.79-83 

PPARγ as a nuclear hormone receptor functions as a transcription factor (Figure 1C). It heterodimerizes with the retinoid X receptor (RXR) and then binds to PPARγ response elements in the promoter regions of target genes.84  Antitumor effects of PPARγ agonists have been associated with exit from the cell cycle by up-regulation of inhibitors of cyclin-dependent kinases (CDKs) such as p18, p21, or p27 associated with reduced phosphorylation of the retinoblastoma protein (Rb).85  The apoptosis pathways are also modestly activated by down-regulation of the antiapoptotic Bcl-2 protein and up-regulation of the proapoptotic proteins Bax and Bad.86,87  Furthermore, PPARγ agonists increased the expression of other tumor suppressor genes such as PTEN and BRCA1.88,89  Several lines of evidence also suggest “off-target” effects of PPARγ agonists, which are not dependent on the PPARγ receptor. Experiments in PPARγ−/− murine cells demonstrated that PPARγ ligands continued to have inhibitory activity on inflammation pathways in macrophages and to cause cell-cycle arrest independently of PPARγ.90-92 

Paradoxically, PPARγ has proneoplastic activity in certain contexts.93  For example, mice carrying a heterozygous mutation of the adenomatous polyposis coli gene (APC) (Apc+/Min mice) demonstrated increased tumor number and size when treated with TZDs.94  This obervation was supported in other animal tumors models of spontaneous colon cancers or genetically induced mammary gland tumors, which showed increased incidence or growth of tumors with TZD treatment or activation of PPARγ.95,96  Nevertheless, the majority of basic research reports attribute antitumor effects to activation of PPARγ and prompted clinical trials of these substances in several human malignancies. Therefore, initial observations of adipocyte maturation in liposarcoma patients78  could not be confirmed in a later trial with 9 patients, which could neither demonstrate terminal differentiation of tumor cells nor achievement of any clinical benefit through treatment with TZDs.97 

Given evidence that TZDs could cause terminal differentiation of breast cancer cell lines, clinical breast cancer studies were undertaken. Among 22 patients with metastatic, refractory breast cancer, no clinical benefit was observed.98  In a study of 38 women with newly diagnosed, early stage breast cancer, neoadjuvant TZD treatment showed that the postoperative pathologic tissue did not show evidence of enhanced differentiation of tumor cells.99  Clinical trials in prostate, colon, lung, and thyroid cancer also did not find a meaningful benefit of TZD therapy.100-105 

In summary, the clinical effects achieved by PPARγ agonists have not been resounding and probably do not merit a primary role for them in cancer therapy. However, their low toxicity profile and the observation that they can achieve additive or synergistic effects combined with other anticancer agents such as retinoids, histone deacetylase (HDAC) inhibitors, or TRAIL ligands86,100-104  make them possible candidates to include in adjuvant or combination therapy trials.106 

The differentiation of hematopoietic stem cells into mature blood cells is intricately controlled by an array of hematopoietic cytokines.107  The discovery of growth factors involved in the regulation of hematopoiesis dates back to the 1960s when the first models for in vitro culturing of hematopoietic progenitors were developed.108,109  Soon, these factors were isolated and produced as recombinant substances for clinical application.107,110  The physiological signal transduction of hematopoietic cytokines consists of their binding to their specific receptor and subsequent activation of downstream pathways such as protein kinase C (PKC), mitogen-activated protein kinase (MAPK), Janus kinases (JAKs), Src kinase pathways, and STATs. Activation of these pathways, in turn, induces transcriptional activation and repression of genes governing the differentiation and lineage commitment of hematopoietic progenitors.

The quickly increasing knowledge about the mechanisms of action of cytokines in the differentiation of hematopoietic progenitors also induced high hopes of using these factors in the treatment of leukemia.111  Some of the hematopoietic leukemia cell lines of myeloid origin such as K562, U937, HL-60, CS-1, KG-1, MUTZ-3, or ex vivo AML or chronic myeloid leukemia (CML) blasts were modestly permissive to induction of in vitro differentiation by EPO, G-CSF, GM-CSF, IL-4, IL-6, SCF, or synergistic combinations of several cytokines.111-115  Molecular mechanisms of cytokine-induced leukemic differentiation were also elucidated. Several prominent proto-oncogenes such as c-myb, c-myc, c-fos, and Ets family transcription factors such as ets-1, fli-1, and TEL2 were found to be differentially regulated upon cytokine-induced differentiation of leukemic cells.116-120  However, when translated to a clinical setting, the approach to treat leukemia by trying to differentiate the malignant cells with cytokines remained rather modest. In single case reports, AML patients have been shown to enter complete hematologic remissions upon treatment with either G-CSF or GM-CSF.121-125  However, in larger trials assessing the therapeutic impact in terms of differentiation of leukemia cells in AML or MDS, hematopoietic cytokines have had limited success.126 

A niche for hematopoietic cytokines in differentiation therapy exists in the treatment of congenital neutropenia disorder. In this disease, several decades of progress have identified pertubations of the G-CSF receptor signaling as one of the underlying causes, and the administration of G-CSF to these patients has overcome a block of myeloid differentiation leading to a substantial prolongation of their survival.127  In summary, the concept of using cytokines as a differentiation treatment against leukemia has been disappointing. Nevertheless, cytokines have gained many other important domains of action such as supportive therapy during cytotoxic chemotherapy or treatment of congenital neutropenia.

In the last few years, increasing information has emerged about the transcription factors governing the differentiation of hematopoietic cells. The function of several of these transcription factors is frequently disrupted in leukemia cells.128  Examples include the CCAAT/enhancer binding protein alpha (C/EBPα) in AML or the paired box gene 5 (Pax5) in ALL (Figure 2).

Figure 2

Block of differentiation by disruption of hematopoietic transcription factors in myeloid and lymphoid leukemia. Depicts the molecular mechanisms leading to disruption of the transcription factor CCAAT/enhancer binding protein alpha (C/EBPα) in myeloid leukemia and the paired box gene 5 (Pax5) in lymphocytic leukemia. Disruption of these transcription factors blocks hematopoietic cells in their early stages of differentiation. Treatment with substances altering the epigenetic settings such as histone deacetylase inhibitors or demethylating agents may partly overcome the block in differentiation. Forced expression of the normal counterpart of the disrupted transcription factor can often re-establish differentiation. In the future, synthesis of small molecules specifically targeting transcription factors or nuclear reprogramming and gene therapy may provide useful tools for correcting defective differentiation in hematologic malignancies.

Figure 2

Block of differentiation by disruption of hematopoietic transcription factors in myeloid and lymphoid leukemia. Depicts the molecular mechanisms leading to disruption of the transcription factor CCAAT/enhancer binding protein alpha (C/EBPα) in myeloid leukemia and the paired box gene 5 (Pax5) in lymphocytic leukemia. Disruption of these transcription factors blocks hematopoietic cells in their early stages of differentiation. Treatment with substances altering the epigenetic settings such as histone deacetylase inhibitors or demethylating agents may partly overcome the block in differentiation. Forced expression of the normal counterpart of the disrupted transcription factor can often re-establish differentiation. In the future, synthesis of small molecules specifically targeting transcription factors or nuclear reprogramming and gene therapy may provide useful tools for correcting defective differentiation in hematologic malignancies.

Close modal

In physiological hematopoiesis, C/EBPα is an important factor for the development of granulocytes129,130  and targeted disruption of C/EBPα leads to a selective block in granulocyte maturation.131  The normal function of C/EBPα is disturbed by a variety of events in AML.132  Several groups have detected spontaneous mutations of the C/EBPα gene, and a summary of these studies revealed that C/EBPα was mutated in 11% of 1565 AML samples.133-135  Mutations in the amino end of the gene produce a dominant negative form of the protein and those in the carboxyl end disturb its DNA binding and heterodimerization ability with other family members.133  Often, some AML cells harbor both kinds of mutations, each affecting a different allele. Interestingly, the presence of CEBPα mutations in AML cells confers a favorable prognosis.135  Other mechanisms of perturbation of C/EBPα include down-regulation of its mRNA levels by AML1-ETO in t(8;21)-positive AMLs,136  hypermethylation of its promoter region, silencing its expression,137  posttranslational inhibition by phosphorylation,138  or dysregulated proteasomal degradation.139  Regardless of the cause, the dysregulation of normal C/EBPα signaling more or less culminates in an arrest of granulocytic differentiation and accumulation of immature blasts in AML. As expected, forced expression of C/EBPα in leukemia cells can reverse the block of differentiation and induce terminal maturation of AML cells.140,141 

A comparable candidate in ALL is the B-lineage–specific transcription factor Pax5. In a recent study analyzing the genomes of 242 cases of childhood ALLs with high-density SNP arrays, Pax5 was the most commonly altered gene, either by deletion or by mutation.142  Both the Downing group (Mullighan et al)142  and we143  showed that Pax5 mutant proteins or Pax5 fusion products displayed reduced transcriptional activity or dominant negative effects. The physiological role of Pax5 lies in the regulation of B-cell gene expression during development of lymphocytes. Its expression is up-regulated beginning at the pro-B progenitor stage and is maintained at high levels throughout B-cell development until it is silenced during the final transition to plasma cells.144  Studies of hematopoiesis in Pax5-deficient mice demonstrated arrested differentiation at very early stages of hematopoietic development before the emergence of B-cell progenitors or B cell–specific gene expression.145,146  Experimental manipulation of Pax5 has shown its effects on the development of lymphocytes.147  In an experiment to challenge the paradigm of irreversible lineage commitment during physiological hematopoiesis, a murine model with a conditional Pax5 deletion allowed mature B cells from peripheral lymphoid organs to dedifferentiate back to early uncommitted progenitors in the bone marrow, which rescued T lymphopoiesis in the thymus of T cell–deficient mice.148  Furthermore, ectopic expression of Pax5 in normal hematopoietic progenitors or myeloid malignancies has yielded heterogeneous results concerning its function and potential to reverse a malignant phenotype.149-151  The mechanisms of leukemogenesis caused by perturbation of Pax5 in lymphoid malignancies is unclear. The leukemogenic effects of Pax5 fusion products and the consequences of re-establishing physiological levels of functional Pax5 in ALL cells with Pax5 deletions is currently being investigated by us and other laboratories.

Given the powerful potential of hematopoietic transcription factors such as C/EBPα and Pax5 to control normal hematopoiesis and their frequent dysregulation in hematologic malignancies, they are obvious therapeutic targets. However, development of drugs that precisely manipulate specific transcription factors has not been achieved successfully. Although, hope lies in large drug screens for small molecules or small interfering RNA molecules that specifically modify endogenous expression or function of transcription factors.135 

As mentioned before, signaling of important hematopoietic transcription factors can be impaired by an aberrant epigenomic environment.137,152  Epigenetic silencing of tumor suppressors and transcription factors governing differentiation occurs in many cancers. In recent years, many compelling mechanisms have been identified, which epigentically alter the genome. The 2 most characterized have led to the development of clinically applicable substances to reverse dysregulated epigenomic changes. One is the aberrant methylation of cytidine-phosphate-guanosine (CpG) dinucleotides, which are accumulated in “CpG islands” of genomic DNA in the promoter regions of genes (DNA methylation). The other is the deacetylation of histones, which results in a positive charge of these proteins, consequently a stronger binding of negatively charged DNA to them and ultimately producing a transcriptional repression by hindering access of transcription complexes to the DNA (histone deacetylation).

Already in the early 1980s, investigators demonstrated that cytidine analogs such as 5′-aza-cytidine or 5′-aza-2′-deoxycytidine (decitabine) were able to differentiate mouse embryo cells to muscle cells by inhibiting methylation of DNA.153  In the last few years, these drugs have established themselves as effective alternatives to cytotoxic chemotherapy in MDS and AML. Both compounds integrate into DNA as alternative nucleotides and trap DNA methyltransferases resulting in the formation of demethylated DNA.154  Due to this mechanism, hypermethylation of DNA in malignant cells is reversed in the course of several cell divisions.155  On the grounds of this gradual process, unlike chemotherapy, demethylating agents are not applied at a maximum tolerable dose but rather in smaller portions over a longer duration to induce differentiation and inhibit proliferation of the malignant cells.153,156  Recent clinical trials using these substances for the treatment of advanced MDS and AML have been promising157  with improvements of survival rates achieved with 5′-aza-cytidine over conventional therapeutic regimens in MDS.158 

Another class of epigenetically active substances are histone-deacetylase (HDAC) inhibitors. HDAC inhibitors manipulate cell growth and differentiation by inhibiting deacetylation of histones and proteins including transcription factors, and thereby reversing transcriptional repression of tumor suppressors or factors responsible for normal differentiation.159  Four classes of histone deacetylases with specific functions are known, and inhibitors against either single classes or a broader spectrum are available. HDAC inhibitors generally display low toxicity and some can be administered orally such as suberoylanilide hydroxamic acid (SAHA, vorinostat) or valproic acid (VPA), which has also widely been used as an anticonvulsive agent. SAHA was originally discovered while screening for differentiation-inducing compounds similar to the polar/planar compounds dimethylsulfoxide (DMSO) and hexamethylene bisacetamide (HMBA). SAHA has demonstrated antitumor effects in various cell lines and in vivo models of leukemia and solid tumors160,161 ; and phase 1 trials have shown efficacy of the drug in a panel of hematologic diseases.162  Furthermore, Vorinostat has successfully been used as a treatment for refractory cutaneous T-cell lymphoma,163  leading to its FDA approval. VPA has been shown to work synergistically with ATRA in cell lines to induce differentiation.164  A pilot study combining the 2 agents in 8 refractory or high-risk AML patients demonstrated clinical benefit in 7 of the patients, with evidence of hyperacetylation of histones and myelomonocytic differentiation of circulating blasts.165  A recently published phase 1 trial of an oral isotype–specific HDACi (MGCD0103, MethylGene, Quebec, QC; Celgene, Summit, NJ) in 29 patients with either AML or MDS, all previously treated with at least 1 prior chemotherapy regimen, produced complete hematologic remissions in 3 patients.166 

In summary, the manipulation of dysregulated transcription factors responsible for hematopoietic differentiation represents a powerful tool to be harnessed for the differentiation therapy of leukemia, pending development of targeted substances. Compounds that already partly achieved this by influencing the epigenetic landscape in favor of growth control and differentiation are already successfully being used clinically and will be developed further, to reduce their toxicity and improve their efficacy.

One of the most seminal achievements in cancer research of the last decade was the development of the tyrosine kinase inhibitor (TKI) imatinib for the treatment of CML with its characteristic Philadelphia chromosome as a drugable target.167  The unsurpassed success of imatinib in CML has led to great efforts to apply the approach of targeted inhibition of tyrosine kinases in other malignancies.

Targeting the dysregulated signaling of the epidermal growth factor receptor (EGFR) pathways has become a major strategy for the treatment of solid tumors. Two “small molecule” agents, which inhibit the intracellular tyrosine kinase activity of EGFR, have reached FDA approval. Gefinitib and erlotinib have shown efficacy for the treatment of a wide range of solid tumors.168  Recently, 2 groups demonstrated that these agents also have the ability to induce apoptosis and differentiation in AML cell lines and primary blasts even though these cells do not express EGFR.169-171  Microarray gene-expression analysis, immunophenotyping, and morphologic assessment have shown that these 2 tyrosine kinase inhibitors induce a differentiation program in myeloid leukemia cells that corresponds to neutrophil maturation. Moreover, a selective induction of apoptosis in CD34+ progenitors derived from either MDS or AML was observed, whereas CD34+ cells from healthy individuals were spared.171  Efforts to elucidate the observed off-target effects have shown that they are mediated at least in part by inhibition of the autophosphorylation of the oncogenic JAK2 kinase.171  A recent case report of a patient suffering from a non–small cell lung cancer (NSCLC) as well as myelogenous leukemia, who was treated with erlotinib monotherapy for 3 months and subsequently displayed a complete remission of his leukemia,172  corroborates the notion that the antileukemic efficacy of erlotinib or gefitinib may not be limited to laboratory experiments. Given the favorable profile of side effects of these substances and the need for milder antileukemic treatments for elderly patients, small molecule tyrosine kinase inhibitors and their off-target effects may represent a new treatment approach for leukemia, which merits further clinical testing.

In the past 3 decades, the arsenal and interrogational power of molecular methods to study malignancy has exploded. Aberrant epigenetic profiles in cancer are assessed with whole-genome CpG methylation and histone modification array techniques.152  Whole-genome analysis of DNA copy number changes and loss of heterozygosity is being performed with high-density single nucleotide polymorphism (SNP) arrays,142,143,173  comparative genomic hybridization, and whole-genome sequencing, which has uncovered mutations in cytogenetically normal AML.174  Candidate genes identified by these powerful screening methods can be analyzed for their function in murine knockout systems, by small interfering RNA inhibition, retroviral and lentiviral overexpression systems, and many other sophisticated new tools, which are constantly accelerating the pace of scientific progress in cancer research. Specifically concerning differentiation therapy, the most important recent development may be the revision of the prevailing conception that differentiation is a unidirectional process. In recent experiments, Takahashi and Yamanaka175  and Jaenisch and Young176  achieved a breakthrough by demonstrating that it is possible to dedifferentiate adult somatic cells to “inducible pluripotent stem cells” (IPS cells) by forced expression of 4 transcription factors: Oct4, Sox2, Klf4, and c-Myc. This nuclear reprogramming was first performed in murine fibroblasts; and after optimization of the method, IPS cells were generated, which were epigenetically and developmentally indistinguishable from embryo-derived stem cells.177-179  When these reprogrammed cells are transferred into a host blastocyst, they take part in normal differentiation to give rise to all 3 embryonic germ layers.176  By further manipulation of transcription factors such as transduction of C/EBPα or knockdown of Pax5, even mature B cells can be reprogrammed into IPS cells.180  This new molecular approach has immense therapeutic potential as it could ultimatively be used to create IPS cells from adult somatic tissue, manipulate, or “repair” these cells in an appropriate fashion and reintroduce them into the patient. Recent studies have shown a proof of principle of this notion by treating murine models of sickle cell anemia, hemophilia, as well as a rat model of Parkinson disease with transplantation of genetically engineered IPS cells.181-183  Recently, IPS cells have successfully been differentiated into hematopoietic progenitor cells184  and reprogrammed cells can be engrafted into irradiated severe combined immunodeficient (SCID) mice.185  Many questions remain such as the oncogenic potential of the factors introduced to reprogram the somatic cells into IPS cells or the fact that the epigenetic profile of the original aberrant somatic cells can be reprogrammed but the abnormal genomic template remains. However, this could even be used as a way to elucidate the relevance and contributions of epigenetic changes in cancer cells bearing specific genomic alterations. Intriguingly, the reprogramming effect of the transcription factors to produce IPS cells is necessary only for a limited time in the dedifferentiation process186  and methods are being refined to induce conditionally the necessary factors or even transiently, thereby ultimately eliminating many of the negative effects in the tissue arising from the IPS cells.

The paradigm that leukemias are characterized by the alteration of 2 sets of genes, those that give the malignancy a proliferative advantage and those associated with a block of differentiation, is still as valid today as it was 3 decades ago (Figure 3). To date, the hope that a variety of common chemicals would be identified that could induce differentiation of leukemia cells similar to the successes of ATRA in APL has not come to fruition on a broad scale. This can be attributed to the fact that candidate substances, which have been assessed for differentiation therapy subsequent to ATRA, have been unable to target a disease-specific lesion comparable with the PML-RARα fusion product in APL or the BCR-ABL chimeric protein in CML. Agents other than ATRA, which have been described in this review, are able to exert therapeutic differentiating effects on multiple deregulated pathways within the leukemic cells and therefore may be useful for improving combination therapies, especially because they often have few side effects. However, these differentiation agents are not targeted and potent enough to achieve the seminal successes of ATRA and imatinib. The future of differentiation therapy may lie in the manipulation of aberrant transcription factors in leukemia as these have emerged as powerful and common abnormalities in AML, ALL, and other cancers. As more knowledge is gathered about their mechanisms of action and their targets, new substances may be developed, which mimic the physiological action of transcription factors or compete for binding sites of mutated leukemogenic factors.

Figure 3

Timeline of milestones in differentiation therapy of leukemia. NHR indicates nuclear hormone receptor; TZD, thiazolidinediones; ATRA, all-trans retinoic acid; ATO, arsenic trioxide; and HDAC, histone deacetylase.

Figure 3

Timeline of milestones in differentiation therapy of leukemia. NHR indicates nuclear hormone receptor; TZD, thiazolidinediones; ATRA, all-trans retinoic acid; ATO, arsenic trioxide; and HDAC, histone deacetylase.

Close modal

This work was supported by the Parker Hughes Fund (Los Angeles, CA) and by grants from the National Institutes of Health (Bethesda, MD). D.N. was supported by a research grant from the Deutsche Forschungsgemeinschaft (DFG, Bonn, Germany; NO 817/1-1). H.P.K. holds the Mark Goodson Chair in Oncology Research at Cedars Sinai Medical Center (Los Angeles, CA) and is a member of the Jonsson Cancer Center and the Molecular Biology Institute of UCLA.

National Institutes of Health

Contribution: H.P.K. designed the review; and D.N., D.S., and H.P.K. wrote and proofread the paper.

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

Correspondence: Daniel Nowak, Division of Hematology and Oncology, Cedars Sinai Medical Center, UCLA School of Medicine, 8700 Beverly Blvd, Los Angeles, CA 90048; e-mail: daniel.nowak@cshs.org.

1
Friend
 
C
Scher
 
W
Holland
 
JG
Sato
 
T
Hemoglobin synthesis in murine virus-induced leukemic cells in vitro: stimulation of erythroid differentiation by dimethyl sulfoxide.
Proc Natl Acad Sci U S A
1971
, vol. 
68
 (pg. 
378
-
382
)
2
Sachs
 
L
Control of normal cell differentiation and the phenotypic reversion of malignancy in myeloid leukaemia.
Nature
1978
, vol. 
274
 (pg. 
535
-
539
)
3
Breitman
 
TR
Selonick
 
SE
Collins
 
SJ
Induction of differentiation of the human promyelocytic leukemia cell line (HL-60) by retinoic acid.
Proc Natl Acad Sci U S A
1980
, vol. 
77
 (pg. 
2936
-
2940
)
4
Breitman
 
TR
Collins
 
SJ
Keene
 
BR
Terminal differentiation of human promyelocytic leukemic cells in primary culture in response to retinoic acid.
Blood
1981
, vol. 
57
 (pg. 
1000
-
1004
)
5
Koeffler
 
HP
Induction of differentiation of human acute myelogenous leukemia cells: therapeutic implications.
Blood
1983
, vol. 
62
 (pg. 
709
-
721
)
6
Rowley
 
JD
Golomb
 
HM
Dougherty
 
C
15/17 translocation, a consistent chromosomal change in acute promyelocytic leukaemia.
Lancet
1977
, vol. 
1
 (pg. 
549
-
550
)
7
Nilsson
 
B
Probable in vivo induction of differentiation by retinoic acid of promyelocytes in acute promyelocytic leukaemia.
Br J Haematol
1984
, vol. 
57
 (pg. 
365
-
371
)
8
Daenen
 
S
Vellenga
 
E
van Dobbenburgh
 
OA
Halie
 
MR
Retinoic acid as antileukemic therapy in a patient with acute promyelocytic leukemia and Aspergillus pneumonia.
Blood
1986
, vol. 
67
 (pg. 
559
-
561
)
9
Huang
 
ME
Ye
 
YC
Chen
 
SR
, et al. 
All-trans retinoic acid with or without low dose cytosine arabinoside in acute promyelocytic leukemia. Report of 6 cases.
Chin Med J (Engl)
1987
, vol. 
100
 (pg. 
949
-
953
)
10
Huang
 
ME
Ye
 
YC
Chen
 
SR
, et al. 
Use of all-trans retinoic acid in the treatment of acute promyelocytic leukemia.
Blood
1988
, vol. 
72
 (pg. 
567
-
572
)
11
Tallman
 
MS
Andersen
 
JW
Schiffer
 
CA
, et al. 
All-trans retinoic acid in acute promyelocytic leukemia: long-term outcome and prognostic factor analysis from the North American Intergroup protocol.
Blood
2002
, vol. 
100
 (pg. 
4298
-
4302
)
12
Fenaux
 
P
Le Deley
 
MC
Castaigne
 
S
, et al. 
Effect of all transretinoic acid in newly diagnosed acute promyelocytic leukemia: results of a multicenter randomized trial: European APL 91 Group.
Blood
1993
, vol. 
82
 (pg. 
3241
-
3249
)
13
Mandelli
 
F
Diverio
 
D
Avvisati
 
G
, et al. 
Molecular remission in PML/RAR alpha-positive acute promyelocytic leukemia by combined all-trans retinoic acid and idarubicin (AIDA) therapy: Gruppo Italiano-Malattie Ematologiche Maligne dell'Adulto and Associazione Italiana di Ematologia ed Oncologia Pediatrica Cooperative Groups.
Blood
1997
, vol. 
90
 (pg. 
1014
-
1021
)
14
Avvisati
 
G
Lo Coco
 
F
Diverio
 
D
, et al. 
AIDA (all-trans retinoic acid + idarubicin) in newly diagnosed acute promyelocytic leukemia: a Gruppo Italiano Malattie Ematologiche Maligne dell'Adulto (GIMEMA) pilot study.
Blood
1996
, vol. 
88
 (pg. 
1390
-
1398
)
15
Lengfelder
 
E
Reichert
 
A
Schoch
 
C
, et al. 
Double induction strategy including high dose cytarabine in combination with all-trans retinoic acid: effects in patients with newly diagnosed acute promyelocytic leukemia: German AML Cooperative Group.
Leukemia
2000
, vol. 
14
 (pg. 
1362
-
1370
)
16
Fenaux
 
P
Chastang
 
C
Chevret
 
S
, et al. 
A randomized comparison of all transretinoic acid (ATRA) followed by chemotherapy and ATRA plus chemotherapy and the role of maintenance therapy in newly diagnosed acute promyelocytic leukemia: The European APL Group.
Blood
1999
, vol. 
94
 (pg. 
1192
-
1200
)
17
Niu
 
C
Yan
 
H
Yu
 
T
, et al. 
Studies on treatment of acute promyelocytic leukemia with arsenic trioxide: remission induction, follow-up, and molecular monitoring in 11 newly diagnosed and 47 relapsed acute promyelocytic leukemia patients.
Blood
1999
, vol. 
94
 (pg. 
3315
-
3324
)
18
Soignet
 
SL
Frankel
 
SR
Douer
 
D
, et al. 
United States multicenter study of arsenic trioxide in relapsed acute promyelocytic leukemia.
J Clin Oncol
2001
, vol. 
19
 (pg. 
3852
-
3860
)
19
Raffoux
 
E
Rousselot
 
P
Poupon
 
J
, et al. 
Combined treatment with arsenic trioxide and all-trans-retinoic acid in patients with relapsed acute promyelocytic leukemia.
J Clin Oncol
2003
, vol. 
21
 (pg. 
2326
-
2334
)
20
Ghavamzadeh
 
A
Alimoghaddam
 
K
Ghaffari
 
SH
, et al. 
Treatment of acute promyelocytic leukemia with arsenic trioxide without ATRA and/or chemotherapy.
Ann Oncol
2006
, vol. 
17
 (pg. 
131
-
134
)
21
Mathews
 
V
George
 
B
Lakshmi
 
KM
, et al. 
Single-agent arsenic trioxide in the treatment of newly diagnosed acute promyelocytic leukemia: durable remissions with minimal toxicity.
Blood
2006
, vol. 
107
 (pg. 
2627
-
2632
)
22
Estey
 
E
Garcia-Manero
 
G
Ferrajoli
 
A
, et al. 
Use of all-trans retinoic acid plus arsenic trioxide as an alternative to chemotherapy in untreated acute promyelocytic leukemia.
Blood
2006
, vol. 
107
 (pg. 
3469
-
3473
)
23
Sanz
 
MA
Grimwade
 
D
Tallman
 
MS
, et al. 
Guidelines on the management of acute promyelocytic leukemia: recommendations from an expert panel on behalf of the European LeukemiaNet.
Blood
 
Prepublished on September 23, 2008, as DOI 10.1182/blood-2008-04-150250
24
Collins
 
SJ
Retinoic acid receptors, hematopoiesis and leukemogenesis.
Curr Opin Hematol
2008
, vol. 
15
 (pg. 
346
-
351
)
25
de Thé
 
H
Chomienne
 
C
Lanotte
 
M
Degos
 
L
Dejean
 
A
The t(15;17) translocation of acute promyelocytic leukaemia fuses the retinoic acid receptor alpha gene to a novel transcribed locus.
Nature
1990
, vol. 
347
 (pg. 
558
-
561
)
26
Kakizuka
 
A
Miller
 
WH
Umesono
 
K
, et al. 
Chromosomal translocation t(15;17) in human acute promyelocytic leukemia fuses RAR alpha with a novel putative transcription factor, PML.
Cell
1991
, vol. 
66
 (pg. 
663
-
674
)
27
Wang
 
ZY
Chen
 
Z
Acute promyelocytic leukemia: from highly fatal to highly curable.
Blood
2008
, vol. 
111
 (pg. 
2505
-
2515
)
28
Carbone
 
R
Botrugno
 
OA
Ronzoni
 
S
, et al. 
Recruitment of the histone methyltransferase SUV39H1 and its role in the oncogenic properties of the leukemia-associated PML-retinoic acid receptor fusion protein.
Mol Cell Biol
2006
, vol. 
26
 (pg. 
1288
-
1296
)
29
Villa
 
R
Pasini
 
D
Gutierrez
 
A
, et al. 
Role of the polycomb repressive complex 2 in acute promyelocytic leukemia.
Cancer Cell
2007
, vol. 
11
 (pg. 
513
-
525
)
30
Di Croce
 
L
Raker
 
VA
Corsaro
 
M
, et al. 
Methyltransferase recruitment and DNA hypermethylation of target promoters by an oncogenic transcription factor.
Science
2002
, vol. 
295
 (pg. 
1079
-
1082
)
31
Scaglioni
 
PP
Pandolfi
 
PP
The theory of APL revisited.
Curr Top Microbiol Immunol
2007
, vol. 
313
 (pg. 
85
-
100
)
32
Grignani
 
F
De Matteis
 
S
Nervi
 
C
, et al. 
Fusion proteins of the retinoic acid receptor-alpha recruit histone deacetylase in promyelocytic leukaemia.
Nature
1998
, vol. 
391
 (pg. 
815
-
818
)
33
Zheng
 
PZ
Wang
 
KK
Zhang
 
QY
, et al. 
Systems analysis of transcriptome and proteome in retinoic acid/arsenic trioxide-induced cell differentiation/apoptosis of promyelocytic leukemia.
Proc Natl Acad Sci U S A
2005
, vol. 
102
 (pg. 
7653
-
7658
)
34
Park
 
DJ
Vuong
 
PT
de Vos
 
S
Douer
 
D
Koeffler
 
HP
Comparative analysis of genes regulated by PML/RAR alpha and PLZF/RAR alpha in response to retinoic acid using oligonucleotide arrays.
Blood
2003
, vol. 
102
 (pg. 
3727
-
3736
)
35
Bentley
 
DL
Groudine
 
M
A block to elongation is largely responsible for decreased transcription of c-myc in differentiated HL60 cells.
Nature
1986
, vol. 
321
 (pg. 
702
-
706
)
36
Park
 
DJ
Chumakov
 
AM
Vuong
 
PT
, et al. 
CCAAT/enhancer binding protein epsilon is a potential retinoid target gene in acute promyelocytic leukemia treatment.
J Clin Invest
1999
, vol. 
103
 (pg. 
1399
-
1408
)
37
Mueller
 
BU
Pabst
 
T
Fos
 
J
, et al. 
ATRA resolves the differentiation block in t(15;17) acute myeloid leukemia by restoring PU.1 expression.
Blood
2006
, vol. 
107
 (pg. 
3330
-
3338
)
38
Ozpolat
 
B
Akar
 
U
Steiner
 
M
, et al. 
Programmed cell death-4 tumor suppressor protein contributes to retinoic acid-induced terminal granulocytic differentiation of human myeloid leukemia cells.
Mol Cancer Res
2007
, vol. 
5
 (pg. 
95
-
108
)
39
Gery
 
S
Park
 
DJ
Vuong
 
PT
, et al. 
RTP801 is a novel retinoic acid-responsive gene associated with myeloid differentiation.
Exp Hematol
2007
, vol. 
35
 (pg. 
572
-
578
)
40
Kitareewan
 
S
Pitha-Rowe
 
I
Sekula
 
D
, et al. 
UBE1L is a retinoid target that triggers PML/RARalpha degradation and apoptosis in acute promyelocytic leukemia.
Proc Natl Acad Sci U S A
2002
, vol. 
99
 (pg. 
3806
-
3811
)
41
Hattori
 
H
Zhang
 
X
Jia
 
Y
, et al. 
RNAi screen identifies UBE2D3 as a mediator of all-trans retinoic acid-induced cell growth arrest in human acute promyelocytic NB4 cells.
Blood
2007
, vol. 
110
 (pg. 
640
-
650
)
42
Chen
 
GQ
Shi
 
XG
Tang
 
W
, et al. 
Use of arsenic trioxide (As2O3) in the treatment of acute promyelocytic leukemia (APL): I, As2O3 exerts dose-dependent dual effects on APL cells.
Blood
1997
, vol. 
89
 (pg. 
3345
-
3353
)
43
Zhang
 
TD
Chen
 
GQ
Wang
 
ZG
Wang
 
ZY
Chen
 
SJ
Chen
 
Z
Arsenic trioxide, a therapeutic agent for APL.
Oncogene
2001
, vol. 
20
 (pg. 
7146
-
7153
)
44
Matsui
 
W
Smith
 
BD
Vala
 
M
, et al. 
Requirement for myeloid growth factors in the differentiation of acute promyelocytic leukaemia.
Br J Haematol
2005
, vol. 
128
 (pg. 
853
-
862
)
45
Miyaura
 
C
Abe
 
E
Kuribayashi
 
T
, et al. 
1 alpha,25-Dihydroxyvitamin D3 induces differentiation of human myeloid leukemia cells.
Biochem Biophys Res Commun
1981
, vol. 
102
 (pg. 
937
-
943
)
46
Koeffler
 
HP
Hirji
 
K
Itri
 
L
1,25-Dihydroxyvitamin D3: in vivo and in vitro effects on human preleukemic and leukemic cells.
Cancer Treat Rep
1985
, vol. 
69
 (pg. 
1399
-
1407
)
47
Kelsey
 
SM
Newland
 
AC
Cunningham
 
J
, et al. 
Sustained haematological response to high-dose oral alfacalcidol in patients with myelodysplastic syndromes.
Lancet
1992
, vol. 
340
 (pg. 
316
-
317
)
48
Hellström
 
E
Robert
 
KH
Samuelsson
 
J
, et al. 
Treatment of myelodysplastic syndromes with retinoic acid and 1 alpha-hydroxy-vitamin D3 in combination with low-dose ara-C is not superior to ara-C alone: results from a randomized study: The Scandinavian Myelodysplasia Group (SMG).
Eur J Haematol
1990
, vol. 
45
 (pg. 
255
-
261
)
49
Slapak
 
CA
Desforges
 
JF
Fogaren
 
T
Miller
 
KB
Treatment of acute myeloid leukemia in the elderly with low-dose cytarabine, hydroxyurea, and calcitriol.
Am J Hematol
1992
, vol. 
41
 (pg. 
178
-
183
)
50
Ferrero
 
D
Bruno
 
B
Pregno
 
P
, et al. 
Combined differentiating therapy for myelodysplastic syndromes: a phase II study.
Leuk Res
1996
, vol. 
20
 (pg. 
867
-
876
)
51
Deeb
 
KK
Trump
 
DL
Johnson
 
CS
Vitamin D signalling pathways in cancer: potential for anticancer therapeutics.
Nat Rev Cancer
2007
, vol. 
7
 (pg. 
684
-
700
)
52
Banach-Petrosky
 
W
Ouyang
 
X
Gao
 
H
, et al. 
Vitamin D inhibits the formation of prostatic intraepithelial neoplasia in Nkx3. 1;Pten mutant mice.
Clin Cancer Res
2006
, vol. 
12
 (pg. 
5895
-
5901
)
53
Huerta
 
S
Irwin
 
RW
Heber
 
D
, et al. 
1alpha,25-(OH)(2)-D(3) and its synthetic analogue decrease tumor load in the Apc(min) Mouse.
Cancer Res
2002
, vol. 
62
 (pg. 
741
-
746
)
54
O'Kelly
 
J
Koeffler
 
HP
Vitamin D analogs and breast cancer: recent results.
Cancer Res
2003
, vol. 
164
 (pg. 
333
-
348
)
55
Shabahang
 
M
Buras
 
RR
Davoodi
 
F
Schumaker
 
LM
Nauta
 
RJ
Evans
 
SR
1,25-Dihydroxyvitamin D3 receptor as a marker of human colon carcinoma cell line differentiation and growth inhibition.
Cancer Res
1993
, vol. 
53
 (pg. 
3712
-
3718
)
56
Simboli-Campbell
 
M
Narvaez
 
CJ
Tenniswood
 
M
Welsh
 
J
1,25-Dihydroxyvitamin D3 induces morphological and biochemical markers of apoptosis in MCF-7 breast cancer cells.
J Steroid Biochem Mol Biol
1996
, vol. 
58
 (pg. 
367
-
376
)
57
Mantell
 
DJ
Owens
 
PE
Bundred
 
NJ
Mawer
 
EB
Canfield
 
AE
1 alpha,25-dihydroxyvitamin D(3) inhibits angiogenesis in vitro and in vivo.
Circ Res
2000
, vol. 
87
 (pg. 
214
-
220
)
58
Wang
 
X
Studzinski
 
GP
Activation of extracellular signal-regulated kinases (ERKs) defines the first phase of 1,25-dihydroxyvitamin D3-induced differentiation of HL60 cells.
J Cell Biochem
2001
, vol. 
80
 (pg. 
471
-
482
)
59
Evans
 
RM
The steroid and thyroid hormone receptor superfamily.
Science
1988
, vol. 
240
 (pg. 
889
-
895
)
60
Carlberg
 
C
Bendik
 
I
Wyss
 
A
, et al. 
Two nuclear signalling pathways for vitamin D.
Nature
1993
, vol. 
361
 (pg. 
657
-
660
)
61
O'Kelly
 
J
Hisatake
 
J
Hisatake
 
Y
Bishop
 
J
Norman
 
A
Koeffler
 
HP
Normal myelopoiesis but abnormal T lymphocyte responses in vitamin D receptor knockout mice.
J Clin Invest
2002
, vol. 
109
 (pg. 
1091
-
1099
)
62
Liu
 
M
Lee
 
MH
Cohen
 
M
Bommakanti
 
M
Freedman
 
LP
Transcriptional activation of the Cdk inhibitor p21 by vitamin D3 leads to the induced differentiation of the myelomonocytic cell line U937.
Genes Dev
1996
, vol. 
10
 (pg. 
142
-
153
)
63
Verlinden
 
L
Verstuyf
 
A
Convents
 
R
Marcelis
 
S
Van Camp
 
M
Bouillon
 
R
Action of 1,25(OH)2D3 on the cell cycle genes, cyclin D1, p21 and p27 in MCF-7 cells.
Mol Cell Endocrinol
1998
, vol. 
142
 (pg. 
57
-
65
)
64
Jensen
 
SS
Madsen
 
MW
Lukas
 
J
Binderup
 
L
Bartek
 
J
Inhibitory effects of 1alpha,25-dihydroxyvitamin D(3) on the G(1)-S phase-controlling machinery.
Mol Endocrinol
2001
, vol. 
15
 (pg. 
1370
-
1380
)
65
Pakkala
 
S
de Vos
 
S
Elstner
 
E
, et al. 
Vitamin D3 analogs: effect on leukemic clonal growth and differentiation, and on serum calcium levels.
Leuk Res
1995
, vol. 
19
 (pg. 
65
-
72
)
66
Trump
 
DL
Potter
 
DM
Muindi
 
J
Brufsky
 
A
Johnson
 
CS
Phase II trial of high-dose, intermittent calcitriol (1,25 dihydroxyvitamin D3) and dexamethasone in androgen-independent prostate cancer.
Cancer
2006
, vol. 
106
 (pg. 
2136
-
2142
)
67
Okamoto
 
R
Akagi
 
T
Koeffler
 
P
Vitamin D compounds and myelodysplastic syndrome.
Leuk Lymphoma
2008
, vol. 
49
 (pg. 
12
-
13
)
68
Koeffler
 
HP
Aslanian
 
N
O'Kelly
 
J
Vitamin D(2) analog (Paricalcitol; Zemplar) for treatment of myelodysplastic syndrome.
Leuk Res
2005
, vol. 
29
 (pg. 
1259
-
1262
)
69
Wang
 
TT
Nestel
 
FP
Bourdeau
 
V
, et al. 
Cutting edge: 1,25-dihydroxyvitamin D3 is a direct inducer of antimicrobial peptide gene expression.
J Immunol
2004
, vol. 
173
 (pg. 
2909
-
2912
)
70
Gombart
 
AF
Borregaard
 
N
Koeffler
 
HP
Human cathelicidin antimicrobial peptide (CAMP) gene is a direct target of the vitamin D receptor and is strongly up-regulated in myeloid cells by 1,25-dihydroxyvitamin D3.
Faseb J
2005
, vol. 
19
 (pg. 
1067
-
1077
)
71
Gombart
 
AF
O'Kelly
 
J
Saito
 
T
Koeffler
 
HP
Regulation of the CAMP gene by 1,25(OH)2D3 in various tissues.
J Steroid Biochem Mol Biol
2007
, vol. 
103
 (pg. 
552
-
557
)
72
Forman
 
BM
Chen
 
J
Evans
 
RM
Hypolipidemic drugs, polyunsaturated fatty acids, and eicosanoids are ligands for peroxisome proliferator-activated receptors alpha and delta.
Proc Natl Acad Sci U S A
1997
, vol. 
94
 (pg. 
4312
-
4317
)
73
Lehmann
 
JM
Moore
 
LB
Smith-Oliver
 
TA
Wilkison
 
WO
Willson
 
TM
Kliewer
 
SA
An antidiabetic thiazolidinedione is a high affinity ligand for peroxisome proliferator-activated receptor gamma (PPAR gamma).
J Biol Chem
1995
, vol. 
270
 (pg. 
12953
-
12956
)
74
Saltiel
 
AR
Olefsky
 
JM
Thiazolidinediones in the treatment of insulin resistance and type II diabetes.
Diabetes
1996
, vol. 
45
 (pg. 
1661
-
1669
)
75
Koeffler
 
HP
Peroxisome proliferator-activated receptor gamma and cancers.
Clin Cancer Res
2003
, vol. 
9
 (pg. 
1
-
9
)
76
Tontonoz
 
P
Hu
 
E
Spiegelman
 
BM
Stimulation of adipogenesis in fibroblasts by PPAR gamma 2, a lipid-activated transcription factor.
Cell
1994
, vol. 
79
 (pg. 
1147
-
1156
)
77
Tontonoz
 
P
Singer
 
S
Forman
 
BM
, et al. 
Terminal differentiation of human liposarcoma cells induced by ligands for peroxisome proliferator-activated receptor gamma and the retinoid X receptor.
Proc Natl Acad Sci U S A
1997
, vol. 
94
 (pg. 
237
-
241
)
78
Demetri
 
GD
Fletcher
 
CD
Mueller
 
E
, et al. 
Induction of solid tumor differentiation by the peroxisome proliferator-activated receptor-gamma ligand troglitazone in patients with liposarcoma.
Proc Natl Acad Sci U S A
1999
, vol. 
96
 (pg. 
3951
-
3956
)
79
Hirase
 
N
Yanase
 
T
Mu
 
Y
, et al. 
Thiazolidinedione induces apoptosis and monocytic differentiation in the promyelocytic leukemia cell line HL60.
Oncology
1999
, vol. 
57
 (pg. 
17
-
26
)
80
Fujimura
 
S
Suzumiya
 
J
Nakamura
 
K
Ono
 
J
Effects of troglitazone on the growth and differentiation of hematopoietic cell lines.
Int J Oncol
1998
, vol. 
13
 (pg. 
1263
-
1267
)
81
Sugimura
 
A
Kiriyama
 
Y
Nochi
 
H
, et al. 
Troglitazone suppresses cell growth of myeloid leukemia cell lines by induction of p21WAF1/CIP1 cyclin-dependent kinase inhibitor.
Biochem Biophys Res Commun
1999
, vol. 
261
 (pg. 
833
-
837
)
82
Konopleva
 
M
Elstner
 
E
McQueen
 
TJ
, et al. 
Peroxisome proliferator-activated receptor gamma and retinoid X receptor ligands are potent inducers of differentiation and apoptosis in leukemias.
Mol Cancer Ther
2004
, vol. 
3
 (pg. 
1249
-
1262
)
83
Zang
 
C
Liu
 
H
Waechter
 
M
, et al. 
Dual PPARalpha/gamma ligand TZD18 either alone or in combination with imatinib inhibits proliferation and induces apoptosis of human CML cell lines.
Cell Cycle
2006
, vol. 
5
 (pg. 
2237
-
2243
)
84
Mangelsdorf
 
DJ
Evans
 
RM
The RXR heterodimers and orphan receptors.
Cell
1995
, vol. 
83
 (pg. 
841
-
850
)
85
Wang
 
T
Xu
 
J
Yu
 
X
Yang
 
R
Han
 
ZC
Peroxisome proliferator-activated receptor gamma in malignant diseases.
Crit Rev Oncol Hematol
2006
, vol. 
58
 (pg. 
1
-
14
)
86
Elstner
 
E
Muller
 
C
Koshizuka
 
K
, et al. 
Ligands for peroxisome proliferator-activated receptorgamma and retinoic acid receptor inhibit growth and induce apoptosis of human breast cancer cells in vitro and in BNX mice.
Proc Natl Acad Sci U S A
1998
, vol. 
95
 (pg. 
8806
-
8811
)
87
Zander
 
T
Kraus
 
JA
Grommes
 
C
, et al. 
Induction of apoptosis in human and rat glioma by agonists of the nuclear receptor PPARgamma.
J Neurochem
2002
, vol. 
81
 (pg. 
1052
-
1060
)
88
Patel
 
L
Pass
 
I
Coxon
 
P
Downes
 
CP
Smith
 
SA
Macphee
 
CH
Tumor suppressor and anti-inflammatory actions of PPARgamma agonists are mediated via upregulation of PTEN.
Curr Biol
2001
, vol. 
11
 (pg. 
764
-
768
)
89
Pignatelli
 
M
Cocca
 
C
Santos
 
A
Perez-Castillo
 
A
Enhancement of BRCA1 gene expression by the peroxisome proliferator-activated receptor gamma in the MCF-7 breast cancer cell line.
Oncogene
2003
, vol. 
22
 (pg. 
5446
-
5450
)
90
Chawla
 
A
Barak
 
Y
Nagy
 
L
Liao
 
D
Tontonoz
 
P
Evans
 
RM
PPAR-gamma dependent and independent effects on macrophage-gene expression in lipid metabolism and inflammation.
Nat Med
2001
, vol. 
7
 (pg. 
48
-
52
)
91
Palakurthi
 
SS
Aktas
 
H
Grubissich
 
LM
Mortensen
 
RM
Halperin
 
JA
Anticancer effects of thiazolidinediones are independent of peroxisome proliferator-activated receptor gamma and mediated by inhibition of translation initiation.
Cancer Res
2001
, vol. 
61
 (pg. 
6213
-
6218
)
92
Shiau
 
CW
Yang
 
CC
Kulp
 
SK
Chen
 
KF
Chen
 
CS
Huang
 
JW
Thiazolidenediones mediate apoptosis in prostate cancer cells in part through inhibition of Bcl-xL/Bcl-2 functions independently of PPARgamma.
Cancer Res
2005
, vol. 
65
 (pg. 
1561
-
1569
)
93
Wang
 
YL
Miao
 
Q
To live or to die: prosurvival activity of PPARgamma in cancers.
PPAR Res
2008
, vol. 
2008
 pg. 
209629
 
94
Lefebvre
 
AM
Chen
 
I
Desreumaux
 
P
, et al. 
Activation of the peroxisome proliferator-activated receptor gamma promotes the development of colon tumors in C57BL/6J-APCMin/+ mice.
Nat Med
1998
, vol. 
4
 (pg. 
1053
-
1057
)
95
Yang
 
K
Fan
 
KH
Lamprecht
 
SA
, et al. 
Peroxisome proliferator-activated receptor gamma agonist troglitazone induces colon tumors in normal C57BL/6J mice and enhances colonic carcinogenesis in Apc1638 N/+ Mlh1+/- double mutant mice.
Int J Cancer
2005
, vol. 
116
 (pg. 
495
-
499
)
96
Saez
 
E
Rosenfeld
 
J
Livolsi
 
A
, et al. 
PPAR gamma signaling exacerbates mammary gland tumor development.
Genes Dev
2004
, vol. 
18
 (pg. 
528
-
540
)
97
Debrock
 
G
Vanhentenrijk
 
V
Sciot
 
R
Debiec-Rychter
 
M
Oyen
 
R
Van Oosterom
 
A
A phase II trial with rosiglitazone in liposarcoma patients.
Br J Cancer
2003
, vol. 
89
 (pg. 
1409
-
1412
)
98
Burstein
 
HJ
Demetri
 
GD
Mueller
 
E
Sarraf
 
P
Spiegelman
 
BM
Winer
 
EP
Use of the peroxisome proliferator-activated receptor (PPAR) gamma ligand troglitazone as treatment for refractory breast cancer: a phase II study.
Breast Cancer Res Treat
2003
, vol. 
79
 (pg. 
391
-
397
)
99
Yee
 
LD
Williams
 
N
Wen
 
P
, et al. 
Pilot study of rosiglitazone therapy in women with breast cancer: effects of short-term therapy on tumor tissue and serum markers.
Clin Cancer Res
2007
, vol. 
13
 (pg. 
246
-
252
)
100
Yamazaki
 
K
Shimizu
 
M
Okuno
 
M
, et al. 
Synergistic effects of RXR alpha and PPAR gamma ligands to inhibit growth in human colon cancer cells: phosphorylated RXR alpha is a critical target for colon cancer management.
Gut
2007
, vol. 
56
 (pg. 
1557
-
1563
)
101
Cesario
 
RM
Stone
 
J
Yen
 
WC
Bissonnette
 
RP
Lamph
 
WW
Differentiation and growth inhibition mediated via the RXR:PPARgamma heterodimer in colon cancer.
Cancer Lett
2006
, vol. 
240
 (pg. 
225
-
233
)
102
Chang
 
TH
Szabo
 
E
Enhanced growth inhibition by combination differentiation therapy with ligands of peroxisome proliferator-activated receptor-gamma and inhibitors of histone deacetylase in adenocarcinoma of the lung.
Clin Cancer Res
2002
, vol. 
8
 (pg. 
1206
-
1212
)
103
Lipkowitz
 
S
Dennis
 
PA
PPARgamma agonists follow an unknown TRAIL in lung cancer.
Cancer Biol Ther
2007
, vol. 
6
 (pg. 
107
-
109
)
104
Zou
 
W
Liu
 
X
Yue
 
P
Khuri
 
FR
Sun
 
SY
PPARgamma ligands enhance TRAIL-induced apoptosis through DR5 upregulation and c-FLIP downregulation in human lung cancer cells.
Cancer Biol Ther
2007
, vol. 
6
 (pg. 
99
-
106
)
105
Hisatake
 
JI
Ikezoe
 
T
Carey
 
M
Holden
 
S
Tomoyasu
 
S
Koeffler
 
HP
Down-Regulation of prostate-specific antigen expression by ligands for peroxisome proliferator-activated receptor gamma in human prostate cancer.
Cancer Res
2000
, vol. 
60
 (pg. 
5494
-
5498
)
106
Veliceasa
 
D
Schulze-Hoepfner
 
FT
Volpert
 
OV
PPARgamma and agonists against cancer: rational design of complementation treatments.
PPAR Res
2008
, vol. 
2008
 pg. 
945275
 
107
Metcalf
 
D
Hematopoietic cytokines.
Blood
2008
, vol. 
111
 (pg. 
485
-
491
)
108
Bradley
 
TR
Metcalf
 
D
The growth of mouse bone marrow cells in vitro.
Aust J Exp Biol Med Sci
1966
, vol. 
44
 (pg. 
287
-
299
)
109
Pluznik
 
DH
Sachs
 
L
The induction of clones of normal mast cells by a substance from conditioned medium.
Exp Cell Res
1966
, vol. 
43
 (pg. 
553
-
563
)
110
Metcalf
 
D
Haemopoietic growth factors.
Med J Aust
1988
, vol. 
148
 (pg. 
516
-
519
)
111
Leung
 
KN
Mak
 
NK
Fung
 
MC
Cytokines in the differentiation therapy of leukemia: from laboratory investigations to clinical applications.
Crit Rev Clin Lab Sci
2005
, vol. 
42
 (pg. 
473
-
514
)
112
Koss
 
A
Lucero
 
G
Koziner
 
B
Granulocyte-colony stimulating factor, granulocyte-macrophage colony stimulating factor and interleukin 4 induce differentiation in the U-937 human monocytic leukemia cell line.
Leuk Lymphoma
1996
, vol. 
22
 (pg. 
163
-
171
follow. 186, color plate XIV-V
113
Goliaei
 
B
Deizadji
 
A
Effects of hyperthermia and granulocyte-macrophage colony-stimulating factor on the differentiation of human leukemic cell line U937.
Leuk Res
1998
, vol. 
22
 (pg. 
705
-
710
)
114
Kamano
 
H
Tanaka
 
T
Ohnishi
 
H
, et al. 
Effects of the antisense myb expression on hemin- and erythropoietin-induced erythroid differentiation of K562 cells.
Biochem Mol Biol Int
1994
, vol. 
34
 (pg. 
85
-
92
)
115
Kamijo
 
R
Takeda
 
K
Nagumo
 
M
Konno
 
K
Effects of combinations of transforming growth factor-beta 1 and tumor necrosis factor on induction of differentiation of human myelogenous leukemic cell lines.
J Immunol
1990
, vol. 
144
 (pg. 
1311
-
1316
)
116
Liebermann
 
DA
Hoffman-Liebermann
 
B
Proto-oncogene expression and dissection of the myeloid growth to differentiation developmental cascade.
Oncogene
1989
, vol. 
4
 (pg. 
583
-
592
)
117
Larsson
 
LG
Pettersson
 
M
Oberg
 
F
Nilsson
 
K
Luscher
 
B
Expression of mad, mxi1, max and c-myc during induced differentiation of hematopoietic cells: opposite regulation of mad and c-myc.
Oncogene
1994
, vol. 
9
 (pg. 
1247
-
1252
)
118
Zervos
 
AS
Gyuris
 
J
Brent
 
R
Mxi1, a protein that specifically interacts with Max to bind Myc-Max recognition sites.
Cell
1993
, vol. 
72
 (pg. 
223
-
232
)
119
Sakurai
 
T
Yamada
 
T
Kihara-Negishi
 
F
, et al. 
Effects of overexpression of the Ets family transcription factor TEL on cell growth and differentiation of K562 cells.
Int J Oncol
2003
, vol. 
22
 (pg. 
1327
-
1333
)
120
Hodge
 
DR
Li
 
D
Qi
 
SM
Farrar
 
WL
IL-6 induces expression of the Fli-1 proto-oncogene via STAT3.
Biochem Biophys Res Commun
2002
, vol. 
292
 (pg. 
287
-
291
)
121
Takamatsu
 
Y
Miyamoto
 
T
Iwasaki
 
H
Makino
 
S
Tamura
 
K
Remission induction by granulocyte colony-stimulating factor in hypoplastic acute myelogenous leukemia complicated by infection: a case report and review of the literature.
Acta Haematol
1998
, vol. 
99
 (pg. 
224
-
230
)
122
Bassan
 
R
Rambaldi
 
A
Amaru
 
R
Motta
 
T
Barbui
 
T
Unexpected remission of acute myeloid leukaemia after GM-CSF.
Br J Haematol
1994
, vol. 
87
 (pg. 
835
-
838
)
123
Fujiwara
 
H
Arima
 
N
Matsushita
 
K
, et al. 
Granulocyte-colony stimulating factor induces differentiation and apoptosis of CD2, CD7 positive hybrid leukemia cells in vivo and ex vivo.
Leuk Res
1997
, vol. 
21
 (pg. 
735
-
741
)
124
Piccaluga
 
PP
Martinelli
 
G
Malagola
 
M
, et al. 
Complete remission in acute myeloid leukemia with granulocyte-colony stimulating factor without chemotherapy: report of cytogenetic remission of a t(9;11)(p22q23) positive AML patient and review of literature.
Haematologica
2003
, vol. 
88
 pg. 
ECR28
 
125
Giralt
 
S
Escudier
 
S
Kantarjian
 
H
, et al. 
Preliminary results of treatment with filgrastim for relapse of leukemia and myelodysplasia after allogeneic bone marrow transplantation.
N Engl J Med
1993
, vol. 
329
 (pg. 
757
-
761
)
126
Ravandi
 
F
Role of cytokines in the treatment of acute leukemias: a review.
Leukemia
2006
, vol. 
20
 (pg. 
563
-
571
)
127
Berliner
 
N
Lessons from congenital neutropenia: 50 years of progress in understanding myelopoiesis.
Blood
2008
, vol. 
111
 (pg. 
5427
-
5432
)
128
Tenen
 
DG
Disruption of differentiation in human cancer: AML shows the way.
Nat Rev Cancer
2003
, vol. 
3
 (pg. 
89
-
101
)
129
Tenen
 
DG
Hromas
 
R
Licht
 
JD
Zhang
 
DE
Transcription factors, normal myeloid development, and leukemia.
Blood
1997
, vol. 
90
 (pg. 
489
-
519
)
130
Friedman
 
AD
Transcriptional regulation of granulocyte and monocyte development.
Oncogene
2002
, vol. 
21
 (pg. 
3377
-
3390
)
131
Zhang
 
DE
Zhang
 
P
Wang
 
ND
Hetherington
 
CJ
Darlington
 
GJ
Tenen
 
DG
Absence of granulocyte colony-stimulating factor signaling and neutrophil development in CCAAT enhancer binding protein alpha-deficient mice.
Proc Natl Acad Sci U S A
1997
, vol. 
94
 (pg. 
569
-
574
)
132
Trivedi
 
AK
Pal
 
P
Behre
 
G
Singh
 
SM
Multiple ways of C/EBPalpha inhibition in myeloid leukaemia.
Eur J Cancer
2008
, vol. 
44
 (pg. 
1516
-
1523
)
133
Pabst
 
T
Mueller
 
BU
Zhang
 
P
, et al. 
Dominant-negative mutations of CEBPA, encoding CCAAT/enhancer binding protein-alpha (C/EBPalpha), in acute myeloid leukemia.
Nat Genet
2001
, vol. 
27
 (pg. 
263
-
270
)
134
Gombart
 
AF
Hofmann
 
WK
Kawano
 
S
, et al. 
Mutations in the gene encoding the transcription factor CCAAT/enhancer binding protein alpha in myelodysplastic syndromes and acute myeloid leukemias.
Blood
2002
, vol. 
99
 (pg. 
1332
-
1340
)
135
Pabst
 
T
Mueller
 
BU
Transcriptional dysregulation during myeloid transformation in AML.
Oncogene
2007
, vol. 
26
 (pg. 
6829
-
6837
)
136
Pabst
 
T
Mueller
 
BU
Harakawa
 
N
, et al. 
AML1-ETO downregulates the granulocytic differentiation factor C/EBPalpha in t(8;21) myeloid leukemia.
Nat Med
2001
, vol. 
7
 (pg. 
444
-
451
)
137
Chim
 
CS
Wong
 
AS
Kwong
 
YL
Infrequent hypermethylation of CEBPA promotor in acute myeloid leukaemia.
Br J Haematol
2002
, vol. 
119
 (pg. 
988
-
990
)
138
Ross
 
SE
Radomska
 
HS
Wu
 
B
, et al. 
Phosphorylation of C/EBPalpha inhibits granulopoiesis.
Mol Cell Biol
2004
, vol. 
24
 (pg. 
675
-
686
)
139
Keeshan
 
K
He
 
Y
Wouters
 
BJ
, et al. 
Tribbles homolog 2 inactivates C/EBPalpha and causes acute myelogenous leukemia.
Cancer Cell
2006
, vol. 
10
 (pg. 
401
-
411
)
140
Radomska
 
HS
Huettner
 
CS
Zhang
 
P
Cheng
 
T
Scadden
 
DT
Tenen
 
DG
CCAAT/enhancer binding protein alpha is a regulatory switch sufficient for induction of granulocytic development from bipotential myeloid progenitors.
Mol Cell Biol
1998
, vol. 
18
 (pg. 
4301
-
4314
)
141
Tavor
 
S
Park
 
DJ
Gery
 
S
Vuong
 
PT
Gombart
 
AF
Koeffler
 
HP
Restoration of C/EBPalpha expression in a BCR-ABL+ cell line induces terminal granulocytic differentiation.
J Biol Chem
2003
, vol. 
278
 (pg. 
52651
-
52659
)
142
Mullighan
 
CG
Goorha
 
S
Radtke
 
I
, et al. 
Genome-wide analysis of genetic alterations in acute lymphoblastic leukaemia.
Nature
2007
, vol. 
446
 (pg. 
758
-
764
)
143
Kawamata
 
N
Ogawa
 
S
Zimmermann
 
M
, et al. 
Cloning of genes involved in chromosomal translocations by high-resolution single nucleotide polymorphism genomic microarray.
Proc Natl Acad Sci U S A
2008
, vol. 
105
 (pg. 
11921
-
11926
)
144
Holmes
 
ML
Pridans
 
C
Nutt
 
SL
The regulation of the B-cell gene expression programme by Pax5.
Immunol Cell Biol
2008
, vol. 
86
 (pg. 
47
-
53
)
145
Urbanek
 
P
Wang
 
ZQ
Fetka
 
I
Wagner
 
EF
Busslinger
 
M
Complete block of early B cell differentiation and altered patterning of the posterior midbrain in mice lacking Pax5/BSAP.
Cell
1994
, vol. 
79
 (pg. 
901
-
912
)
146
Nutt
 
SL
Urbanek
 
P
Rolink
 
A
Busslinger
 
M
Essential functions of Pax5 (BSAP) in pro-B cell development: difference between fetal and adult B lymphopoiesis and reduced V-to-DJ recombination at the IgH locus.
Genes Dev
1997
, vol. 
11
 (pg. 
476
-
491
)
147
Cobaleda
 
C
Busslinger
 
M
Developmental plasticity of lymphocytes.
Curr Opin Immunol
2008
, vol. 
20
 (pg. 
139
-
148
)
148
Cobaleda
 
C
Jochum
 
W
Busslinger
 
M
Conversion of mature B cells into T cells by dedifferentiation to uncommitted progenitors.
Nature
2007
, vol. 
449
 (pg. 
473
-
477
)
149
Sekine
 
R
Kitamura
 
T
Tsuji
 
T
Tojo
 
A
Efficient retroviral transduction of human B-lymphoid and myeloid progenitors: marked inhibition of their growth by the Pax5 transgene.
Int J Hematol
2008
, vol. 
87
 (pg. 
351
-
362
)
150
Souabni
 
A
Jochum
 
W
Busslinger
 
M
Oncogenic role of Pax5 in the T-lymphoid lineage upon ectopic expression from the immunoglobulin heavy-chain locus.
Blood
2007
, vol. 
109
 (pg. 
281
-
289
)
151
Anderson
 
K
Rusterholz
 
C
Mansson
 
R
, et al. 
Ectopic expression of PAX5 promotes maintenance of biphenotypic myeloid progenitors coexpressing myeloid and B-cell lineage-associated genes.
Blood
2007
, vol. 
109
 (pg. 
3697
-
3705
)
152
Figueroa
 
ME
Reimers
 
M
Thompson
 
RF
, et al. 
An integrative genomic and epigenomic approach for the study of transcriptional regulation.
PLoS ONE
2008
, vol. 
3
 pg. 
e1882
 
153
Jones
 
PA
Taylor
 
SM
Cellular differentiation, cytidine analogs and DNA methylation.
Cell
1980
, vol. 
20
 (pg. 
85
-
93
)
154
Egger
 
G
Liang
 
G
Aparicio
 
A
Jones
 
PA
Epigenetics in human disease and prospects for epigenetic therapy.
Nature
2004
, vol. 
429
 (pg. 
457
-
463
)
155
Daskalakis
 
M
Nguyen
 
TT
Nguyen
 
C
, et al. 
Demethylation of a hypermethylated P15/INK4B gene in patients with myelodysplastic syndrome by 5-Aza-2′-deoxycytidine (decitabine) treatment.
Blood
2002
, vol. 
100
 (pg. 
2957
-
2964
)
156
Issa
 
JP
Garcia-Manero
 
G
Giles
 
FJ
, et al. 
Phase 1 study of low-dose prolonged exposure schedules of the hypomethylating agent 5-aza-2′-deoxycytidine (decitabine) in hematopoietic malignancies.
Blood
2004
, vol. 
103
 (pg. 
1635
-
1640
)
157
Garcia-Manero
 
G
Demethylating agents in myeloid malignancies.
Curr Opin Oncol
2008
, vol. 
20
 (pg. 
705
-
710
)
158
Itzykson
 
R
Gardin
 
C
Fenaux
 
P
Meeting report: myelodysplastic syndromes at ASH 2007.
Leukemia
2008
, vol. 
22
 (pg. 
893
-
897
)
159
Bolden
 
JE
Peart
 
MJ
Johnstone
 
RW
Anticancer activities of histone deacetylase inhibitors.
Nat Rev Drug Discov
2006
, vol. 
5
 (pg. 
769
-
784
)
160
Zhang
 
C
Richon
 
V
Ni
 
X
Talpur
 
R
Duvic
 
M
Selective induction of apoptosis by histone deacetylase inhibitor SAHA in cutaneous T-cell lymphoma cells: relevance to mechanism of therapeutic action.
J Invest Dermatol
2005
, vol. 
125
 (pg. 
1045
-
1052
)
161
Sakajiri
 
S
Kumagai
 
T
Kawamata
 
N
Saitoh
 
T
Said
 
JW
Koeffler
 
HP
Histone deacetylase inhibitors profoundly decrease proliferation of human lymphoid cancer cell lines.
Exp Hematol
2005
, vol. 
33
 (pg. 
53
-
61
)
162
O'Connor
 
OA
Heaney
 
ML
Schwartz
 
L
, et al. 
Clinical experience with intravenous and oral formulations of the novel histone deacetylase inhibitor suberoylanilide hydroxamic acid in patients with advanced hematologic malignancies.
J Clin Oncol
2006
, vol. 
24
 (pg. 
166
-
173
)
163
Duvic
 
M
Talpur
 
R
Ni
 
X
, et al. 
Phase 2 trial of oral vorinostat (suberoylanilide hydroxamic acid, SAHA) for refractory cutaneous T-cell lymphoma (CTCL).
Blood
2007
, vol. 
109
 (pg. 
31
-
39
)
164
Kuendgen
 
A
Gattermann
 
N
Valproic acid for the treatment of myeloid malignancies.
Cancer
2007
, vol. 
110
 (pg. 
943
-
954
)
165
Cimino
 
G
Lo-Coco
 
F
Fenu
 
S
, et al. 
Sequential valproic acid/all-trans retinoic acid treatment reprograms differentiation in refractory and high-risk acute myeloid leukemia.
Cancer Res
2006
, vol. 
66
 (pg. 
8903
-
8911
)
166
Garcia-Manero
 
G
Assouline
 
S
Cortes
 
J
, et al. 
Phase 1 study of the oral isotype specific histone deacetylase inhibitor MGCD0103 in leukemia.
Blood
2008
, vol. 
112
 (pg. 
981
-
989
)
167
Druker
 
BJ
Translation of the Philadelphia chromosome into therapy for CML.
Blood
2008
, vol. 
112
 (pg. 
4808
-
4817
)
168
Ciardiello
 
F
Tortora
 
G
EGFR antagonists in cancer treatment.
N Engl J Med
2008
, vol. 
358
 (pg. 
1160
-
1174
)
169
Stegmaier
 
K
Corsello
 
SM
Ross
 
KN
Wong
 
JS
Deangelo
 
DJ
Golub
 
TR
Gefitinib induces myeloid differentiation of acute myeloid leukemia.
Blood
2005
, vol. 
106
 (pg. 
2841
-
2848
)
170
Boehrer
 
S
Ades
 
L
Galluzzi
 
L
, et al. 
Erlotinib and gefitinib for the treatment of myelodysplastic syndrome and acute myeloid leukemia: a preclinical comparison.
Biochem Pharmacol
2008
, vol. 
76
 (pg. 
1417
-
1425
)
171
Boehrer
 
S
Ades
 
L
Braun
 
T
, et al. 
Erlotinib exhibits antineoplastic off-target effects in AML and MDS: a preclinical study.
Blood
2008
, vol. 
111
 (pg. 
2170
-
2180
)
172
Chan
 
G
Pilichowska
 
M
Complete remission in a patient with acute myelogenous leukemia treated with erlotinib for non small-cell lung cancer.
Blood
2007
, vol. 
110
 (pg. 
1079
-
1080
)
173
Kawamata
 
N
Ogawa
 
S
Zimmermann
 
M
, et al. 
Molecular allelokaryotyping of pediatric acute lymphoblastic leukemias by high-resolution single nucleotide polymorphism oligonucleotide genomic microarray.
Blood
2008
, vol. 
111
 (pg. 
776
-
784
)
174
Ley
 
TJ
Mardis
 
ER
Ding
 
L
, et al. 
DNA sequencing of a cytogenetically normal acute myeloid leukaemia genome.
Nature
2008
, vol. 
456
 (pg. 
66
-
72
)
175
Takahashi
 
K
Yamanaka
 
S
Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors.
Cell
2006
, vol. 
126
 (pg. 
663
-
676
)
176
Jaenisch
 
R
Young
 
R
Stem cells, the molecular circuitry of pluripotency and nuclear reprogramming.
Cell
2008
, vol. 
132
 (pg. 
567
-
582
)
177
Maherali
 
N
Sridharan
 
R
Xie
 
W
, et al. 
Directly reprogrammed fibroblasts show global epigenetic remodeling and widespread tissue contribution.
Cell Stem Cell
2007
, vol. 
1
 (pg. 
55
-
70
)
178
Okita
 
K
Ichisaka
 
T
Yamanaka
 
S
Generation of germline-competent induced pluripotent stem cells.
Nature
2007
, vol. 
448
 (pg. 
313
-
317
)
179
Wernig
 
M
Meissner
 
A
Foreman
 
R
, et al. 
In vitro reprogramming of fibroblasts into a pluripotent ES-cell-like state.
Nature
2007
, vol. 
448
 (pg. 
318
-
324
)
180
Hanna
 
J
Markoulaki
 
S
Schorderet
 
P
, et al. 
Direct reprogramming of terminally differentiated mature B lymphocytes to pluripotency.
Cell
2008
, vol. 
133
 (pg. 
250
-
264
)
181
Hanna
 
J
Wernig
 
M
Markoulaki
 
S
, et al. 
Treatment of sickle cell anemia mouse model with iPS cells generated from autologous skin.
Science
2007
, vol. 
318
 (pg. 
1920
-
1923
)
182
Wernig
 
M
Zhao
 
JP
Pruszak
 
J
, et al. 
Neurons derived from reprogrammed fibroblasts functionally integrate into the fetal brain and improve symptoms of rats with Parkinson's disease.
Proc Natl Acad Sci U S A
2008
, vol. 
105
 (pg. 
5856
-
5861
)
183
Xu
 
D
Alipio
 
Z
Yang
 
J
, et al. 
Phenotypic correction of hemophilia a using an Ips-based cellular therapy [abstract].
Blood
2008
, vol. 
112
  
Abstract 514
184
Takayama
 
N
Eto
 
K
Nakauchi
 
H
Yamanaka
 
S
Generation of blood cells from human Ips cells in vitro through the hematopoietic progenitors concentrated within the unique structures. Ips-Sac [abstract].
Blood
2008
, vol. 
112
  
Abstract 1992
185
Alipio
 
Z
Xu
 
D
Yang
 
J
, et al. 
Reprogrammed murine fibroblasts differentiated into hematopoietic progenitors are able to successfully engraft and repopulate the bone marrow [abstract].
Blood
2008
, vol. 
112
  
Abstract 389
186
Brambrink
 
T
Foreman
 
R
Welstead
 
GG
, et al. 
Sequential expression of pluripotency markers during direct reprogramming of mouse somatic cells.
Cell Stem Cell
2008
, vol. 
2
 (pg. 
151
-
159
)
Sign in via your Institution