Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Feb 23.
Published in final edited form as: Biochem Pharmacol. 2010 Apr 9;80(5):602–612. doi: 10.1016/j.bcp.2010.04.003

The bone marrow microenvironment as a sanctuary for minimal residual disease in CML

Rajesh R Nair 1, Joel Tolentino 1, Lori A Hazlehurst 1,*
PMCID: PMC3285111  NIHMSID: NIHMS344497  PMID: 20382130

Abstract

Bcr-abl kinase inhibitors have provided proof of principal that targeted therapy holds great promise for the treatment of cancer. However, despite the success of these agents in treating chronic myelogenous leukemia (CML), the majority of patients continue to present with minimal residual disease contained within the bone marrow microenvironment. These clinical observations suggest that the bone marrow microenvironment may provide survival signals that contribute to the failure to eliminate minimal residual disease. The bone marrow microenvironment is comprised of multiple sub-domains which vary in cellular composition and gradients of soluble factors and matrix composition. Experimental evidence indicates that exposure of tumor cells to either bone marrow derived soluble factors or matrixes can confer a multi-drug resistance phenotype. Together, these data indicate that targeting such pathways may be a viable approach for increasing the efficacy of chemotherapy. Moreover, we propose that personalized medicine must go beyond understanding predictive models inherent to tumors but rather build predictive models that consider diversity in response due to interactions with the tumor microenvironment. This review will focus on CML, however, understanding the contribution of the bone marrow microenvironment could contribute to rationale combination therapy in other types of leukemia, multiple myeloma and solid tumors which metastasize to the bone.

Keywords: Chronic myeloid leukemia, Drug resistance, Tumor microenvironment, CAM-DR, BCR-ABL

1. Introduction

Chronic Myeloid leukemia (CML) is a hematopoietic stem cell disease that is characterized by the existence of the BCR-ABL fusion oncogene resulting from the reciprocal translocation between chromosomes 9 and 22 (t(9;22)(q34;q11)) [1,2]. This single chromosomal abnormality called the Philadelphia chromosome (Ph) is the cytogenetic hallmark of more than 90% CML cases and about 25–30% acute lymphatic leukemia (ALL) cases. The expression of BCR-ABL oncogene, a chimeric bcr-abl protein with deregulated tyrosine kinase activity, is necessary and sufficient for the pathogenesis of CML [3]. The bcr-abl oncoprotein, unlike the primarily nuclear c-abl, is distributed throughout the cytoplasm and interacts with various proteins involved in signal transduction pathways leading to deregulated proliferation, differentiation and survival [4]. Progression of CML follows three distinct phases, namely the early indolent chronic phase, the intermediate accelerated phase and the terminal blast phase. The treatment goal for CML patients is to maintain remission and prevent relapse and disease progression to accelerated phase or blast phase.

STI571 or Imatinib mesylate (IM) was the first molecularly targeted therapy that was rationally designed to specifically inhibit bcr-abl tyrosine kinase activity [5]. IM interacts with the ATP-binding site of bcr-abl, thereby suppressing downstream signaling. The recent phase 3 International Randomized Study of Interferon and STI571 (IRIS) trial has confirmed the long-term efficacy and safety of IM, and it is currently the recommended first line therapy for CML in chronic phase [6-9]. However, despite the effectiveness and good tolerability of IM, drug resistance does emerge due to both intrinsic/innate and extrinsic/acquired factors. Intrinsic factors include point mutations in the abl-kinase domain, BCR-ABL gene amplification, BCR-ABL overexpression, clonal evolution and persistence of CML stem cells [10-13]. Extrinsic factors include decreased drug bioavailability, microenvironmental factors like cytokines, and growth factors leading to decreased apoptosis and increased survival, independent of bcr-abl mediated signaling and development of multidrug resistance in sanctuary sites like the bone marrow (BM) [12].

To overcome IM resistance in CML, more selective second generation bcr-abl kinase inhibitors like Nilotinib (NI) and a dual kinase inhibitor of bcr-abl and src kinases called Dasatinib (DA) were developed [14,15]. However, in clinical practice NI and DA failed to achieve sustained remission in IM-resistant CML blast crisis and ALL [16,17]. Hence, even though the targeted therapy against bcr-abl kinase induces hematologic and cytogenetic remission in chronic phase CML patients, the majority of patients still harbor minimal residual disease (MRD) that, if unchecked leads to relapse of CML.

For many hematopoietic tumors, including CML, following drug treatment, MRD is found in the BM compartment. Indeed, measuring Wilms’ tumor 1 (WT1) gene transcript a marker for the diagnosis of MRD, has demonstrated that the WT1 expression level in peripheral blood was approximately 10 times lower than that in BM, regardless of the type of leukemia [18]. Consistent with this observation is the finding that bcr-abl transcripts in BM samples exceed the levels in the corresponding peripheral blood samples by at least 1 log unit in patients with bcr-abl positive ALL [19]. These clinical observations suggest that the BM microenvironment provides a very important protective sanctuary that contributes to drug resistance. In the following review, we describe the BM microenvironment including its various components and their functions, followed by a brief description of how the CML cells interact and manipulate this BM microenvironment to evade cell death and maintain MRD. Finally, we enumerate the various strategies utilized to overcome MRD, and thus potentially add to the arsenal of therapeutic tools that can be utilized along with standard therapy in order to seek a cure against CML.

2. Bone marrow architecture/microenvironment

Bone marrow (BM) is a hematopoietic organ that resides within the protected confines of the bones and is the major location for hematopoiesis. In adults, 4.6% of body weight is due to the BM that is distributed throughout the vertebrae, ribs, pelvis, skull and proximal ends of the long bones [20]. The fundamental objective of the BM is to maintain the numbers of differentiated hematopoietic cells in the peripheral blood at a constant level throughout the life time of a human being. The BM does this by producing an estimated 500 billion cells per day [21].

The structure of the BM consists of a rigid bone cortex enclosing a cavity containing the arterial vascular system, a complex sinusoidal system, hematopoietic cells and the stroma. Arterial vessels enter the BM and divide into arterioles and capillaries that span throughout the BM. These small vessels then supply sinusoids, which are radially distributed around a draining central sinus which in turn follows into the emissary vein and finally joins the systemic venous circulation [21]. The BM sinusoids are unique, in that they are completely devoid of supporting connective tissue and, instead, the sinusoidal wall is made up of an overlapping single layer of endothelial cells [22]. Because of the lack of connective tissues, the major supporting structure for the sinusoids is the surrounding BM cells. Even with this blood supply, direct measurement of oxygen levels has revealed that BM, in general, is hypoxic (1–2% O2) [23]. The cellular component of the BM can be divided into hematopoietic cells and mesenchymal-derived cells. The mesenchymal-derived cells, along with the macrophages and the extracellular matrix (ECM), form the BM stroma. The BM hematopoietic cells consist of hematopoietic stem cells (HSCs), hematopoietic progenitor cells (HPCs) and mature plasma cells. The HSCs and HPCs are held within the BM stroma not only through the adhesive interactions between VCAM-1, expressed by stromal cells, and integrin α4β1 expressed by the HSCs and HPCs [24,25], but also by the chemotactic interaction between the chemokine CXCL12 and its receptor CXCR4 expressed on the HSCs and HPCs [26,27]. Interestingly, BM is also innervated by myelinated and unmyelinated nerve fibers [28]. The following sections will discuss the components of BM and describe how it may contribute to tumor cell survival and drug resistance.

3. Hematopoietic stem cell niche

HSCs have the ability to differentiate into all myeloid and lymphoid cell lineages and reproduce the entire hematopoietic and immune systems. In steady-state conditions the majority of the HSCs reside in the BM with a few circulating in the peripheral blood [29,30]. Schofield identified the first HSCs in the BM, and he later proposed that HSCs needed to be localized in a particular niche within the BM to maintain their self-renewal capacity [31,32]. Indeed a large body of evidence suggests that the HSCs are localized in a nonrandom manner at the bone–BM interface (Osteoblastic niche) and around the blood vessels (vascular niche) [33-36]. In contrast, the more committed progenitors accumulate in the center of the BM [33,34].

3.1. Osteoblastic niche

The majority of HSCs in the endosteum are found to be in direct contact with the osteoblasts, and there is genetic evidence that increased bone formation leads to an increased number of HSCs in the BM [37,38]. For example, transgenic mice carrying an inducible deletion of the bone morphogenic protein receptor 1A, or mice expressing a constitutively active mutant of the parathyroid hormone receptor 1 in the osteoblasts, result in an increased number of osteoblasts and doubling of HSCs despite the fact that these mice have reduced BM cavity due to increased bone formation [35,37,38]. On the other hand, Visnjic et al. showed that in transgenic mice carrying the herpesvirus thymidine kinase suicide gene driven by an osteoblast-specific promoter, such that administration of ganciclovir specifically induces osteoblast death, a decrease in osteoblasts results in a reduction of HSCs in the BM [39]. These studies demonstrate the role of osteoblasts in the maintenance and survival of HSCs.

The HSCs home in on the endosteum with the help of a chemotactic Ca2+ sensing receptor that senses the Ca2+ gradient within the BM. The gradient is created by the release of Ca2+ in the BM due to the constitutive osteoclasts and osteoblasts-mediated bone remodeling [40]. On reaching the endosteum, the HSCs are held in the endosteum by the expression of adhesive molecules like osteopontin, N-cadherin, transmembrane c-KIT ligand stem cell factor and the polysaccharide hyaluronic acid [40-43] (Fig. 1). In addition to this, the interaction is strengthened by the cytokines and chemokines that are produced by the osteoblasts, like chemokine CXCL12, angiopoietin-1, the ligand for the tyrosine kinase receptor Tie-2 expressed by HSCs, and jagged-1, a ligand for Notch1 receptor also expressed on HSCs [35,37,42]. Finally, these interactions regulate HSC self-renewal by controlling the survival, quiescence and proliferation of HSCs.

Fig. 1.

Fig. 1

Schematic representation of the molecular interactions between hematopoietic stem cells and their niches in the bone marrow. In normal physiological conditions, the hematopoietic stem cells (HSCs) reside either at the endosteum, lodged with the osteoblasts (OBT) (osteoblastic niche) or at the sinusoidal vessels (SV) (vascular niche). The osteoblastic niche provides the microenvironment for HSC maintenance and quiescence and the vascular niche provides the microenvironment for HSC prolirferation and differentiation. The oxygen and the calcium gradient might play a crucial role in maintenance of the different niches within the bone marrow.

3.2. Vascular niche

In addition to the osteoblastic niche, a vascular niche was identified in the BM by using the signaling lymphocyte activation molecule (SLAM) family of receptors (CD150+CD48CD41) [44]. The SLAM family are cell-surface proteins that act as adhesion and signaling receptors and are differentially expressed among hematopoietic progenitors in a way that correlates with progenitor primitiveness [45]. Interestingly, the HSCs in the vascular niche are devoid of N-cadherins and were specifically located adjacent to cells expressing high levels of CXCL12, which in turn surrounded the sinusoidal endothelial cells. The importance of CXCL12-CXCR4 interaction in the maintenance of the vascular niche was demonstrated by showing that a depletion of CXCR4 led to a reduction of the HSC pool [46]. One of the major physiological differences between the two niches is the high oxygen levels in the vascular niche compared to the hypoxic region of the osteoblastic niche [47] (Fig. 1). It has been postulated that the quiescent HSC located in the hypoxic region of the osteoblastic niche, on stimulation, could move to the vascular niche where the increased oxygen levels would induce the HSCs to undergo differentiation and provide the peripheral blood supply with the necessary mature hematopoietic cells [48]. When the differentiated hematopoietic cells are not required, HSCs move back to the endosteal niche and revert back to their original quiescent state, there by maintaining a constant number of differentiated hematopoietic cells in the peripheral blood [48].

4. Cancer stem cells and drug resistance

Cancer stem cells have been described in a number of cancers including acute myeloid leukemia (AML), breast cancer, brain cancer, colon cancer, pancreatic cancer and CML [49-54]. CML, is a stem cell derived hematopoietic cancer that can be effectively treated with the use of bcr-abl kinase inhibitors. However, failure to eliminate MRD found in the BM compartment due to the existence of the CML stem cells (that closely resemble normal HSCs), leads to the relapse of the disease [55]. In vitro, it was demonstrated that osteoblasts or osteoblast-like cells seeded on the bio-derived bone scaffold could maintain stem/progenitor cells from CML bone marrow primary cells in a long term culture [56]. In vivo, the CML stem cells utilize CD44 to home into and subsequently engraft in to the BM HSC niche [57]. Once within the BM, the CML stem cells utilize the survival pathways that are active in normal HSCs.

In CML stem cells, the resistance to apoptosis begins with quiescence, which favors survival against bcr-abl tyrosine kinase inhibitors that rely on cell cycle-dependent apoptotic machinery to induce cell death [58]. This was nicely demonstrated in an in vitro study where quiescent IM-resistant CML cells were rendered sensitive to IM by simply inducing cell cycle progression [59]. Tissue specific-stem cells, including HSCs, have an inbuilt characteristic of up regulating genes that translate to proteins involved in drug efflux and detoxification. Not surprisingly, when compared to mature CML cells, CML stem cells have decreased levels of the organic cation transporter-1 (OCT-1), a transporter involved in the active uptake of IM, and increased levels of drug-efflux-related surface molecules, including the multi-drug transporter, MDR1 [60,61]. Furthermore, it has been shown that pretreatment OCT-1 expression, but not expression of drug efflux transporter, was the most powerful predictor of complete cytogenic response achievement in CML [62].

Moreover, CML stem cell populations not only have higher BCR-ABL transcript levels than mature CML cells, but they are also composed of different subclones carrying different drug resistance bcr-abl kinase domain mutation [63-67]. Thus, the progeny of CML stem cells, containing wild type bcr-abl, initially succumb to therapy with bcr-abl kinase inhibitors. However, with longer treatment there can be the emergence of bcr-abl kinase mutants through the process of subclone selection which in turn leads to drug resistance [61,64].

Another important feature of HSCs is their capacity for self-renewal through the Wnt/β-catenin and hedgehog activation pathways [68,69]. Sonic hedgehog was shown to induce expansion of CML stem cells and the Wnt/β-catenin pathway was found to be activated in granulocyte–macrophage progenitors, which resulted in the acquisition of ‘stemness’ in these CML cells [70,71]. This property of self-renewal ensures that there is always a reservoir of CML stem cells that can be activated to proceed and maintain a CML disease state. In summary, in CML a complete cure can only be established by abolishing MRD which in turn can only be achieved by the eradication of all the CML stem cells and their sub clone population. However, since the pathway for survival and self-renewal in HSCs and the CML stem cells are the same, elucidating the differentially regulated pathway will be critical for devising therapies that eradicate CML stem cells while sparing normal stem cells.

5. Overcoming cancer stem cell mediated drug resistance

Most of the conventional drugs currently used target actively dividing cells and not the quiescent cancer stem cells. Thus, even if the tumor burden is decreased after chemotherapy because of death in the actively dividing tumor cells, drug therapy leaves the quiescent tumorogenic cancer stem cells to survive leading to the reemergence of the disease state. In light of this, an attractive strategy would be to coax the cancer stem cells into cell division. As proof of principle, Ito et al. showed an improvement in the treatment outcome of CML by enhancing the cycling of quiescent leukemia stem cells with the induction of oxidative stress, using Arsenic trioxide (As2O3) [72]. In this study, not only did the number of quiescent leukemic stem cells decrease significantly but more of the stem cells entered the cell cycle when compared to normal stem cells following As2O3 treatment. The mechanism for this observation was attributed to the As2O3-induced degradation of promyelocytic leukemia protein (PML), which was found to be essential for HSC maintenance. Furthermore, in vivo addition of cytosine arabinoside (Ara-C) following As2O3 treatment induced significant cell death in the leukemic stem cell population resulting in complete remission of CML in recipient mice in a serial transplantation assay [72]. Surprisingly, Holtz et al. has shown that combining IM with As2O3 or Ara-C did not induce apoptosis in non-proliferating CML CD34+ progenitors, indicating the use of caution when combining different drug regimens [73].

In addition to inducing cell cycling, oxidative stress also has an effect on the long term maintenance of HSCs. For example, repression of p16lnk4a and p19Arf senescence pathways promotes HSCs cell survival and self-renewal [74,75]. Oxidative stress can affect the long term HSC capacity for survival and self-renewal thorough the upregulation of tumor suppressors p16lnk4a and p19Arf. Treatment with a p38 mitogen-activated protein kinase (MAPK) inhibitor or the antioxidant N-acetyl-l-cysteine blocked the oxidative stress-induced increase of p16lnk4a and p19Arf [76,77]. This suggests that oxidative stress induces the HSC-specific phosphorylation of p38 MAPK and this activation leads to defects in the maintenance of HSC capacity for self-renewal. However, it should be noted that oxidative stress does not differentiate between normal HSCs and CML stem cells, and a targeted therapy involving conjugation of oxidative stress inducing molecule with an antibody that specifically recognize CML stem cells might hold promise. An interesting candidate for targeted therapy may be interleukin (IL)-3 receptors, which is expressed in high abundance on the leukemic stem and very rarely on normal HSCs [78]. Another strategy to specifically target the CML stem cells was suggested in a study by Dierks et al., wherein they show that Smo, a seven-transmembrane domain receptor protein involved in the hedgehog pathway, is specifically upregulated in CML cells and is essential for the expansion of the CML stem cell pool [79]. Interestingly, pharmacological inhibiton of Smo did not impact regular hematopoiesis, but it drastically reduced the CML stem cells in vivo. These results indicate that Smo, is a specific and ‘druggable’ target for CML stem cells that can help eradicate CML.

IM-mediated inhibition of the bcr-abl kinase activity in CML cells leads to suppression of all bcr-abl downstream signaling. This in turn induces a pro-apoptotic milieu that overcomes the pro-survival signaling of bcr-abl, ultimately resulting in cell death [80,81]. But, CML stem cells have high transcripts of BCR-ABL and resistance can arise due to inability of the drug to inhibit all the bcr-abl kinase activity. Dose escalation of bcr-abl kinase inhibitors might be a clinical strategy to circumvent this problem. Indeed, retrospective analysis of IM patient dose escalation in the IRIS trials supports this strategy [82]. Unfortunately, a more thorough investigation, namely the Tyrosine Kinase Inhibitor Optimization and Selectivity (TOPS) study demonstrated that at the end of one year there was no significant difference in the cytogenic response between patients getting the standard dose of IM (400 mg/day) and the higher dose of IM (800 mg/ml) [83]. Switching to a more potent drug than IM, such as NI (30-fold more potent) or DA (325-fold more potent), is an option in IM-resistant CML patients [84,85].

The phenomenon of the ‘emergence’ of CML stem cells with point mutations in the ATP kinase domain is exhibited in patients receiving IM treatment [13,86]. Of the more than 40 different mutations that have been described, the most worrisome mutation is T315I, which renders the CML cell insensitive to any of the clinically available BCR-ABL kinase inhibitors [87-90]. Multi-kinase inhibitors like MK-0457 (VX-680), a potent aurora kinase and JAK2 inhibitor, PHA-739358 a pan-auroroa kinase inhibitor, and DSA compounds which are potent inhibitors of ABL, BCR-ABL T315I and src have shown very encouraging results in CML cells carrying the T315I mutations [91-93]. A novel compound referred to as DCC-2036 is a switch pocket inhibitor and binds to a regulatory site distant from the catalytic kinase domain of ABL and thus successfully circumvents the binding issues with mutations in the kinase domain [94]. Similarly, a selective allosteric bcr-abl inhibitor, GNF-2 and its N-hydroxyethyl carboxamide analogue, GNF-5 were shown to display enhanced inhibitory activity against T315I mutant in cellular assays and also in a murine BM transplantation model, when used in combination with IM [95]. Future and ongoing clinical trials will tell whether any of these agents will help eradicate CML.

The intracellular concentrations of bcr-abl kinase inhibitors in the CML stem cells can be increased by combining them with modulators of drug transporters. The efficiency of such combination in reversing drug resistance was demonstrated by Illmer et al., who showed using a novel high-performance liquid chromatography-based method, that intracellular levels of IM decreases in P-glycoprotein (Pgp)-positive leukemic cells. However, modulation of Pgp by cyclosporin A readily restored IM cytotoxicity in these cells [96]. Also, OCT-1 has been demonstrated to have a significant impact on the drug availability through inhibition of IM influx [97]. Since neither NI nor DA’s cellular uptake is significantly affected by OCT-1 activity, examining the OCT-1 status of the patient prior to start of therapy may lead to a better clinical outcome of the patient [98,99].

Before the advent of bcr-abl tyrosine kinase inhibitors, inter-feron-α (IFNα) was used to treat CML with variable efficiency. Interestingly, IFNα is the only anti-leukemic drug that has more effect on CML stem cells than the conventional bcr-abl tyrosine kinase inhibitors. Whereas IFNα has little effect against the bulk of CML tumor burden, in vitro it can induce CML stem cells to differentiate [100]. To date, the only long term sustainable remission has come from IFNα therapy. For example, 8% of patients treated with IFNα alone have maintained long-term disease-free status without any therapy [101]. Also, patients treated with IFNα who reached complete cytogenic response with follow up IM treatment remained disease free after more than 2 years without any type of therapy [102]. Taken together, these data indicate that a combination of IFNα with bcr-abl tyrosine kinase inhibitors would perhaps eliminate CML stem cells and thus eradicate CML.

Allogenic HSC transplant can also eliminate the primary refractory CML stem cells. HSC transplant in patients with drug resistant CML is the only established stem cell eradicating therapy in CML [103]. It has been hypothesized that in drug resistant patients one of the therapies detailed above can be utilized to induce remission which can then be followed by an allogenic transplant. After transplantation, these patients could be continued on a maintenance therapy utilizing novel bcr-abl tyrosine kinase inhibitors. However, this hypothesis and the success of such therapies remain to be investigated.

6. Bone marrow stroma

The BM stroma is a heterogeneous population of cells that as a whole contribute to hematopoiesis. The different components of the BM stroma include adipocytes, reticular cells, macrophages, vascular endothelial cells, smooth muscle cells and mesenchymal stromal cells (MSCs) [104]. This heterogeneous population is responsible for not only the production and deposition of a complex extracellular matrix (ECM), but also for the production and concentration of cytokines, chemokines and growth factors. The hematopoietic cells interact with the stroma through specific cell surface receptors, secreted growth factors and with adhesive ligands present on the ECM or stromal cells (Fig. 2). Thus, the BM stroma creates a permissive environment in which the process of myelopoiesis, lymphopoiesis and immunoregulation can occur in close proximity [105].

Fig. 2.

Fig. 2

Schematic representation of the mechanism for the stroma-mediated resistance in the bone marrow. The tumor cells within the bone marrow (BM) is bathed in a mileu of cytokines (IL-6, IL-11, IL-1, TNF-α, IFNs) growth factors (VEGF, FGF, G-CSF, GM-CSF, TGFs) and chemokine factors (SCF, SDF-1) secreted by the BM stromal cells (paracrine) or the tumor cells (autocrine). In addition, the tumor cells interact with the stromal compartment with the help of integrins (VLA-4, ICAM, VCAM) and N-cadherins. All this interactions culminates in the activation of signaling pathways (MAPK, JAK/STAT, PI3K-Akt) within the tumor cells which ultimately leads to the proliferation and survival of the tumor cell in the BM stroma.

6.1. Mesenchymal stromal cells (MSCs)

Mesenchymal stromal cells (MSCs) are an important component of the BM stroma and are estimated to represent 0.01–0.0001% of the nucleated cells in adult human BM [106]. MSCs were first identified by Friedenstein et al. who described it as undifferentiated cells in the BM which were able to differentiate in vitro into bone- and cartilage-like colonies [107]. Later studies showed that MSCs were able to differentiate into various cell types, including osteoblasts, chondroblasts, adipocytes, myocytes, astrocytes, neurons, endothelial cells and lung epithelial cells [108-110]. Because of their plasticity they are often referred to as mesenchymal stem cells. However for this review we will refer to them as MSCs in accordance with the recommendation of the International Society for Cellular Therapy [111]. MSCs are characterized by the expression of a large number of adhesion molecules and stromal cell markers (CD73, CD105, CD44, CD29, CD90) in the absence of hematopoietic markers (CD34, CD45, CD14) and endothelial markers (CD34, CD31, von Willebrand factor) [112,113].

One of the main functions of MSCs is to provide, within the BM, a microenvironment that is conducive for hematopoiesis. It does so by not only differentiating into the various cellular components of the BM stroma, but also by providing a scaffold for cell to cell interactions and by secreting cytokines and growth factors. MSCs rely on the transforming growth factor-β (TGF-β) family of proteins in governing its development and differentiation. Even though TGF-β itself promotes the expansion of undifferentiated MSCs, other TGF-β family members in cooperation with specific transcriptional regulators promote the differentiation of MSCs. For example, transcription factor Runx2 drives MSCs to differentiate into osteoblasts and the transcription factor peroxisome proliferator-activated receptor (PPARγ) drives the differentiation of MSCs into adipocytes [114,115]. Interestingly, the transcriptional coactivator PDZ-binding motif (TAZ) acts as a key modulator that ensures that this differentiation occurs independent of each other [116]. In vitro, TAZ was demonstrated to co-activate Runx2-dependent transcription while repressing PPARγ-dependent transcription thus ensuring only the differentiation of MSCs into osteoblasts.

Almost all the information of cytokine production from MSCs is derived from in vitro cell culture studies. MSCs have been shown to secrete cytokines and growth factors such as macrophage-CSF, Flt-3L, stem cell factor (SCF), interleukins like IL-6, IL-7, IL-11, IL-12, IL-14 and IL-15 [117,118]. MSCs are also considered the major source for the secretion of the homeostatic chemokine, CXCL12 which has a role in both homing of the circulating HSCs and in immune responses [119]. MSCs can also be stimulated to induce secretion of cytokines. For example, upon stimulation with IL-1α, MSCs have been demonstrated to secrete IL-1α, leukemia inhibitory factor, G-CSF and GM-CSF [120]. MSCs play a fundamental role in lymphopoiesis by aiding the development of the naïve B cells in the BM. The developmental process involves the interaction of MSCs with naïve B cells and providing the B cells with SCF and IL-7 [121,122]. MSCs are also implicated in the development of T cells via the process of extrathymic T cell lymphopoiesis [123]. MSCs, through the production of CXCL12, assists in the homing of CD8+ T cells into the BM and thus mediate the induction of adaptive immunity in response to antigen [124].

In contrast, MSCs can also exert immunosuppressive activity through not only suppression of T cell proliferation but also by negatively modulating the functions of B cells, natural killer cells and dendritic cells [125-129]. Not surprisingly, the immunosuppressive properties of MSCs were utilized by Le Blanc et al. to treat a young patient with a severe grade IV acute graft-vs-host disease [130]. The exact mechanism for MSC-mediated immunosuppression remains to be clarified. However, in vitro, soluble factors like hepatocyte growth factor, prostaglandin E2, TGF-β1, indoleamine 2,3-dioxygenase and IL-10 have all been suggested to play a role [131,132].

6.2. Extracellular matrix (ECM)

The main constituents of the bone marrow ECM include fibronectin (FN), hualuronan, collagen types I and IV, laminin, glycosaminoglycans heparin sulfate and chondroitin sulfate. In addition to providing the structural scaffold for cellular elements of the BM, ECM also exerts its effect on cell growth, differentiation, motility and viability. The ECM molecules function as a reservoir of many secreted growth factors, cytokines, matrix metalloproteinases and other processing enzymes whose availability is regulated by arrangement of the matrix, thus creating stromal niches within the BM [133].

The BM cells sense their location within the BM through specific interactions within the ECM. Such interactions are very important for cell homeostasis, since inappropriate cell/ECM interaction leads to cell death through anoikis [134]. Most of the interactions between the ECM and the BM cells occur through specific cell-surface receptors named integrins. The integrins are composed of two distinct subunits, the α (18 α subunits) and the β (8 β subunits) subunits which form at least 24 different combinations of α/β sub units in humans [135]. The integrins function not only as mechano-transducers that transform mechanical forces created by the ECM into chemical signals, but they can also cross-talk with receptor tyrosine kinases in response to growth factors. Interestingly, integrins have the ability to communicate over the plasma membrane in both directions, distinguished as outside-in and inside-out signaling [136,137]. Thus, integrin signaling regulates many cellular processes such as proliferation, survival, migration and invasion.

7. Bone marrow stroma and drug resistance

Paget proposed the concept that the microenvironment of a developing tumor is a crucial regulator of tumor growth in his famous ‘seed and soil’ hypothesis [138]. This is clearly demonstrated in the study showing that BM stroma is abnormal in CML, as it has a reduced ability to support growth of normal progenitors cultured in contact with the stroma [139]. In contrast, the growth and proliferation of CML progenitors in contact with CML stroma was unimpaired, again showing that abnormal interactions between the CML cells and the stroma play an instrumental part in promoting survival and maintenance of MRD. This observation was further validated by the finding showing that leukemic cells create an abnormal stromal niche by secreting SCF. The high concentration of SCF directs the normal HSC to home into this niche which in turn makes the maintenance and self-renewal capacity of the normal HSC highly defective [140]. Another stromal niche is created due to the secretion of the chemokine CXCL12 within the BM by the stromal cells, predominantly MSCs [46]. Circulating CML cells, including the CML stem cells express CXCR4, which is the designated chemokine receptor for CXCL12 [141,142]. CML cells use this receptor for homing into the BM and thus evade cell death induced by bcr-abl tyrosine kinase inhibitors. Indeed, Vianello et al. have shown that MSCs in the BM protect and preserve the CML progenitor cells from IM-induced cell death via the CXCR4–CXCL12 axis [143]. Thus by manipulating the BM stroma, the CML cells control the population and fate of normal HSCs.

Cytokine signaling from the BM microenvironment can also lead to leukemic cell survival in the presence of bcr-abl kinase inhibitors. For example, Williams et al. have shown in their study that in spite of efficient inhibition of bcr-abl kinase activity in the tumor cells, mice with BCR-ABL-expressing Arf-null ALL are resistant to IM treatment [144]. The failure of the leukemic cells to respond was determined to be due to IL-7-mediated rescue within the host hematopoietic microenvironment by enabling cells to remain in cycle even when exposed to micromolar IM concentration. Interestingly, when the same resistant leukemic cells where isolated from the in vivo microenvironment and exposed to IM in vitro they became sensitive indicating that resistance was reversible [144]. Our lab addressed the contribution of soluble factors derived from the BM microenvironment in mediating resistance to bcr-abl kinase inhibitors in CML [145]. We showed that stable soluble factors secreted by HS-5 stromal cell line were sufficient to cause resistance to IM, NI and DA in CML cells. Specifically, soluble factors-mediated activation of STAT3 in the CML cells was sufficient to cause resistance to bcr-abl kinase inhibitors and may contribute to the failure of bcr-abl kinase inhibitors to eradicate MRD associated with CML [145]. Similarly, Wang et al. showed that conditioned media of IM-resistant LAMA84 cells protected IM-naïve LAMA cells and CML progenitors from IM- or NI-induced cell death [146]. The protection was due to the GM-CSF (from the conditioned media of IM-resistant LAMA84 cells) -mediated activation of janus kinases 2 (JAK-2)/signal transducer and activator of transcription 5 (STAT5) signaling pathway. Furthermore, in the same study they demonstrated that IM-resistant patient samples showed elevated mRNA and protein levels of GM-CSF, thus suggesting that GM-CSF also contributed to the IM and NI resistance in vivo.

The phenomenon of cell adhesion-mediated drug resistance (CAM-DR) has been observed in various tumor cells [147-149]. In CML cells, adhesion to FN via VLA-5 integrin induces CAM-DR [150]. In this study, K562 cells adhered to FN via the VLA-5 integrin, which in turn provided significant resistance to apoptosis induced by a number of DNA damaging agents including melphalan, mitoxantrone and γ-irradiation. In addition, cellular adhesion to FN also inhibited apoptosis in CML cells exposed to IM [150]. Decreased proliferation in CML cells adhered to FN via integrins might be one of the contributing factors for the manifestation of drug resistance [151]. Interestingly, in CML cells, β1 integrin mediated adhesion reduced the levels of Bim, a proapoptotic protein that counteracts the effects of antiapoptototic proteins Bcl-2 and Bcl-xL [152]. Furthermore, this decrease in Bim levels was attributed to increased degradation of Bim protein in CML cells adhered to FN. The importance of this finding was divulged in studies showing that incubation of CML cells with bcr-abl tyrosine kinase inhibitors results in cell death that is partially mediated by an increase in levels of Bim expression [153,154]. Thus, CML binding to the FN in the BM stroma might evade cell death induced by bcr-abl kinase inhibitor treatment through down regulation of Bim [152].

8. Overcoming bone marrow stroma-mediated drug resistance

Strategies to overcome BM stroma-mediated drug resistance should include not only disrupting the interaction between the CML cells and the BM stromal cells, but also, inhibit the signaling pathways mediated by cytokines and growth factors in the BM. Signals derived from the BM stroma can effectively reconstitute the downstream signaling pathway of bcr-abl protein, such that CML cells can achieve bcr-abl independent growth in the BM, making them resistant to bcr-abl tyrosine kinase inhibitors. Targeting the downstream pathway in such cases may offer an alternative pathway to induce cell death in resistant CML cells. For example, the PI3K-Akt-mTOR pathway is shown to be activated in CML cells after IM treatment and this activation was shown to be important in mediating cell survival, thereby resulting in the development of resistance [155]. This same pathway leads to the over expression of vascular endothelial growth factor in the CML BM, which in turn increases the microvascular density and the vascular niche within the BM and thus may be essential for the growth of CML progenitors [156]. Studies have shown that RAD001, a derivative of rapamycin, orthe PI3-kinase inhibitor, LY294002, significantly inhibits the PI3K-mTOR pathway resulting in, not only inhibition of the growth of CML cells, but also prolonging the survival of mice transplanted with BM retrovirally transduced with BCR-ABL [156,157]. Similarly, bcr-abl is known to couple with Ras pathway and play a critical role in the oncogenic transformation in CML. Ras activation requires the post-translational modification, involving the addition of 15-carbon farnesyl isoprenoid moiety to a conserved cysteine residue in a carboxy-terminal CAAX motif. This whole process called prenylation is catalyzed by farnesyltransferase (FT). FT inhibitors (FTI) like tipifarnib and lonafarnib were rationally designed drugs designed to target FT activity. In IM-resistant CML patients, tipifarnib showed modest activity when used either alone or in combination with IM [158,159]. Additionally, lonafarnib demonstrated selective activity against primary cells from patients with CML and also showed efficacy in a murine model of CML [160]. Finally, inhibitors of MAPK kinases (MEK1/2), PD18435273 and U0126, when combined with IM or DA, results in synergistic cell death in IM-resistant CML cell lines [161]. However, this effect was not reproduced in IM-resistant cell lines with T315I mutation in the ABL kinase domain, thereby limiting its utility [162].

Heat shock protein 90 (Hsp90) is a molecular chaperone that interacts and stabilizes signal transducing proteins like Raf, Akt and bcr-abl. 17-allylamino-17-demethoxygeldanamycin (17-AGG), a protein that binds Hsp90 and inhibits its ability to act as a chaperone results in down regulation of Raf, Akt and bcr-abl mediated downstream signaling leading to apoptosis in CML cell lines [163]. Also, 17-AGG targets the P-glycoprotein multi-drug resistance pump and has been successfully used in combination with IM to synergistically inhibit cell growth and induce cell death in IM-resistant CML cell lines [164]. Increased proteosomal activity in CML cells leads to resistance to apoptosis by decreasing the levels of pro-apoptotic protein Bim [152-154]. To overcome this, proteosome inhibitors like bortezomib have been used in IM-resistant CML cell lines to successfully inhibit proliferation, induce cell cycle arrest and promote apoptosis in these cells [165]. Growth factors and the cytokines within the BM stroma can synergize with bcr-abl activity to regulate the cell cycling of the CML cells via the cyclin-dependent kinases. Two approaches have been utilized to inhibit this pathway, namely the use of either histone deacetylase inhibitors (HDI) or cyclin-dependent kinase inhibitors. Inhibtion of histone deacetylase by HDI causes the hyperacetylation of histones resulting in transcriptional upregulation of cyclin-dependent kinase inhibitor p21 leading to cell-arrest and apoptosis in CML cells [166]. HDI like suberoylanilide hydroxamic acid (SAHA), LBH589 and valproate have been used in combination with IM and NI and have shown to enhance BCR-ABL tyrosine kinase inhibitor-induced growth arrest and apoptosis in CML cells and in IM-resistant cell lines expressing T315I mutation [167-170]. Also, the multiple cyclin-dependent kinase inhibitor, falvopiridol has been used in combination with IM to cause selective apoptosis in CML and IM-resistant CML cells [168].

The CXCR4–CXCL12 axis functions as a major migration mechanism for CML cells to home into the BM. Use of CXCR4 antagonist offers an attractive approach to mobilize the CML cells from the BM and hence make them sensitive to bcr-abl tyrosine kinase inhibitor. One of such inhibitors that has been tested in CML is the bicyclam non-peptide antoagonist AMD3100. AMD3100 restored the sensitivity of CML progenitors cells to cell death by IM [143,171]. More importantly, CML progenitor burden within the BM was drastically reduced with the treatment with AMD3100 [143]. However, further studies have to be performed to warrant the use of inhibitors of the CXCR4–CXCL12 axis in combination with bcr-abl tyrosine kinase inhibitors.

Production of cytokines like IL-3, G-CSF and GM-CSF within the BM has been suggested to play a role in CML pathogenesis [146,172,173]. Studies have shown these cytokines have a role in development of resistance towards bcr-abl kinase inhibitors via the JAK/STAT pathway [145,146]. Indeed, use of JAK-2 inhibitor, AG490 reduced the GM-CSF-induced STAT5 activation and thus reversed the potential of GM-CSF to confer IM or NI resistance in CML [146]. Similarly, our lab has shown that reducing STAT3 levels with siRNA sensitized HS-5 stromal cell conditioned media-cultured BCR-ABL kinase inhibitor resistant K562 cells to IM, NI and DA [145]. Finally, it remains to be tested if the CAM-DR seen in CML can be reversed either by a VLA-4 inhibitor or by a novel integrin inhibitor like LFA703 [174]. In summary, we propose that effective testing of combination therapies should include strategies to test in multicellular models that include components of the bone marrow.

9. Conclusion

Survival among the CML patients is closely linked to the reduction of BCR-ABL positive cell burden in patients. Since the presence of bcr-abl protein can be cytogenetically identified by the presence of Philadelphia chromosome, the goal of treatment in CML is to effectively eradicate Philadelphia chromosome-positive cells. The current targeted molecular therapy utilizing bcr-abl kinase inhibitors is effective in inducing a complete hematological and cytogenic remission in a high percentage of patients, but the problem of resistance and intolerance to these agents still exists. We propose that personalized medicine must go beyond an understanding of tumor derived pathways and begin to delineate how variability within the context of tumor cell-microenvironment interactions may predict drug response and lead to the optimization of combination therapy. Taken together, this knowledge will provide the blueprint for a rational combination of drugs that is tailored to a specific individual patient, with the goal of helping overcome emergence of drug resistance and leading to a better patient outcome.

Abbreviations

CML

chronic myeloid leukemia

ALL

acute lymphatic leukemia

IM

imatinib mesylate

BM

bone marrow

NI

nilotinib

DA

dasatinib

MRD

minimal residual disease

WT1

Wilms’ tumor 1

HSC

hematopoietic stem cell

HPC

hematopoietic progenitor cell

ECM

extracellular matrix

MSC

mesenchymal stromal cell

AML

acute myeloid leukemia

SLAM

signaling lymphocyte activation molecule

SDF-1

stromal cell-derived factor-1

OCT-1

organic cation transporter-1

Pgp

P-glycoprotein

GMP

granulocyte–macrophage progenitors

CAM-DR

cell adhesion-mediated drug resistance

SCF

stem cell factor

IL

interleukin

INFα

interferon-α

TGF-β

transforming growth factor-β

References

  • 1.Rowley JD. Letter: a new consistent chromosomal abnormality in chronic myelogenous leukaemia identified by quinacrine fluorescence and Giemsa staining. Nature. 1973;243:290–3. doi: 10.1038/243290a0. [DOI] [PubMed] [Google Scholar]
  • 2.Bartram CR, de Klein A, Hagemeijer A, van Agthoven T, Geurts van Kessel A, Bootsma D, et al. Translocation of c-ab1 oncogene correlates with the presence of a Philadelphia chromosome in chronic myelocytic leukaemia. Nature. 1983;306:277–80. doi: 10.1038/306277a0. [DOI] [PubMed] [Google Scholar]
  • 3.Heisterkamp N, Stephenson JR, Groffen J, Hansen PF, de Klein A, Bartram CR, et al. Localization of the c-ab1 oncogene adjacent to a translocation break point in chronic myelocytic leukaemia. Nature. 1983;306:239–42. doi: 10.1038/306239a0. [DOI] [PubMed] [Google Scholar]
  • 4.Sattler M, Griffin JD. Molecular mechanisms of transformation by the BCR-ABL oncogene. Semin Hematol. 2003;40:4–10. doi: 10.1053/shem.2003.50034. [DOI] [PubMed] [Google Scholar]
  • 5.Druker BJ, Tamura S, Buchdunger E, Ohno S, Segal GM, Fanning S, et al. Effects of a selective inhibitor of the Abl tyrosine kinase on the growth of Bcr-Abl positive cells. Nat Med. 1996;2:561–6. doi: 10.1038/nm0596-561. [DOI] [PubMed] [Google Scholar]
  • 6.Baccarani M, Saglio G, Goldman J, Hochhaus A, Simonsson B, Appelbaum F, et al. Evolving concepts in the management of chronic myeloid leukemia: recommendations from an expert panel on behalf of the European LeukemiaNet. Blood. 2006;108:1809–20. doi: 10.1182/blood-2006-02-005686. [DOI] [PubMed] [Google Scholar]
  • 7.Druker BJ, Talpaz M, Resta DJ, Peng B, Buchdunger E, Ford JM, et al. Efficacy and safety of a specific inhibitor of the BCR-ABL tyrosine kinase in chronic myeloid leukemia. N Engl J Med. 2001;344:1031–7. doi: 10.1056/NEJM200104053441401. [DOI] [PubMed] [Google Scholar]
  • 8.Hochhaus A, Druker B, Sawyers C, Guilhot F, Schiffer CA, Cortes J, et al. Favorable long-term follow-up results over 6 years for response, survival, and safety with imatinib mesylate therapy in chronic-phase chronic myeloid leukemia after failure of interferon-alpha treatment. Blood. 2008;111:1039–43. doi: 10.1182/blood-2007-07-103523. [DOI] [PubMed] [Google Scholar]
  • 9.Hochhaus A, O’Brien SG, Guilhot F, Druker BJ, Branford S, Foroni L, et al. Six-year follow-up of patients receiving imatinib for the first-line treatment of chronic myeloid leukemia. Leukemia. 2009;23:1054–61. doi: 10.1038/leu.2009.38. [DOI] [PubMed] [Google Scholar]
  • 10.Hochhaus A, Kreil S, Corbin AS, La Rosee P, Muller MC, Lahaye T, et al. Molecular and chromosomal mechanisms of resistance to imatinib (STI571) therapy. Leukemia. 2002;16:2190–6. doi: 10.1038/sj.leu.2402741. [DOI] [PubMed] [Google Scholar]
  • 11.Hochhaus A, La Rosee P. Imatinib therapy in chronic myelogenous leukemia: strategies to avoid and overcome resistance. Leukemia. 2004;18:1321–31. doi: 10.1038/sj.leu.2403426. [DOI] [PubMed] [Google Scholar]
  • 12.Apperley JF. Part I: mechanisms of resistance to imatinib in chronic myeloid leukaemia. Lancet Oncol. 2007;8:1018–29. doi: 10.1016/S1470-2045(07)70342-X. [DOI] [PubMed] [Google Scholar]
  • 13.Cortes JE, Talpaz M, Giles F, O’Brien S, Rios MB, Shan J, et al. Prognostic significance of cytogenetic clonal evolution in patients with chronic myelogenous leukemia on imatinib mesylate therapy. Blood. 2003;101:3794–800. doi: 10.1182/blood-2002-09-2790. [DOI] [PubMed] [Google Scholar]
  • 14.Weisberg E, Manley PW, Breitenstein W, Bruggen J, Cowan-Jacob SW, Ray A, et al. Characterization of AMN107, a selective inhibitor of native and mutant Bcr-Abl. Cancer Cell. 2005;7:129–41. doi: 10.1016/j.ccr.2005.01.007. [DOI] [PubMed] [Google Scholar]
  • 15.Shah NP, Tran C, Lee FY, Chen P, Norris D, Sawyers CL. Overriding imatinib resistance with a novel ABL kinase inhibitor. Science. 2004;305:399–401. doi: 10.1126/science.1099480. [DOI] [PubMed] [Google Scholar]
  • 16.Kantarjian H, Giles F, Wunderle L, Bhalla K, O’Brien S, Wassmann B, et al. Nilotinib in imatinib-resistant CML and Philadelphia chromosome-positive ALL. N Engl J Med. 2006;354:2542–51. doi: 10.1056/NEJMoa055104. [DOI] [PubMed] [Google Scholar]
  • 17.Talpaz M, Shah NP, Kantarjian H, Donato N, Nicoll J, Paquette R, et al. Dasatinib in imatinib-resistant Philadelphia chromosome-positive leukemias. N Engl J Med. 2006;354:2531–41. doi: 10.1056/NEJMoa055229. [DOI] [PubMed] [Google Scholar]
  • 18.Inoue K, Ogawa H, Yamagami T, Soma T, Tani Y, Tatekawa T, et al. Long-term follow-up of minimal residual disease in leukemia patients by monitoring WT1 (Wilms tumor gene) expression levels. Blood. 1996;88:2267–78. [PubMed] [Google Scholar]
  • 19.Martin H, Atta J, Bruecher J, Elsner S, Schardt C, Stadler M, et al. In patients with BCR-ABL-positive ALL in CR peripheral blood contains less residual disease than bone marrow: implications for autologous BMT. Ann Hematol. 1994;68:85–7. doi: 10.1007/BF01715137. [DOI] [PubMed] [Google Scholar]
  • 20.Lord BI, Testa NG, Hendry JH. The relative spatial distributions of CFUs and CFUc in the normal mouse femur. Blood. 1975;46:65–72. [PubMed] [Google Scholar]
  • 21.Fliedner TM, Graessle D, Paulsen C, Reimers K. Structure and function of bone marrow hemopoiesis: mechanisms of response to ionizing radiation exposure. Cancer Biother Radiopharm. 2002;17:405–26. doi: 10.1089/108497802760363204. [DOI] [PubMed] [Google Scholar]
  • 22.Tavassoli M. Structure and function of sinusoidal endothelium of bone marrow. Prog Clin Biol Res. 1981;59B:249–56. [PubMed] [Google Scholar]
  • 23.Cipolleschi MG, Dello Sbarba P, Olivotto M. The role of hypoxia in the maintenance of hematopoietic stem cells. Blood. 1993;82:2031–7. [PubMed] [Google Scholar]
  • 24.Scott LM, Priestley GV, Papayannopoulou T. Deletion of alpha4 integrins from adult hematopoietic cells reveals roles in homeostasis, regeneration, and homing. Mol Cell Biol. 2003;23:9349–60. doi: 10.1128/MCB.23.24.9349-9360.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Ulyanova T, Scott LM, Priestley GV, Jiang Y, Nakamoto B, Koni PA, et al. VCAM-1 expression in adult hematopoietic and nonhematopoietic cells is controlled by tissue-inductive signals and reflects their developmental origin. Blood. 2005;106:86–94. doi: 10.1182/blood-2004-09-3417. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Broxmeyer HE, Orschell CM, Clapp DW, Hangoc G, Cooper S, Plett PA, et al. Rapid mobilization of murine and human hematopoietic stem and progenitor cells with AMD3100, a CXCR4 antagonist. J Exp Med. 2005;201:1307–18. doi: 10.1084/jem.20041385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Liles WC, Broxmeyer HE, Rodger E, Wood B, Hubel K, Cooper S, et al. Mobilization of hematopoietic progenitor cells in healthy volunteers by AMD3100, a CXCR4 antagonist. Blood. 2003;102:2728–30. doi: 10.1182/blood-2003-02-0663. [DOI] [PubMed] [Google Scholar]
  • 28.Weiss RA, Root WS. Innervation of the vessels of the marrow cavity of certain bones. Am J Physiol. 1959;197:1255–7. doi: 10.1152/ajplegacy.1959.197.6.1255. [DOI] [PubMed] [Google Scholar]
  • 29.Abkowitz JL, Robinson AE, Kale S, Long MW, Chen J. Mobilization of hematopoietic stem cells during homeostasis and after cytokine exposure. Blood. 2003;102:1249–53. doi: 10.1182/blood-2003-01-0318. [DOI] [PubMed] [Google Scholar]
  • 30.Wright DE, Wagers AJ, Gulati AP, Johnson FL, Weissman IL. Physiological migration of hematopoietic stem and progenitor cells. Science. 2001;294:1933–6. doi: 10.1126/science.1064081. [DOI] [PubMed] [Google Scholar]
  • 31.Schofield R. A comparative study of the repopulating potential of grafts from various haemopoietic sources: CFU repopulation. Cell Tissue Kinet. 1970;3:119–30. doi: 10.1111/j.1365-2184.1970.tb00259.x. [DOI] [PubMed] [Google Scholar]
  • 32.Schofield R. The relationship between the spleen colony-forming cell and the haemopoietic stem cell. Blood Cells. 1978;4:7–25. [PubMed] [Google Scholar]
  • 33.Lo Celso C, Wu JW, Lin CP. In vivo imaging of hematopoietic stem cells and their microenvironment. J Biophotonics. 2009;2:619–31. doi: 10.1002/jbio.200910072. [DOI] [PubMed] [Google Scholar]
  • 34.Nilsson SK, Johnston HM, Coverdale JA. Spatial localization of transplanted hemopoietic stem cells: inferences for the localization of stem cell niches. Blood. 2001;97:2293–9. doi: 10.1182/blood.v97.8.2293. [DOI] [PubMed] [Google Scholar]
  • 35.Calvi LM, Adams GB, Weibrecht KW, Weber JM, Olson DP, Knight MC, et al. Osteoblastic cells regulate the haematopoietic stem cell niche. Nature. 2003;425:841–6. doi: 10.1038/nature02040. [DOI] [PubMed] [Google Scholar]
  • 36.Kiel MJ, Radice GL, Morrison SJ. Lack of evidence that hematopoietic stem cells depend on N-cadherin-mediated adhesion to osteoblasts for their maintenance. Cell Stem Cell. 2007;1:204–17. doi: 10.1016/j.stem.2007.06.001. [DOI] [PubMed] [Google Scholar]
  • 37.Arai F, Hirao A, Ohmura M, Sato H, Matsuoka S, Takubo K, et al. Tie2/angiopoietin-1 signaling regulates hematopoietic stem cell quiescence in the bone marrow niche. Cell. 2004;118:149–61. doi: 10.1016/j.cell.2004.07.004. [DOI] [PubMed] [Google Scholar]
  • 38.Zhang J, Niu C, Ye L, Huang H, He X, Tong WG, et al. Identification of the haematopoietic stem cell niche and control of the niche size. Nature. 2003;425:836–41. doi: 10.1038/nature02041. [DOI] [PubMed] [Google Scholar]
  • 39.Visnjic D, Kalajzic Z, Rowe DW, Katavic V, Lorenzo J, Aguila HL. Hematopoiesis is severely altered in mice with an induced osteoblast deficiency. Blood. 2004;103:3258–64. doi: 10.1182/blood-2003-11-4011. [DOI] [PubMed] [Google Scholar]
  • 40.Adams GB, Chabner KT, Alley IR, Olson DP, Szczepiorkowski ZM, Poznansky MC, et al. Stem cell engraftment at the endosteal niche is specified by the calcium-sensing receptor. Nature. 2006;439:599–603. doi: 10.1038/nature04247. [DOI] [PubMed] [Google Scholar]
  • 41.Nilsson SK, Haylock DN, Johnston HM, Occhiodoro T, Brown TJ, Simmons PJ. Hyaluronan is synthesized by primitive hemopoietic cells, participates in their lodgment at the endosteum following transplantation, and is involved in the regulation of their proliferation and differentiation in vitro. Blood. 2003;101:856–62. doi: 10.1182/blood-2002-05-1344. [DOI] [PubMed] [Google Scholar]
  • 42.Semerad CL, Christopher MJ, Liu F, Short B, Simmons PJ, Winkler I, et al. G-CSF potently inhibits osteoblast activity and CXCL12 mRNA expression in the bone marrow. Blood. 2005;106:3020–7. doi: 10.1182/blood-2004-01-0272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Stier S, Ko Y, Forkert R, Lutz C, Neuhaus T, Grunewald E, et al. Osteopontin is a hematopoietic stem cell niche component that negatively regulates stem cell pool size. J Exp Med. 2005;201:1781–91. doi: 10.1084/jem.20041992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Kiel MJ, Yilmaz OH, Iwashita T, Yilmaz OH, Terhorst C, Morrison SJ. SLAM family receptors distinguish hematopoietic stem and progenitor cells and reveal endothelial niches for stem cells. Cell. 2005;121:1109–21. doi: 10.1016/j.cell.2005.05.026. [DOI] [PubMed] [Google Scholar]
  • 45.Engel P, Eck MJ, Terhorst C. The SAP and SLAM families in immune responses and X-linked lymphoproliferative disease. Nat Rev Immunol. 2003;3:813–21. doi: 10.1038/nri1202. [DOI] [PubMed] [Google Scholar]
  • 46.Sugiyama T, Kohara H, Noda M, Nagasawa T. Maintenance of the hematopoietic stem cell pool by CXCL12-CXCR4 chemokine signaling in bone marrow stromal cell niches. Immunity. 2006;25:977–88. doi: 10.1016/j.immuni.2006.10.016. [DOI] [PubMed] [Google Scholar]
  • 47.Parmar K, Mauch P, Vergilio JA, Sackstein R, Down JD. Distribution of hematopoietic stem cells in the bone marrow according to regional hypoxia. Proc Natl Acad Sci USA. 2007;104:5431–6. doi: 10.1073/pnas.0701152104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Iwasaki H, Suda T. Cancer stem cells and their niche. Cancer Sci. 2009;100:1166–72. doi: 10.1111/j.1349-7006.2009.01177.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Al-Hajj M, Wicha MS, Benito-Hernandez A, Morrison SJ, Clarke MF. Prospective identification of tumorigenic breast cancer cells. Proc Natl Acad Sci USA. 2003;100:3983–8. doi: 10.1073/pnas.0530291100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Bonnet D, Dick JE. Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat Med. 1997;3:730–7. doi: 10.1038/nm0797-730. [DOI] [PubMed] [Google Scholar]
  • 51.Dalerba P, Dylla SJ, Park IK, Liu R, Wang X, Cho RW, et al. Phenotypic characterization of human colorectal cancer stem cells. Proc Natl Acad Sci USA. 2007;104:10158–63. doi: 10.1073/pnas.0703478104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Li C, Heidt DG, Dalerba P, Burant CF, Zhang L, Adsay V, et al. Identification of pancreatic cancer stem cells. Cancer Res. 2007;67:1030–7. doi: 10.1158/0008-5472.CAN-06-2030. [DOI] [PubMed] [Google Scholar]
  • 53.Singh SK, Hawkins C, Clarke ID, Squire JA, Bayani J, Hide T, et al. Identification of human brain tumour initiating cells. Nature. 2004;432:396–401. doi: 10.1038/nature03128. [DOI] [PubMed] [Google Scholar]
  • 54.Wang JC, Lapidot T, Cashman JD, Doedens M, Addy L, Sutherland DR, et al. High level engraftment of NOD/SCID mice by primitive normal and leukemic hematopoietic cells from patients with chronic myeloid leukemia in chronic phase. Blood. 1998;91:2406–14. [PubMed] [Google Scholar]
  • 55.Eaves C, Udomsakdi C, Cashman J, Barnett M, Eaves A. The biology of normal and neoplastic stem cells in CML. Leuk Lymphoma. 1993;11(Suppl. 1):245–53. doi: 10.3109/10428199309047894. [DOI] [PubMed] [Google Scholar]
  • 56.Hou L, Liu T, Tan J, Meng W, Deng L, Yu H, et al. Long-term culture of leukemic bone marrow primary cells in biomimetic osteoblast niche. Int J Hematol. 2009 doi: 10.1007/s12185-009-0392-4. [DOI] [PubMed] [Google Scholar]
  • 57.Krause DS, Lazarides K, von Andrian UH, Van Etten RA. Requirement for CD44 in homing and engraftment of BCR-ABL-expressing leukemic stem cells. Nat Med. 2006;12:1175–80. doi: 10.1038/nm1489. [DOI] [PubMed] [Google Scholar]
  • 58.Valent P, Deininger M. Clinical perspectives of concepts on neoplastic stem cells and stem cell-resistance in chronic myeloid leukemia. Leuk Lymphoma. 2008;49:604–9. doi: 10.1080/10428190801923212. [DOI] [PubMed] [Google Scholar]
  • 59.Graham SM, Jorgensen HG, Allan E, Pearson C, Alcorn MJ, Richmond L, et al. Primitive, quiescent, Philadelphia-positive stem cells from patients with chronic myeloid leukemia are insensitive to STI571 in vitro. Blood. 2002;99:319–25. doi: 10.1182/blood.v99.1.319. [DOI] [PubMed] [Google Scholar]
  • 60.Brendel C, Scharenberg C, Dohse M, Robey RW, Bates SE, Shukla S, et al. Imatinib mesylate and nilotinib (AMN107) exhibit high-affinity interaction with ABCG2 on primitive hematopoietic stem cells. Leukemia. 2007;21:1267–75. doi: 10.1038/sj.leu.2404638. [DOI] [PubMed] [Google Scholar]
  • 61.Jiang X, Zhao Y, Smith C, Gasparetto M, Turhan A, Eaves A, et al. Chronic myeloid leukemia stem cells possess multiple unique features of resistance to BCR-ABL targeted therapies. Leukemia. 2007;21:926–35. doi: 10.1038/sj.leu.2404609. [DOI] [PubMed] [Google Scholar]
  • 62.Wang L, Giannoudis A, Lane S, Williamson P, Pirmohamed M, Clark RE. Expression of the uptake drug transporter hOCT1 is an important clinical determinant of the response to imatinib in chronic myeloid leukemia. Clin Pharmacol Ther. 2008;83:258–64. doi: 10.1038/sj.clpt.6100268. [DOI] [PubMed] [Google Scholar]
  • 63.Sorel N, Bonnet ML, Guillier M, Guilhot F, Brizard A, Turhan AG. Evidence of ABL-kinase domain mutations in highly purified primitive stem cell populations of patients with chronic myelogenous leukemia. Biochem Biophys Res Commun. 2004;323:728–30. doi: 10.1016/j.bbrc.2004.08.169. [DOI] [PubMed] [Google Scholar]
  • 64.Roche-Lestienne C, Preudhomme C. Mutations in the ABL kinase domain pre-exist the onset of imatinib treatment. Semin Hematol. 2003;40:80–2. doi: 10.1053/shem.2003.50046. [DOI] [PubMed] [Google Scholar]
  • 65.Chu S, Xu H, Shah NP, Snyder DS, Forman SJ, Sawyers CL, et al. Detection of BCR-ABL kinase mutations in CD34+ cells from chronic myelogenous leukemia patients in complete cytogenetic remission on imatinib mesylate treatment. Blood. 2005;105:2093–8. doi: 10.1182/blood-2004-03-1114. [DOI] [PubMed] [Google Scholar]
  • 66.Copland M, Hamilton A, Elrick LJ, Baird JW, Allan EK, Jordanides N, et al. Dasatinib (BMS-354825) targets an earlier progenitor population than imatinib in primary CML but does not eliminate the quiescent fraction. Blood. 2006;107:4532–9. doi: 10.1182/blood-2005-07-2947. [DOI] [PubMed] [Google Scholar]
  • 67.Jamieson CH. Chronic myeloid leukemia stem cells. Hematol Am Soc Hematol Educ Program. 2008:436–42. doi: 10.1182/asheducation-2008.1.436. [DOI] [PubMed] [Google Scholar]
  • 68.Reya T, Duncan AW, Ailles L, Domen J, Scherer DC, Willert K, et al. A role for Wnt signalling in self-renewal of haematopoietic stem cells. Nature. 2003;423:409–14. doi: 10.1038/nature01593. [DOI] [PubMed] [Google Scholar]
  • 69.Stein MI, Zhu J, Emerson SG. Molecular pathways regulating the self-renewal of hematopoietic stem cells. Exp Hematol. 2004;32:1129–36. doi: 10.1016/j.exphem.2004.08.012. [DOI] [PubMed] [Google Scholar]
  • 70.Jamieson CH, Ailles LE, Dylla SJ, Muijtjens M, Jones C, Zehnder JL, et al. Granulocyte–macrophage progenitors as candidate leukemic stem cells in blast-crisis CML. N Engl J Med. 2004;351:657–67. doi: 10.1056/NEJMoa040258. [DOI] [PubMed] [Google Scholar]
  • 71.Sengupta A, Banerjee D, Chandra S, Banerji SK, Ghosh R, Roy R, et al. Deregulation and cross talk among Sonic hedgehog, Wnt, Hox and Notch signaling in chronic myeloid leukemia progression. Leukemia. 2007;21:949–55. doi: 10.1038/sj.leu.2404657. [DOI] [PubMed] [Google Scholar]
  • 72.Ito K, Bernardi R, Morotti A, Matsuoka S, Saglio G, Ikeda Y, et al. PML targeting eradicates quiescent leukaemia-initiating cells. Nature. 2008;453:1072–8. doi: 10.1038/nature07016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Holtz MS, Forman SJ, Bhatia R. Nonproliferating CML CD34+ progenitors are resistant to apoptosis induced by a wide range of proapoptotic stimuli. Leukemia. 2005;19:1034–41. doi: 10.1038/sj.leu.2403724. [DOI] [PubMed] [Google Scholar]
  • 74.Molofsky AV, Pardal R, Iwashita T, Park IK, Clarke MF, Morrison SJ. Bmi-1 dependence distinguishes neural stem cell self-renewal from progenitor proliferation. Nature. 2003;425:962–7. doi: 10.1038/nature02060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Park IK, Qian D, Kiel M, Becker MW, Pihalja M, Weissman IL, et al. Bmi-1 is required for maintenance of adult self-renewing haematopoietic stem cells. Nature. 2003;423:302–5. doi: 10.1038/nature01587. [DOI] [PubMed] [Google Scholar]
  • 76.Giafis N, Katsoulidis E, Sassano A, Tallman MS, Higgins LS, Nebreda AR, et al. Role of the p38 mitogen-activated protein kinase pathway in the generation of arsenic trioxide-dependent cellular responses. Cancer Res. 2006;66:6763–71. doi: 10.1158/0008-5472.CAN-05-3699. [DOI] [PubMed] [Google Scholar]
  • 77.Ito K, Hirao A, Arai F, Takubo K, Matsuoka S, Miyamoto K, et al. Reactive oxygen species act through p38 MAPK to limit the lifespan of hematopoietic stem cells. Nat Med. 2006;12:446–51. doi: 10.1038/nm1388. [DOI] [PubMed] [Google Scholar]
  • 78.Florian S, Sonneck K, Hauswirth AW, Krauth MT, Schernthaner GH, Sperr WR, et al. Detection of molecular targets on the surface of CD34+/CD383− stem cells in various myeloid malignancies. Leuk Lymphoma. 2006;47:207–22. doi: 10.1080/10428190500272507. [DOI] [PubMed] [Google Scholar]
  • 79.Dierks C, Beigi R, Guo GR, Zirlik K, Stegert MR, Manley P, et al. Expansion of Bcr-Abl-positive leukemic stem cells is dependent on Hedgehog pathway activation. Cancer Cell. 2008;14:238–49. doi: 10.1016/j.ccr.2008.08.003. [DOI] [PubMed] [Google Scholar]
  • 80.Chu S, Holtz M, Gupta M, Bhatia R. BCR/ABL kinase inhibition by imatinib mesylate enhances MAP kinase activity in chronic myelogenous leukemia CD34+ cells. Blood. 2004;103:3167–74. doi: 10.1182/blood-2003-04-1271. [DOI] [PubMed] [Google Scholar]
  • 81.Deacon K, Mistry P, Chernoff J, Blank JL, Patel R. p38 Mitogen-activated protein kinase mediates cell death and p21-activated kinase mediates cell survival during chemotherapeutic drug-induced mitotic arrest. Mol Biol Cell. 2003;14:2071–87. doi: 10.1091/mbc.E02-10-0653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Kantarjian HM, Larson RA, Guilhot F, O’Brien SG, Mone M, Rudoltz M, et al. Efficacy of imatinib dose escalation in patients with chronic myeloid leukemia in chronic phase. Cancer. 2009;115:551–60. doi: 10.1002/cncr.24066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Cortes JE, Baccarani M, Guilhot F, Druker BJ, Branford S, Kim DW, et al. Phase III, randomized, open-label study of daily imatinib mesylate 400 mg versus 800 mg in patients with newly diagnosed, previously untreated chronic myeloid leukemia in chronic phase using molecular end points: tyrosine kinase inhibitor optimization and selectivity study. J Clin Oncol. 2010;28:424–30. doi: 10.1200/JCO.2009.25.3724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Golemovic M, Verstovsek S, Giles F, Cortes J, Manshouri T, Manley PW, et al. AMN107, a novel aminopyrimidine inhibitor of Bcr-Abl, has in vitro activity against imatinib-resistant chronic myeloid leukemia. Clin Cancer Res. 2005;11:4941–7. doi: 10.1158/1078-0432.CCR-04-2601. [DOI] [PubMed] [Google Scholar]
  • 85.Lombardo LJ, Lee FY, Chen P, Norris D, Barrish JC, Behnia K, et al. Discovery of N-(2-chloro-6-methyl-phenyl)-2-(6-(4-(2-hydroxyethyl)-piperazin-1-yl)-2-methylpyrimidin-4-ylamino)thiazole-5-carboxamide (BMS-354825), a dual Src/Abl kinase inhibitor with potent antitumor activity in preclinical assays. J Med Chem. 2004;47:6658–61. doi: 10.1021/jm049486a. [DOI] [PubMed] [Google Scholar]
  • 86.Jabbour E, Kantarjian H, Jones D, Talpaz M, Bekele N, O’Brien S, et al. Frequency and clinical significance of BCR-ABL mutations in patients with chronic myeloid leukemia treated with imatinib mesylate. Leukemia. 2006;20:1767–73. doi: 10.1038/sj.leu.2404318. [DOI] [PubMed] [Google Scholar]
  • 87.Branford S, Rudzki Z, Walsh S, Parkinson I, Grigg A, Szer J, et al. Detection of BCR-ABL mutations in patients with CML treated with imatinib is virtually always accompanied by clinical resistance, and mutations in the ATP phosphate-binding loop (P-loop) are associated with a poor prognosis. Blood. 2003;102:276–83. doi: 10.1182/blood-2002-09-2896. [DOI] [PubMed] [Google Scholar]
  • 88.Hughes T, Deininger M, Hochhaus A, Branford S, Radich J, Kaeda J, et al. Monitoring CML patients responding to treatment with tyrosine kinase inhibitors: review and recommendations for harmonizing current methodology for detecting BCR-ABL transcripts and kinase domain mutations and for expressing results. Blood. 2006;108:28–37. doi: 10.1182/blood-2006-01-0092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Jabbour E, Kantarjian HM, Jones D, Reddy N, O’Brien S, Garcia-Manero G, et al. Characteristics and outcome of chronic myeloid leukemia patients with F317L BCR-ABL kinase domain mutation after therapy with tyrosine kinase inhibitors. Blood. 2008;112:4839–42. doi: 10.1182/blood-2008-04-149948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Soverini S, Colarossi S, Gnani A, Rosti G, Castagnetti F, Poerio A, et al. Contribution of ABL kinase domain mutations to imatinib resistance in different subsets of Philadelphia-positive patients: by the GIMEMA Working Party on Chronic Myeloid Leukemia. Clin Cancer Res. 2006;12:7374–9. doi: 10.1158/1078-0432.CCR-06-1516. [DOI] [PubMed] [Google Scholar]
  • 91.Giles FJ, Cortes J, Jones D, Bergstrom D, Kantarjian H, Freedman SJ. MK-0457, a novel kinase inhibitor, is active in patients with chronic myeloid leukemia or acute lymphocytic leukemia with the T315I BCR-ABL mutation. Blood. 2007;109:500–2. doi: 10.1182/blood-2006-05-025049. [DOI] [PubMed] [Google Scholar]
  • 92.Gontarewicz A, Balabanov S, Keller G, Colombo R, Graziano A, Pesenti E, et al. Simultaneous targeting of Aurora kinases and Bcr-Abl kinase by the small molecule inhibitor PHA-739358 is effective against imatinib-resistant BCR-ABL mutations including T315I. Blood. 2008;111:4355–64. doi: 10.1182/blood-2007-09-113175. [DOI] [PubMed] [Google Scholar]
  • 93.Seeliger MA, Ranjitkar P, Kasap C, Shan Y, Shaw DE, Shah NP, et al. Equally potent inhibition of c-Src and Abl by compounds that recognize inactive kinase conformations. Cancer Res. 2009;69:2384–92. doi: 10.1158/0008-5472.CAN-08-3953. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Pinilla-Ibarz J, Quintas-Cardama A. New agents in the treatment of chronic myelogenous leukemia. J Natl Compr Canc Netw. 2009;7:1028–37. doi: 10.6004/jnccn.2009.0066. [DOI] [PubMed] [Google Scholar]
  • 95.Zhang J, Adrian FJ, Jahnke W, Cowan-Jacob SW, Li AG, Iacob RE, et al. Targeting Bcr-Abl by combining allosteric with ATP-binding-site inhibitors. Nature. 2010;463:501–6. doi: 10.1038/nature08675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Illmer T, Schaich M, Platzbecker U, Freiberg-Richter J, Oelschlagel U, von Bonin M, et al. P-glycoprotein-mediated drug efflux is a resistance mechanism of chronic myelogenous leukemia cells to treatment with imatinib mesylate. Leukemia. 2004;18:401–8. doi: 10.1038/sj.leu.2403257. [DOI] [PubMed] [Google Scholar]
  • 97.Thomas J, Wang L, Clark RE, Pirmohamed M. Active transport of imatinib into and out of cells: implications for drug resistance. Blood. 2004;104:3739–45. doi: 10.1182/blood-2003-12-4276. [DOI] [PubMed] [Google Scholar]
  • 98.Hiwase DK, Saunders V, Hewett D, Frede A, Zrim S, Dang P, et al. Dasatinib cellular uptake and efflux in chronic myeloid leukemia cells: therapeutic implications. Clin Cancer Res. 2008;14:3881–8. doi: 10.1158/1078-0432.CCR-07-5095. [DOI] [PubMed] [Google Scholar]
  • 99.White DL, Saunders VA, Dang P, Engler J, Zannettino AC, Cambareri AC, et al. OCT-1-mediated influx is a key determinant of the intracellular uptake of imatinib but not nilotinib (AMN107): reduced OCT-1 activity is the cause of low in vitro sensitivity to imatinib. Blood. 2006;108:697–704. doi: 10.1182/blood-2005-11-4687. [DOI] [PubMed] [Google Scholar]
  • 100.Angstreich GR, Matsui W, Huff CA, Vala MS, Barber J, Hawkins AL, et al. Effects of imatinib and interferon on primitive chronic myeloid leukaemia progenitors. Br J Haematol. 2005;130:373–81. doi: 10.1111/j.1365-2141.2005.05606.x. [DOI] [PubMed] [Google Scholar]
  • 101.Talpaz M, Estrov Z, Kantarjian H, Ku S, Foteh A, Kurzrock R. Persistence of dormant leukemic progenitors during interferon-induced remission in chronic myelogenous leukemia Analysis by polymerase chain reaction of individual colonies. J Clin Invest. 1994;94:1383–9. doi: 10.1172/JCI117473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Rousselot P, Huguet F, Rea D, Legros L, Cayuela JM, Maarek O, et al. Imatinib mesylate discontinuation in patients with chronic myelogenous leukemia in complete molecular remission for more than 2 years. Blood. 2007;109:58–60. doi: 10.1182/blood-2006-03-011239. [DOI] [PubMed] [Google Scholar]
  • 103.Gratwohl A, Heim D. Current role of stem cell transplantation in chronic myeloid leukaemia. Best Pract Res Clin Haematol. 2009;22:431–43. doi: 10.1016/j.beha.2009.05.002. [DOI] [PubMed] [Google Scholar]
  • 104.Clark BR, Keating A. Biology of bone marrow stroma. Ann N Y Acad Sci. 1995;770:70–8. doi: 10.1111/j.1749-6632.1995.tb31044.x. [DOI] [PubMed] [Google Scholar]
  • 105.Verfaillie C, Hurley R, Bhatia R, McCarthy JB. Role of bone marrow matrix in normal and abnormal hematopoiesis. Crit Rev Oncol Hematol. 1994;16:201–24. doi: 10.1016/1040-8428(94)90071-x. [DOI] [PubMed] [Google Scholar]
  • 106.Dazzi F, Ramasamy R, Glennie S, Jones SP, Roberts I. The role of mesenchymal stem cells in haemopoiesis. Blood Rev. 2006;20:161–71. doi: 10.1016/j.blre.2005.11.002. [DOI] [PubMed] [Google Scholar]
  • 107.Friedenstein AJ, Deriglasova UF, Kulagina NN, Panasuk AF, Rudakowa SF, Luria EA, et al. Precursors for fibroblasts in different populations of hematopoietic cells as detected by the in vitro colony assay method. Exp Hematol. 1974;2:83–92. [PubMed] [Google Scholar]
  • 108.Al-Khaldi A, Eliopoulos N, Martineau D, Lejeune L, Lachapelle K, Galipeau J. Postnatal bone marrow stromal cells elicit a potent VEGF-dependent neoangiogenic response in vivo. Gene Ther. 2003;10:621–9. doi: 10.1038/sj.gt.3301934. [DOI] [PubMed] [Google Scholar]
  • 109.Pittenger MF, Mackay AM, Beck SC, Jaiswal RK, Douglas R, Mosca JD, et al. Multilineage potential of adult human mesenchymal stem cells. Science. 1999;284:143–7. doi: 10.1126/science.284.5411.143. [DOI] [PubMed] [Google Scholar]
  • 110.Reyes M, Lund T, Lenvik T, Aguiar D, Koodie L, Verfaillie CM. Purification and ex vivo expansion of postnatal human marrow mesodermal progenitor cells. Blood. 2001;98:2615–25. doi: 10.1182/blood.v98.9.2615. [DOI] [PubMed] [Google Scholar]
  • 111.Horwitz EM, Le Blanc K, Dominici M, Mueller I, Slaper-Cortenbach I, Marini FC, et al. Clarification of the nomenclature for MSC: The International Society for Cellular Therapy position statement. Cytotherapy. 2005;7:393–5. doi: 10.1080/14653240500319234. [DOI] [PubMed] [Google Scholar]
  • 112.Kemp KC, Hows J, Donaldson C. Bone marrow-derived mesenchymal stem cells. Leuk Lymphoma. 2005;46:1531–44. doi: 10.1080/10428190500215076. [DOI] [PubMed] [Google Scholar]
  • 113.Roorda BD, ter Elst A, Kamps WA, de Bont ES. Bone marrow-derived cells and tumor growth: contribution of bone marrow-derived cells to tumor microenvironments with special focus on mesenchymal stem cells. Crit Rev Oncol Hematol. 2009;69:187–98. doi: 10.1016/j.critrevonc.2008.06.004. [DOI] [PubMed] [Google Scholar]
  • 114.Nakashima K, de Crombrugghe B. Transcriptional mechanisms in osteoblast differentiation and bone formation. Trends Genet. 2003;19:458–66. doi: 10.1016/S0168-9525(03)00176-8. [DOI] [PubMed] [Google Scholar]
  • 115.Rosen ED, Walkey CJ, Puigserver P, Spiegelman BM. Transcriptional regulation of adipogenesis. Genes Dev. 2000;14:1293–307. [PubMed] [Google Scholar]
  • 116.Hong JH, Hwang ES, McManus MT, Amsterdam A, Tian Y, Kalmukova R, et al. TAZ, a transcriptional modulator of mesenchymal stem cell differentiation. Science. 2005;309:1074–8. doi: 10.1126/science.1110955. [DOI] [PubMed] [Google Scholar]
  • 117.Majumdar MK, Thiede MA, Haynesworth SE, Bruder SP, Gerson SL. Human marrow-derived mesenchymal stem cells (MSCs) express hematopoietic cytokines and support long-term hematopoiesis when differentiated toward stromal and osteogenic lineages. J Hematother Stem Cell Res. 2000;9:841–8. doi: 10.1089/152581600750062264. [DOI] [PubMed] [Google Scholar]
  • 118.Majumdar MK, Thiede MA, Mosca JD, Moorman M, Gerson SL. Phenotypic and functional comparison of cultures of marrow-derived mesenchymal stem cells (MSCs) and stromal cells. J Cell Physiol. 1998;176:57–66. doi: 10.1002/(SICI)1097-4652(199807)176:1<57::AID-JCP7>3.0.CO;2-7. [DOI] [PubMed] [Google Scholar]
  • 119.Baggiolini M. Chemokines and leukocyte traffic. Nature. 1998;392:565–8. doi: 10.1038/33340. [DOI] [PubMed] [Google Scholar]
  • 120.Deans RJ, Moseley AB. Mesenchymal stem cells: biology and potential clinical uses. Exp Hematol. 2000;28:875–84. doi: 10.1016/s0301-472x(00)00482-3. [DOI] [PubMed] [Google Scholar]
  • 121.Torlakovic E, Tenstad E, Funderud S, Rian E. CD10+ stromal cells form B-lymphocyte maturation niches in the human bone marrow. J Pathol. 2005;205:311–7. doi: 10.1002/path.1705. [DOI] [PubMed] [Google Scholar]
  • 122.Milne CD, Fleming HE, Zhang Y, Paige CJ. Mechanisms of selection mediated by interleukin-7, the preBCR, and hemokinin-1 during B-cell development. Immunol Rev. 2004;197:75–88. doi: 10.1111/j.0105-2896.2004.0103.x. [DOI] [PubMed] [Google Scholar]
  • 123.Barda-Saad M, Rozenszajn LA, Globerson A, Zhang AS, Zipori D. Selective adhesion of immature thymocytes to bone marrow stromal cells: relevance to T cell lymphopoiesis. Exp Hematol. 1996;24:386–91. [PubMed] [Google Scholar]
  • 124.Mazo IB, Honczarenko M, Leung H, Cavanagh LL, Bonasio R, Weninger W, et al. Bone marrow is a major reservoir and site of recruitment for central memory CD8+ T cells. Immunity. 2005;22:259–70. doi: 10.1016/j.immuni.2005.01.008. [DOI] [PubMed] [Google Scholar]
  • 125.Le Blanc K, Tammik C, Rosendahl K, Zetterberg E, Ringden O. HLA expression and immunologic properties of differentiated and undifferentiated mesenchymal stem cells. Exp Hematol. 2003;31:890–6. doi: 10.1016/s0301-472x(03)00110-3. [DOI] [PubMed] [Google Scholar]
  • 126.Tse WT, Pendleton JD, Beyer WM, Egalka MC, Guinan EC. Suppression of allogeneic T-cell proliferation by human marrow stromal cells: implications in transplantation. Transplantation. 2003;75:389–97. doi: 10.1097/01.TP.0000045055.63901.A9. [DOI] [PubMed] [Google Scholar]
  • 127.Di Nicola M, Carlo-Stella C, Magni M, Milanesi M, Longoni PD, Matteucci P, et al. Human bone marrow stromal cells suppress T-lymphocyte proliferation induced by cellular or nonspecific mitogenic stimuli. Blood. 2002;99:3838–43. doi: 10.1182/blood.v99.10.3838. [DOI] [PubMed] [Google Scholar]
  • 128.Krampera M, Glennie S, Dyson J, Scott D, Laylor R, Simpson E, et al. Bone marrow mesenchymal stem cells inhibit the response of naive and memory antigen-specific T cells to their cognate peptide. Blood. 2003;101:3722–9. doi: 10.1182/blood-2002-07-2104. [DOI] [PubMed] [Google Scholar]
  • 129.Djouad F, Plence P, Bony C, Tropel P, Apparailly F, Sany J, et al. Immunosuppressive effect of mesenchymal stem cells favors tumor growth in allogeneic animals. Blood. 2003;102:3837–44. doi: 10.1182/blood-2003-04-1193. [DOI] [PubMed] [Google Scholar]
  • 130.Le Blanc K, Rasmusson I, Sundberg B, Gotherstrom C, Hassan M, Uzunel M, et al. Treatment of severe acute graft-versus-host disease with third party haploidentical mesenchymal stem cells. Lancet. 2004;363:1439–41. doi: 10.1016/S0140-6736(04)16104-7. [DOI] [PubMed] [Google Scholar]
  • 131.Aggarwal S, Pittenger MF. Human mesenchymal stem cells modulate allogeneic immune cell responses. Blood. 2005;105:1815–22. doi: 10.1182/blood-2004-04-1559. [DOI] [PubMed] [Google Scholar]
  • 132.Beyth S, Borovsky Z, Mevorach D, Liebergall M, Gazit Z, Aslan H, et al. Human mesenchymal stem cells alter antigen-presenting cell maturation and induce T-cell unresponsiveness. Blood. 2005;105:2214–9. doi: 10.1182/blood-2004-07-2921. [DOI] [PubMed] [Google Scholar]
  • 133.Marastoni S, Ligresti G, Lorenzon E, Colombatti A, Mongiat M. Extracellular matrix: a matter of life and death. Connect Tissue Res. 2008;49:203–6. doi: 10.1080/03008200802143190. [DOI] [PubMed] [Google Scholar]
  • 134.Gilmore AP. Anoikis. Cell Death Differ. 2005;12(Suppl. 2):1473–7. doi: 10.1038/sj.cdd.4401723. [DOI] [PubMed] [Google Scholar]
  • 135.Hynes RO. Integrins: a family of cell surface receptors. Cell. 1987;48:549–54. doi: 10.1016/0092-8674(87)90233-9. [DOI] [PubMed] [Google Scholar]
  • 136.Zhu J, Carman CV, Kim M, Shimaoka M, Springer TA, Luo BH. Requirement of alpha and beta subunit transmembrane helix separation for integrin outside-in signaling. Blood. 2007;110:2475–83. doi: 10.1182/blood-2007-03-080077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Luo BH, Carman CV, Springer TA. Structural basis of integrin regulation and signaling. Annu Rev Immunol. 2007;25:619–47. doi: 10.1146/annurev.immunol.25.022106.141618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Paget S. The distribution of secondary growths in cancer of the breast. 1889. Cancer Metastasis Rev. 1989;8:98–101. [PubMed] [Google Scholar]
  • 139.Bhatia R, McGlave PB, Dewald GW, Blazar BR, Verfaillie CM. Abnormal function of the bone marrow microenvironment in chronic myelogenous leukemia: role of malignant stromal macrophages. Blood. 1995;85:3636–45. [PubMed] [Google Scholar]
  • 140.Colmone A, Amorim M, Pontier AL, Wang S, Jablonski E, Sipkins DA. Leukemic cells create bone marrow niches that disrupt the behavior of normal hematopoietic progenitor cells. Science. 2008;322:1861–5. doi: 10.1126/science.1164390. [DOI] [PubMed] [Google Scholar]
  • 141.Oberlin E, Amara A, Bachelerie F, Bessia C, Virelizier JL, Arenzana-Seisdedos F, et al. The CXC chemokine SDF-1 is the ligand for LESTR/fusin and prevents infection by T-cell-line-adapted HIV-1. Nature. 1996;382:833–5. doi: 10.1038/382833a0. [DOI] [PubMed] [Google Scholar]
  • 142.Peled A, Petit I, Kollet O, Magid M, Ponomaryov T, Byk T, et al. Dependence of human stem cell engraftment and repopulation of NOD/SCID mice on CXCR4. Science. 1999;283:845–8. doi: 10.1126/science.283.5403.845. [DOI] [PubMed] [Google Scholar]
  • 143.Vianello F, Villanova F, Tisato V, Lymperi S, Ho KK, Gomes AR, et al. Bone marrow mesenchymal stromal cells non-selectively protect chronic myeloid leukemia cells from imatinib-induced apoptosis via the CXCR4/CXCL12 axis. Haematologica. 2010 doi: 10.3324/haematol.2009.017178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Williams RT, den Besten W, Sherr CJ. Cytokine-dependent imatinib resistance in mouse BCR-ABL + Arf-null lymphoblastic leukemia. Genes Dev. 2007;21:2283–7. doi: 10.1101/gad.1588607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Bewry NN, Nair RR, Emmons MF, Boulware D, Pinilla-Ibarz J, Hazlehurst LA. Stat3 contributes to resistance toward BCR-ABL inhibitors in a bone marrow microenvironment model of drug resistance. Mol Cancer Ther. 2008;7:3169–75. doi: 10.1158/1535-7163.MCT-08-0314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Wang Y, Cai D, Brendel C, Barett C, Erben P, Manley PW, et al. Adaptive secretion of granulocyte–macrophage colony-stimulating factor (GM-CSF) mediates imatinib and nilotinib resistance in BCR/ABL+ progenitors via JAK-2/STAT-5 pathway activation. Blood. 2007;109:2147–55. doi: 10.1182/blood-2006-08-040022. [DOI] [PubMed] [Google Scholar]
  • 147.Damiano JS, Cress AE, Hazlehurst LA, Shtil AA, Dalton WS. Cell adhesion mediated drug resistance (CAM-DR): role of integrins and resistance to apoptosis in human myeloma cell lines. Blood. 1999;93:1658–67. [PMC free article] [PubMed] [Google Scholar]
  • 148.Damiano JS, Dalton WS. Integrin-mediated drug resistance in multiple myeloma. Leuk Lymphoma. 2000;38:71–81. doi: 10.3109/10428190009060320. [DOI] [PubMed] [Google Scholar]
  • 149.Sethi T, Rintoul RC, Moore SM, MacKinnon AC, Salter D, Choo C, et al. Extracellular matrix proteins protect small cell lung cancer cells against apoptosis: a mechanism for small cell lung cancer growth and drug resistance in vivo. Nat Med. 1999;5:662–8. doi: 10.1038/9511. [DOI] [PubMed] [Google Scholar]
  • 150.Damiano JS, Hazlehurst LA, Dalton WS. Cell adhesion-mediated drug resistance (CAM-DR) protects the K562 chronic myelogenous leukemia cell line from apoptosis induced by BCR/ABL inhibition, cytotoxic drugs, and gamma-irradiation. Leukemia. 2001;15:1232–9. doi: 10.1038/sj.leu.2402179. [DOI] [PubMed] [Google Scholar]
  • 151.Lundell BI, McCarthy JB, Kovach NL, Verfaillie CM. Activation-dependent alpha5beta1 integrin-mediated adhesion to fibronectin decreases proliferation of chronic myelogenous leukemia progenitors and K562 cells. Blood. 1996;87:2450–8. [PubMed] [Google Scholar]
  • 152.Hazlehurst LA, Argilagos RF, Dalton WS. Beta1 integrin mediated adhesion increases Bim protein degradation and contributes to drug resistance in leukaemia cells. Br J Haematol. 2007;136:269–75. doi: 10.1111/j.1365-2141.2006.06435.x. [DOI] [PubMed] [Google Scholar]
  • 153.Luciano F, Jacquel A, Colosetti P, Herrant M, Cagnol S, Pages G, et al. Phosphorylation of Bim-EL by Erk1/2 on serine 69 promotes its degradation via the proteasome pathway and regulates its proapoptotic function. Oncogene. 2003;22:6785–93. doi: 10.1038/sj.onc.1206792. [DOI] [PubMed] [Google Scholar]
  • 154.Kuribara R, Honda H, Matsui H, Shinjyo T, Inukai T, Sugita K, et al. Roles of Bim in apoptosis of normal and Bcr-Abl-expressing hematopoietic progenitors. Mol Cell Biol. 2004;24:6172–83. doi: 10.1128/MCB.24.14.6172-6183.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Burchert A, Wang Y, Cai D, von Bubnoff N, Paschka P, Muller-Brusselbach S, et al. Compensatory PI3-kinase/Akt/mTor activation regulates imatinib resistance development. Leukemia. 2005;19:1774–82. doi: 10.1038/sj.leu.2403898. [DOI] [PubMed] [Google Scholar]
  • 156.Mayerhofer M, Valent P, Sperr WR, Griffin JD, Sillaber C. BCR/ABL induces expression of vascular endothelial growth factor and its transcriptional activator, hypoxia inducible factor-1alpha, through a pathway involving phosphoinositide 3-kinase and the mammalian target of rapamycin. Blood. 2002;100:3767–75. doi: 10.1182/blood-2002-01-0109. [DOI] [PubMed] [Google Scholar]
  • 157.Dengler J, vonBubnoff N, Decker T, Peschel C, Duyster J. Combination of imatinib with rapamycin or RAD001 acts synergistically only in Bcr-Abl-positive cells with moderate resistance to imatinib. Leukemia. 2005;19:1835–8. doi: 10.1038/sj.leu.2403848. [DOI] [PubMed] [Google Scholar]
  • 158.Cortes J, Albitar M, Thomas D, Giles F, Kurzrock R, Thibault A, et al. Efficacy of the farnesyl transferase inhibitor R115777 in chronic myeloid leukemia and other hematologic malignancies. Blood. 2003;101:1692–7. doi: 10.1182/blood-2002-07-1973. [DOI] [PubMed] [Google Scholar]
  • 159.Cortes J, Quintas-Cardama A, Garcia-Manero G, O’Brien S, Jones D, Faderl S, et al. Phase 1 study of tipifarnib in combination with imatinib for patients with chronic myelogenous leukemia in chronic phase after imatinib failure. Cancer. 2007;110:2000–6. doi: 10.1002/cncr.23006. [DOI] [PubMed] [Google Scholar]
  • 160.Peters DG, Hoover RR, Gerlach MJ, Koh EY, Zhang H, Choe K, et al. Activity of the farnesyl protein transferase inhibitor SCH66336 against BCR/ABL-induced murine leukemia and primary cells from patients with chronic myeloid leukemia. Blood. 2001;97:1404–12. doi: 10.1182/blood.v97.5.1404. [DOI] [PubMed] [Google Scholar]
  • 161.Yu C, Krystal G, Varticovksi L, McKinstry R, Rahmani M, Dent P, et al. Pharmacologic mitogen-activated protein/extracellular signal-regulated kinase kinase/mitogen-activated protein kinase inhibitors interact synergistically with STI571 to induce apoptosis in Bcr/Abl-expressing human leukemia cells. Cancer Res. 2002;62:188–99. [PubMed] [Google Scholar]
  • 162.Nguyen TK, Rahmani M, Harada H, Dent P, Grant S. MEK1/2 inhibitors sensitize Bcr/Abl+ human leukemia cells to the dual Abl/Src inhibitor BMS-354/825. Blood. 2007;109:4006–15. doi: 10.1182/blood-2006-09-045039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Nimmanapalli R, O’Bryan E, Bhalla K. Geldanamycin and its analogue 17-allylamino-17-demethoxygeldanamycin lowers Bcr-Abl levels and induces apoptosis and differentiation of Bcr-Abl-positive human leukemic blasts. Cancer Res. 2001;61:1799–804. [PubMed] [Google Scholar]
  • 164.Radujkovic A, Schad M, Topaly J, Veldwijk MR, Laufs S, Schultheis BS, et al. Synergistic activity of imatinib and 17-AAG in imatinib-resistant CML cells overexpressing BCR-ABL—inhibition of P-glycoprotein function by 17-AAG. Leukemia. 2005;19:1198–206. doi: 10.1038/sj.leu.2403764. [DOI] [PubMed] [Google Scholar]
  • 165.Gatto S, Scappini B, Pham L, Onida F, Milella M, Ball G, et al. The proteasome inhibitor PS-341 inhibits growth and induces apoptosis in Bcr/Abl-positive cell lines sensitive and resistant to imatinib mesylate. Haematologica. 2003;88:853–63. [PubMed] [Google Scholar]
  • 166.Nimmanapalli R, Fuino L, Stobaugh C, Richon V, Bhalla K. Cotreatment with the histone deacetylase inhibitor suberoylanilide hydroxamic acid (SAHA) enhances imatinib-induced apoptosis of Bcr-Abl-positive human acute leukemia cells. Blood. 2003;101:3236–9. doi: 10.1182/blood-2002-08-2675. [DOI] [PubMed] [Google Scholar]
  • 167.Yu C, Rahmani M, Almenara J, Subler M, Krystal G, Conrad D, et al. Histone deacetylase inhibitors promote STI571-mediated apoptosis in STI571-sensitive and -resistant Bcr/Abl+ human myeloid leukemia cells. Cancer Res. 2003;63:2118–26. [PubMed] [Google Scholar]
  • 168.Yu C, Krystal G, Dent P, Grant S. Flavopiridol potentiates STI571-induced mitochondrial damage and apoptosis in BCR-ABL-positive human leukemia cells. Clin Cancer Res. 2002;8:2976–84. [PubMed] [Google Scholar]
  • 169.Fiskus W, Pranpat M, Bali P, Balasis M, Kumaraswamy S, Boyapalle S, et al. Combined effects of novel tyrosine kinase inhibitor AMN107 and histone deacetylase inhibitor LBH589 against Bcr-Abl-expressing human leukemia cells. Blood. 2006;108:645–52. doi: 10.1182/blood-2005-11-4639. [DOI] [PubMed] [Google Scholar]
  • 170.Morotti A, Cilloni D, Messa F, Arruga F, Defilippi I, Carturan S, et al. Valproate enhances imatinib-induced growth arrest and apoptosis in chronic myeloid leukemia cells. Cancer. 2006;106:1188–96. doi: 10.1002/cncr.21725. [DOI] [PubMed] [Google Scholar]
  • 171.Dillmann F, Veldwijk MR, Laufs S, Sperandio M, Calandra G, Wenz F, et al. Plerixafor inhibits chemotaxis toward SDF-1 and CXCR4-mediated stroma contact in a dose-dependent manner resulting in increased susceptibility of BCR-ABL+ cell to Imatinib and Nilotinib. Leuk Lymphoma. 2009;50:1676–86. doi: 10.1080/10428190903150847. [DOI] [PubMed] [Google Scholar]
  • 172.Jiang X, Lopez A, Holyoake T, Eaves A, Eaves C. Autocrine production and action of IL-3 and granulocyte colony-stimulating factor in chronic myeloid leukemia. Proc Natl Acad Sci USA. 1999;96:12804–9. doi: 10.1073/pnas.96.22.12804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Zhang X, Ren R. Bcr-Abl efficiently induces a myeloproliferative disease and production of excess interleukin-3 and granulocyte–macrophage colony-stimulating factor in mice: a novel model for chronic myelogenous leukemia. Blood. 1998;92:3829–40. [PubMed] [Google Scholar]
  • 174.Weitz-Schmidt G, Welzenbach K, Brinkmann V, Kamata T, Kallen J, Bruns C, et al. Statins selectively inhibit leukocyte function antigen-1 by binding to a novel regulatory integrin site. Nat Med. 2001;7:687–92. doi: 10.1038/89058. [DOI] [PubMed] [Google Scholar]

RESOURCES