This paper introduces novel therapeutic strategies focusing on a molecular marker relevant to a particular hematologic malignancy. Four different approaches targeting specific molecules in unique pathways will be presented. The common theme will be rational target selection in a strategy that has reached the early phase of human clinical trial in one malignancy, but with a much broader potential applicability to the technology.

In Section I Dr. Richard Klasa presents preclinical data on the use of antisense oligonucleotides directed at the bcl-2 gene message to specifically downregulate Bcl-2 protein expression in non-Hodgkin's lymphomas and render the cells more susceptible to the induction of apoptosis.

In Section II Dr. Alan List reviews the targeting of vascular endothelial growth factor (VEGF) and its receptor in anti-angiogenesis strategies for acute myeloid leukemia (AML) and myelodysplastic syndromes (MDS).

In Section III Dr. Bruce Cheson describes recent progress in inhibiting cell cycle progression by selectively disrupting cyclin D1 with structurally unique compounds such as flavopiridol in mantle cell lymphoma as well as describing a new class of agents that affect proteasome degradation pathways.

Richard J. Klasa, MD*

*

Division of Medical Oncology, British Columbia Cancer Agency, 600 West 10th Avenue, Vancouver BC V52 4E6 Canada

Dr. Klasa is on the international advisory boards of Schering AG and Hoffmann-LaRoche, and is on the speakers' bureau and advisory board for Berlex Canada.

Hematologic malignancies in general and non-Hodgkin's lymphomas (NHLs) in particular are frequently associated with gain of function mutations, many characterized by balanced chromosomal translocations. Genome-wide surveys of gene expression are identifying both known and new transcripts that are overexpressed in different histological subtypes of lymphoma.1 As we develop molecular classifications of these diseases it is assumed that a few key genes will account for the particular survival advantage conferred on malignant lymphocytes as compared to their normal counterparts. These genes and their protein products would provide rational targets for the development of therapeutic strategies to reverse this upregulation associated with the malignant phenotype.

Non-Hodgkin's Lymphoma and Bcl-2

Over the past quarter century cytogenetic analysis has identified a number of reciprocal translocations that frequently occur in histologically identifiable subtypes of NHL.2 The transposition of the bcl-2 gene to the immunoglobulin heavy chain promoter region in the t(14;18) translocation is associated with > 90% of follicular lymphomas (FL) at diagnosis and 10% of diffuse large B-cell lymphomas (DLBCL), making it the most frequent event identified in NHL. Additionally, 50% of DLBCL overexpress the BCL-2 protein through other mechanisms, such as gene duplication, and are associated with a poorer prognosis after anthracycline-based combination chemotherapy.3 BCL-2 is also overexpressed in mantle cell lymphoma (MCL), chronic lymphocytic leukemia (CLL), multiple myeloma (MM) and acute myelogenous leukemia (AML).4 This same widespread pattern of distribution is also seen in a variety of solid tumors including melanoma, small cell lung carcinoma, and colon carcinoma as well as prostate and breast carcinoma, especially once the last two are hormone independent. The obvious conclusion is that BCL-2 overexpression, by whatever means, confers a fundamental advantage to malignant cells and that disruption of this overexpression might have therapeutic potential.

Bcl-2 is an anti-apoptotic member of a large family of genes involved in the regulation of programmed cell death.5,6 Pro-apoptotic (BAX and BCL-Xs) and anti-apoptotic (BCL-2, BCL-XL) molecules reside within the inner mitochondrial membrane and can homo- and heterodimerize upon appropriate stimulus. These interactions control the release of substances such as cytochrome C from the mitochondria into the cytosol through the opening or closing of specific pores in the membrane, with permeability determined by the relative abundance of the different molecules. Cytochrome C is central to the activation of caspases that initiate the apoptotic process. Thus, an overabundance of BCL-2 can prevent or retard activation of the apoptotic machinery and allow survival under conditions that might otherwise be lethal to a cell (Table 1 ).

Antisense Oligonucleotide

Reverse complementary or “antisense” oligonucleotides (ASOs) are short sequences of single stranded deoxyribonucleotides complementary to the coding regions of a gene that are designed to hybridize by Watson-Crick base pairing to messenger-RNA (m-RNA) sequences and thus facilitate their degradation.7,8 Naturally occurring antisense sequences have been identified as regulators of gene expression in a number of systems, supporting their potential for therapeutic development.9,10 The formation of a heteroduplex of m-RNA with the DNA of the ASO engages RNaseH, an enzyme that proceeds to specifically cleave off the m-RNA moiety, destroying the message and putatively leaving the therapeutic ASO molecule able to hybridize to another message sequence.11 This results in a reduction in the target m-RNA pool, which subsequently leads to reduction in the specific protein encoded (Figure 1; see color page 551). The presence of the ASO may also prevent the m-RNA from appropriately docking with the ribosomal machinery that would allow translation into a functional protein. The end result is loss of expression of that protein in the cell.

ASOs of 16-24 bases in length provide target specificity while shorter or longer sequences can result in random hybridization within the transcript repertoire. Selecting the target areas within a messenger RNA must ultimately take into account its tertiary structure, which will determine the accessibility of an area for hybridization. These target areas are defined in oligonucleotide arrays where the entire antisense sequence to an m-RNA is displayed in overlapping segments on a slide. The intensity of hybridization of the labeled message determines the candidate therapeutic ASOs.12 Screening of oligonucleotide libraries has also identified RNA sites that are most accessible to hybridization and correlated these sites with protein downregulation and biological function.13,14 More empirically, the first 6 codons of the open reading frame downstream of the AUG start site have repeatedly been found to be accessible to hybridization and have been chosen for initial development of ASOs against a number of genes.

As organisms have developed a sophisticated system for dealing with rogue strands of DNA both inside and outside the cell, the development of therapeutic molecules required chemical modifications to confer nuclease resistance and a favorable pharmacokinetic profile.15,16 Substitutions in the phosphodiester linkage of the bases in the ASO backbone has yielded a number of molecules now in clinical development with phosphorothioates being the most widely studied first generation molecules (Figure 2 ). The sulfur substitution yields an ASO that is nuclease resistant and capable of entering the cell. It demonstrates good hybridization kinetics and has little in the way of non-sequence-dependent effects or toxicities at concentrations required to downregulate the target message. Additionally, although in tissue culture a delivery system such as cationic lipid is required for efficient intracellular penetration of these highly charged molecules, in vivo ASOs have been shown to be active in free form, possibly due to interaction with blood lipoproteins.17,18 

The correlation of a biologic effect with the specific downregulation of target message and protein in vivo has been a major focus of the development of ASOs. However, ASOs can be very potent immune stimulators, by virtue of unmethylated CpG motifs presented in the context of certain flanking sequences, and therapeutic activity could in part be attributed to nonspecific systemic immune effects rather then to a specific ASO/mRNA interaction.19–,22 Additionally, structures such as guanosine quartets can demonstrate sequence specific but non-antisense biological activity in vitro.23,24 Experimental designs therefore strive, by the use of appropriate control oligonucleotides (sense, missense, one and two base mismatch mis-sense) and various strains of immunodeficient animals, to isolate effects that can be attributed to the specific downregulation of target message and protein (Figure 3 ).

ASOs directed at bcl-2

An 18-mer phosphorothiolated oligonucleotide, G3139, directed against the first six codons of the open reading frame of the bcl-2 gene message has been developed by Genta Pharmaceuticals (Figure 3). Studies of G3139 utilizing the BCL-2 overexpressing lymphoma cell lines DoHH-2 and SU-DHL-4 in vitro have shown downregulation of message and protein expression.25 Tumor xenograft models in SCID mice have demonstrated therapeutic activity that is specific when compared to control animals as well as animals treated with reverse polarity sense, 2-base mismatch mis-sense and non-sense oligonucleotides.25,26 Extensive pharmakokinetic as well as toxicity studies have been performed identifying a dose range with a good therapeutic index.15,16 These findings supported the development of clinical trials using G3139 alone as treatment for BCL-2 overexpressing follicular lymphomas.

Further therapeutic potential is suggested by in vitro experiments confirming that bcl-2 plays a major role in the response of malignant cells to various stresses which produce cellular damage, including chemotherapy (Figure 4 ).4,27,28 Malignant cell lines transfected with a bcl-2 gene with resultant overexpression of the protein product demonstrated increased resistance to various chemotherapeutic agents.29–,32 Additionally, cell lines overexpressing BCL-2 were rendered more sensitive to killing by chemotherapeutic agents when either antisense oligonucleotides directed at the bcl-2 message were introduced into the culture or the cells were transfected with a vector bearing the antisense sequence.33,34 This has been correlated with a demonstration of downregulation of bcl-2 expression. With this as background, we set out to test the in vivo combination of ASOs targeting bcl-2 with a cytotoxic agent commonly used in the treatment of lymphoma.

ASOs to bcl-2and Chemotherapy

Escalating doses of G3139, cyclophosphamide and the combination of both agents were evaluated in severe combined immunodeficiency (SCID) mice bearing a systemic human DoHH2 lymphoma xenograft.35 Experiments confirmed that G3139 was able to downregulate BCL-2 expression in vitro and that treatment with G3139 alone resulted in prolongation of median survival and cure of some animals. (Figure 5 , left panel). This effect was dose and schedule dependent with no long-term survivors seen when a dose of 5mg/kg was given daily for 14 consecutive days as opposed to > 40% when the dose was increased to 12.5 mg/kg on the same daily schedule or either 5 or 12.5 mg/kg were administered for 14 treatments on alternate days (28-day schedule). Similarly, cyclophosphamide treatment alone resulted in no long-term survivors at lower doses but was able to cure animals at high doses. The addition of G3139 to low dose cyclophosphamide resulted in the cure of the majority of animals (Figure 5, right panel). The interaction between the two agents does show dose-response correlations; for the two low doses of cyclophosphamide tested, increasing the dose of G3139 from 2.5 to 5 mg/kg resulted in longer median survivals and overall increase in long-term survivors. A rather striking result was achieved when a completely ineffective dose of cyclophosphamide (15 mg/kg, median survival 36 days and no long-term survivors) was combined with a modestly effective dose of G3139 (2.5 mg/kg, 61 day median survival and 16% long-term survivors) to produce a 72-day median survival and 50% long-term survivors. Mice sacrificed at 90 days showed no histological evidence of any disease in all tissues analyzed, including immunoperoxidase staining for BCL-2, or molecular detection of bcl-2, by PCR. This suggests that chemotherapy at very modest doses could be made much more effective without increasing toxicity when combined with an antisense oligonucleotide. Such an increase in the efficacy of currently available agents could significantly alter the prognosis of a large number of moderately chemotherapy sensitive human tumors, resulting in longer median survivals and increasing the potential for cure.

The model has direct relevance to the clinical situation faced in NHL, where patients typically present with a chemotherapy-sensitive tumor at diagnosis that regresses only to recur within months to years post-treatment. The DoHH2 cell line was derived from a follicular lymphoma carrying a t(14;18), which results in bcl-2 gene overexpression. The aggressive nature of the disease in this model is, however, more suggestive of a transformation to a higher-grade histology, a common event in follicular lymphoma. Indeed, a recent re-exploration of the molecular and cytogenetic features of the cell line, using more sensitive detection techniques, has revealed a second translocation involving the c-myc oncogene with a resultant derivative chromosome 8 carrying t(8;14;18).36 We have recently described this phenomenon of double translocation of both bcl-2 and c-myc in a subset of patients with small non-cleaved cell (Burkitt-like) lymphoma, which represents a very aggressive form of the disease.37 

Clinical Studies

G3139 has been studied as a single agent in a phase 1 trial in heavily pretreated (median of 4 prior regimens) patients with relapsed NHL.38,39 Twenty-one patients with follicular (9), small lymphocytic (8), diffuse large B-cell (3) or mantle cell (1) lymphomas that expressed BCL-2 were treated at 8 dose levels ranging from 4.6 to 195.8 mg/m2/day by continuous subcutaneous infusion for 14 consecutive days. No significant toxicity was seen up to doses of 110 mg/m2/day. One complete and 2 minor responses as well as 9 disease stabilizations were seen in this heavily pretreated group. BCL-2 protein was decreased in 7 out of 16 samples examined, including 2 from accessible tumor sites and 5 samples of peripheral blood or marrow mononuclear cells.

One study combining a chemotherapeutic agent with G3139 has been reported in metastatic melanoma,40 and studies are ongoing in a number of other solid tumors (melanoma, prostate carcinoma) and hematological malignancies (myeloma, chronic lymphocytic leukemia, and acute myeloid leukemia). A phase 1 study at our institution in relapsed follicular lymphomas combining escalating doses of both cyclophosphamide and G3139 has not identified any unexpected toxicity. The last patient enrolled has received cyclophosphamide 750 mg/m2 with 2.3 mg/kg/day of G3139 by continuous intravenous infusion for 14 consecutive days (Table 2 ).

Conclusions

The identification of overexpression or aberrant expression of genes that result in a gain of function, through genome wide surveys of cells and tissues in varying states, will provide unprecedented insight into the biology of hematological malignancies. Specific downregulation of such overexpression with antisense oligonucleotides allows disruption of single gene function at the messenger RNA and protein level and the study of downstream events in the involved molecular pathways both in vitro and in vivo. Genes that are critical to the differential growth and survival advantage enjoyed by malignant cells are being identified and are logical therapeutic targets. The development of ASOs directed at the bcl-2 gene provides a model by which a systemic therapy for a metastatic disease has been taken from the laboratory through preclinical studies to early phase clinical trials, building on knowledge of this particular gene's role in cellular apoptosis. Combining multiple antisense strategies with other therapeutic modalities has the potential to increase the specificity of the treatments available to our patients, thus improving their efficacy and reducing toxicity.

Alan F. List, MD*

*

Arizona Cancer Center, 1515 N Campbell, Room 3945, PO Box 245004, Tucson AZ 85774

The seminal observations that the growth and metastatic potential of solid tumors is dependent upon the formation of new blood vessels triggered an enormous expansion in angiogenic research that has yielded novel therapeutics targeting an array of angiogenic molecules. Investigations of the relevance of angiogenesis in hematologic malignancies are still at an early stage, but accumulating evidence indicates that the angiogenic profile of many hematologic malignancies is distinct from that of solid tumors. As progeny of a common endothelial and hematopoietic stem cell, hematologic malignancies may elaborate and respond to angiogenic factors in a paracrine or autocrine fashion, contributing to tumor cell survival and expansion, adhesion, bone resorption and immune suppression.

Angiogenesis

Blood vessel development is characterized by two distinct biologic processes, vasculogenesis, and angiogenesis. Vasculogenesis, which is largely restricted to embryonic development, involves de novo endothelial cell differentiation from mesodermal precursors as the prerequisite for coordinated blood vessel generation.1 Angiogenesis, the process of new blood vessel formation from preexisting vessels, is responsible for the generation of neovasculature in adult life, and occurs physiologically during wound healing and within female reproductive organs during the menstrual cycle, as well as in pathologic conditions such as proliferative retinopathy, arthritic synovium, and human malignancies.2 Angiogenesis is a multistep process that includes both activation and resolution phases. The activation phase is responsible for the sequential events of basement membrane degradation, endothelial cell proliferation and migration, and capillary lumen formation. The resolution phase is responsible for the maturation and stabilization of the newly formed microvasculature through the recruitment of pericytes, promotion of basement membrane reconstitution, and subsequent extinction of the endothelial cell mitogenic response. A large number of pro-angiogenic molecules and endogenous angiogenesis inhibitors that coordinate the angiogenic response have been delineated (Table 3 ).3 Vascular endothelial growth factor (VEGF), first identified in 1989 and later isolated from the HL-60 myeloid leukemia cell line, is a critical regulator of vascular development that is responsible for activation of endothelial cell proliferation during vasculogenesis and the direction of capillary sprouting during angiogenesis.4 Indeed, gene inactivation studies indicate that VEGF is essential to the neoplastic angiogenic response.5 

VEGF and Receptor Tyrosine Kinases

The VEGF-A molecule is a disulfide-linked homodimer represented by five different isoforms generated by alternate exon splicing of gene message.6 The corresponding VEGF monomers range in size from 121 to 206 amino acids, with smaller molecules (i.e., 121, 145, and 165 amino acids) representing the secreted and diffusable isoforms, whereas the larger proteins (189, 206 amino acids) are sequestered by heparin sulfate residues present on cell surfaces or within the extracellular matrix.7,8 Although all isoforms are biologically active, the VEGF165 isoform predominates and is recognized as a more potent and bioavailable endothelial cell mitogen.9,10 Recent investigations indicate that the VEGF family is composed of five members in addition to the prototype, VEGF-A, including VEGF-B, VEGF-C, VEGF-D, VEGF-E and PIGF (placental growth factor).6 Within the arterial wall, VEGF is produced by smooth muscle cells in response to oxidative stress and other stimuli.11 Autocrine production of VEGF and corresponding receptor upregulation is also demonstrable in endothelial cells in response to hypoxia, nitric oxide, VEGF deprivation and other cellular stresses.12,13 

Trophic response to the VEGF family members is directed by selective interaction with structurally homologous type III receptor tyrosine kinases (RTKs), including VEGFR-1, originally termed fms-like tyrosine kinase (FLT-1),14 VEGFR-2 or kinase insert domain-containing receptor (KDR)/fetal liver kinase-1 (flk-1),15 and the recently characterized VEGFR-3 or FLT-4 receptor (Figure 6 ).16 Each of these receptors contains seven extracellular immunoglobulin homology domains that create the ligand-binding site, a single short transmembrane-spanning sequence, and a cytoplasmic tail that contains the tyrosine kinase domain akin to that of the c-kit and PDGF-receptors.17,18 The external ligand-binding component of VEGFR-3 differs from the other VEGF receptors in that the fifth immunoglobulin domain is cleaved during receptor processing to yield covalent, disulfide-linked subunits.19 In adults, VEGFR-1 and VEGFR-2 expression is limited to the vascular endothelium, monocytes (VEGFR-1), and primitive hematopoietic precursors (VEGFR-2), whereas VEGFR-3 is restricted to lymphatic endothelium.20–,24 The critical role for these receptors and that of VEGF has been demonstrated in gene inactivation studies. VEGF-A and the VEGFR-2 receptor are essential to embryonic vasculogenesis and definitive hematopoiesis.25–,27 In vitro investigations indicate that VEGFR-2 signaling is responsible for proliferation and differentiation of endothelial cells in response to VEGF-A stimulation, whereas VEGFR-1 has been implicated as a decoy receptor and potential activator of endothelial permeability.28 Ligand binding to VEGFR-3 is restricted to VEGF-C and VEGF-D, which together direct lymphangiogenesis.17,29,30 The actions of the remaining VEGF family members are less well-defined; however, receptor recognition for VEGF-B appears restricted to VEGFR-1 where it is believed to regulate extracellular matrix (ECM) degradation and endothelial cell migration via protease elaboration.31 VEGF-E is an alternate ligand for VEGFR-2 and has biologic activities similar to VEGF-A.32 Neuropilin-1 (NRP-1), a cell surface glycoprotein, acts as a co-receptor for VEGF-A165 to enhance its binding to VEGFR-2 on endothelial cells. Unlike the VEGF receptors, NRP-1 lacks an intracellular domain and therefore is not a direct mediator of cytokine signaling.33 

VEGF-A exerts its biologic effects via interaction with either VEGFR-1 or VEGFR-2. VEGF receptors dimerize in response to ligand engagement, activating the tyrosine kinase located on the intracellular cytoplasmic tail. The resultant auto-phosphorylation of these receptors on specific tyrosine residues creates docking sites for cytoplasmic signaling molecules.34 Activation of VEGF receptors in endothelial cells triggers recruitment of adaptor and signaling proteins that contain Src homology domain-2 (Sh-2) including Shc and Sck.35,36 Recruitment and phosphorylation of these adaptor molecules permit their binding to the p85 subunit of phosphatidyl inositol-3 (PI-3) kinase and to phospholipase-Cg, activating the PI-3 kinase and ERK (stress activated protein kinase-1, 2) kinase signaling cascades.37–,39 PI-3 kinase phosphorylates Akt (protein kinase-B), a serine/threonine kinase involved in antiapoptotic signaling.40 Akt activation permits its translocation to the plasma membrane, where it phosphorylates and inactivates a variety of pro-apoptotic molecules, including BAD, caspase-9, mitochondrial Raf, and the forkhead transcription factor-1 (FKHR-1).41 Similarly, through PI-3 kinase activation, VEGF-A activates focal adhesion kinase in endothelial cells, a step that is critical to the recruitment and activation of cell adhesion molecules containing β1, β2, and β3 integrins.42,43 Indeed VEGFR-2 interaction with its ligand triggers receptor association with PI-3 kinase, vascular endothelial (VE)-cadherin and αVβ3, components that are essential to the creation of an active multimeric signaling complex. Thus, VEGFR-2 activation assures proper juxtaposition of the receptor with cytoskeletal proteins.44,45 

VEGF Regulation

Transcriptional regulation of VEGF is influenced by a number of signals that converge upon the oxygen sensitive transcription factor, hypoxia-inducing-factor (HIF)-1a. HIF is a heterodimeric protein composed of α and β subunits.46 The β subunit is expressed constitutively, whereas the a subunit is a member of a family of DNA-binding proteins that contain trans-activation domains within the carboxy-terminus. Under normal oxygen conditions, HIF-1a is degraded rapidly by the ubiquitin-proteasome pathway.47 Hypoxia stabilizes the HIF-1a protein by inhibiting its proteasome degradation. The von Hippel-Lindau (VHL) tumor suppressor gene is one component of the ubiquitin ligase complex that ubiquinates and promotes the degradation of HIF-1a under normoxic conditions. Inactivation of VHL, as occurs in the von Hippel-Lindau syndrome, stabilizes HIF-1a and sustains tumor production of VEGF.48 Inactivation of a negative regulator of HIF-1a such as VHL is only one of many signals which converge upon HIF-1a to activate VEGF expression. Alternate signals include sustained RAS activation, β-integrin activation, stress-induced ERK signaling, and inactivation of p53.49–,52 In addition to its regulatory effects on VEGF, HIF-1a is a transcriptional activator of other genes including erythropoietin, inducible nitric oxide synthetase (iNOS), insulin growth factor (IGF-2), glycolytic enzymes, and tumor growth factor-β (TGF-β).53 

Hemangioblast and Hematopoiesis

During embryogenesis, pluripotential hematopoietic stem cells originate from cell clusters in the ventral floor of the pre-umbilical dorsal aorta.54 These common hematopoietic/endothelial (HE) precursors share an antigenic phenotype characterized by expression of the progenitor cell antigen, CD34, and VEGFR-2.27,55 Murine gene knockout studies indicate that VEGFR-2 expression is required for vasculogenesis and for establishment of definitive hematopoiesis. Although HE progenitors can be generated in vitro from VEGFR-2 deficient embryonal stem cells, VEGFR-2 signaling is essential for progenitor migration and endothelial commitment. Post-natally, hemangioblasts, or HE progenitors, persist and contribute to long-term hematopoiesis and to the generation of circulating endothelial cells. Confirmation of leukemia specific gene expression (i.e., bcr/abl in chronic myeloid leukemia) and the identification of donor-specific DNA alleles (i.e., in allograft recipients) within vascular endothelium provides definitive proof that bone marrow-derived hemangioblast progeny contribute to the maintenance of endothelium in adults.56–,58 These primitive hemangioblast progenitors represent less than 2% of circulating CD34+ cells and display a surface phenotype characterized by CD34+, AC133+, VEGFR-2+, platelet endothelial cell adhesion molecule (PECAM), and the stromal derived factor-1 (SDF-1) receptor, CXCR-4.59 Under appropriate culture conditions, hematopoietic commitment of these precursors is demonstrable by acquisition of the leukocyte common antigen, CD45, and β1 integrins, whereas angioblasts display VE-cadherin.

VEGF receptorexpression extends beyond the primitive progenitor compartment during adult hematopoiesis. VEGFR-1 is expressed by monocytes and committed CD34+ progenitors, and is responsible for activation of migration and maturation skewing, respectively.22,23,60 With the exception of primitive hematopoietic progenitors, VEGFR-2 expression appears restricted to megakaryocytes.20 Within the adult bone marrow, cellular expression of angiogenic molecules outside of stromal elements is restricted to erythroblastic islands. Both VEGF-A and PIGF are secreted by erythroblasts within erythroid islands,61 suggesting that these cytokines may serve to recruit nurse macrophages and promote β-integrin activation in a paracrine fashion.

Laboratory and animal studies employing recombinant human VEGF (rhu-VEGF) have shown that the cytokine exerts both growth-promoting and growth-suppressive effects on hematopoietic progenitors that are lineage- and maturation-dependent. Rhu-VEGF impairs maturation of dendritic cells from CD34+ cells and promotes expansion of immature granulocyte-macrophage progenitors in animal models60 while inhibiting the formation of primitive erythroid and multipotent progenitors.62 Similarly, VEGF promotes osteoclast differentiation and bone remodeling and, as a consequence, contributes to bone resorption in animal models.63,64 In receptor-competent hematopoietic cells VEGF promotes resistance to radiation via induction of the antiapoptotic BCL-2 homologue, MCL-1, and activation of the Krüppel-type zinc finger transcription factor, ZK7.65–,67 

Hematologic Malignancies

Tumor cells elaborate angiogenic factors as an essential feature of the malignant phenotype providing paracrine signaling for tumor growth. Hematopoietic malignancies, as progeny of receptor-competent HE progenitors, are distinguished from other tumors by the potential for autocrine growth stimulation and thereby offer the prospect of clinical benefit with antiangiogenic therapy. Thus, in AML and MDS, for example, myeloblasts and leukemic monocytes secrete VEGF and display VEGF receptors.68,69 In leukemia cell lines, rhu-VEGF directly promotes colony-forming capacity and β-1 integrin activation via the PI-3 kinase/Akt signal pathway, analogous to its effects in endothelial cells.69–,71 Similarly, neutralizing VEGF suppresses TNFα generation in MDS bone marrow stroma and promotes the recovery of primitive progenitors.69,Table 4  summarizes the angiogenic profile of a number of hematologic malignancies and the relationship to clinical and biological features. Although basic fibroblast growth factor (bFGF) production is demonstrable in the majority of hematologic malignancies, expression on the malignant cells of functional bFGF receptors appears limited to chronic myeloid leukemia (CML)72,73 and MM.74 The relevance of tumor-derived angiogenic molecules to disease pathobiology is demonstrated by their linkage to prognosis and/or disease behavior. In NHL, for example, VEGF expression is common; however, co-expression of VEGF receptor(s) appears limited to intermediate- or high-grade lymphomas and adversely impacts overall survival.75–,77 Similarly, in AML, elevated cellular VEGF content is an independent prognostic variable associated with resistance to induction chemotherapy and inferior overall and disease-free survival. Leukemia-derived VEGF appears essential for bone marrow engraftment of AML cells in SCID mouse models.78,79 Indeed, rhu-VEGF promotes clonogenic growth of receptor competent AML cell lines and serves as an autocrine signal implicated in the central coalescence of immature myeloid precursors (i.e., abnormal localized immature precursors [ALIP]) in advanced MDS.69–,71 The encouraging results of treatment with the antiangiogenic agent thalidomide in patients with refractory myeloma suggests that angiogenesis represents an appropriate therapeutic target in the hematologic malignancies.80 

Antiangiogenic Agents

The explosion in our understanding of neoplastic angiogenesis in the past decade has generated an ever-expanding catalogue of potential targets and therapeutics in clinical development. The angiogenesis antagonists can be classified into five distinct types, including the protease inhibitors that impact extracellular matrix remodeling, inhibitors of activated endothelial cell proliferation, inhibitors of survival signaling via vascular adhesion molecules, and agents that interfere with the generation of angiogenic molecules or receptor activity (Table 5 ). Among protease inhibitors, the matrix metalloprotease (MMP) inhibitors have undergone extensive testing in patients with solid tumors and are now being tested in hematologic malignancies. Most MMPs are soluble proenzymes that upon activation initiate proteolysis of basement membrane constituents such as collagen, laminin and proteoglycans.3 MMPs are classified according to their domain structure and substrate specificity. A leading MMP candidate for antiangiogenic therapy in hematologic malignancies is the synthetic gelatinase inhibitor, AG3340 (Prinomastat™),81 which, because of its selectivity, has limited clinical toxicity. The biological targets of gelatinases extend beyond basement membrane dissolution, and include the disruption of integrin interactions and the generation of soluble TNFα and fas ligand via liberation from membrane bound isoforms.82–,84 A common structural feature of MMP inhibitors is the presence of a metal binding group, in this case hydroxamate, that chelates the zinc ion present within the catalytic domain of the enzyme. AG3340 is currently completing phase II investigations in patients with MDS.

The inhibitors of endothelial cell activation have undergone the most extensive testing in hematologic malignancies. These agents, such as endostatin and synthetic pharmacophores, interfere with angiogenic factor-induced endothelial cell migration and proliferation and may possess additional biological effects. Thalidomide, a synthetic phamacophore, is a sedative and potent teratogen that has both antiinflammatory and antiangiogenic properties. Its biological effects include inhibition of bFGF- and VEGF-induced angiogenesis, suppression of TNFα generation by potentiating mRNA degradation, and the modulation of cell adhesion molecule activation.85,86 Thalidomide and its analogs augment natural killer cell cytotoxicity and exert direct cytostatic effects in myeloma cell lines unrelated to their antiangiogenic properties.87,88 Indeed, in a study involving 84 patients with refractory myeloma, durable antitumor responses were reported in up to 30% of patients.80 In MDS, approximately 30% of patients may achieve red blood cell transfusion independence following treatment with thalidomide.89 Although phase II and phase III testing of thalidomide in both myeloma and MDS are underway, its use is limited by neurologic toxicity. The thalidomide analog, CC-5013 (Revimid™), displays greater in vitro potency, costimulates the Th 1 immune response, and appears to be devoid of neurotoxicity. CC-5013 recently entered clinical trials in myeloma and MDS. Other angiogenic inhibitors that are completing testing in hematologic malignancies include arsenic trioxide and the farnesyl transferase inhibitors (FTI), SCH66336 (Schering Pharmaceuticals) and R115777 (Janssen). The FTIs inhibit ras protooncogene activation, an upstream transcriptional regulator of VEGF essential for endothelial cell proliferation. Moreover, gene inactivation studies indicate that VEGF is necessary for ras-induced cell transformation, suggesting that tumorigenicity of the ras oncogene is VEGF-dependent.90 Preliminary results of phase I studies of R115777 have shown single agent activity in AML.91 

Agents that directly target angiogenic factors or their receptors offer the prospect for greater activity in receptor-competent hematologic malignancies by interrupting autocrine receptor signaling. The humanized monoclonal anti-VEGF antibody, Bevacizumab (Genentech) produces sustained neutralization of circulating VEGF and is now in phase II testing in MDS, lymphoma, AML, and solid tumors. The receptor tyrosine kinase inhibitors (RTKI) represent a particularly exciting class of synthetic, small molecule inhibitors of angiogenic receptor signaling. The first receptor antagonist to enter clinical testing in hematologic malignancies is SU5416 (Sugen), which impairs ligand-induced autophosphorylation of the VEGFR-1 and VEGFR-2 receptors and c-Kit. SU5416 inhibits VEGF-induced clonogenic response in leukemia cell lines and promotes apoptosis in myeloblasts from AML patients.70,92 Although it shows promising clinical activity in initial phase II testing,93 the utility of SU5416 is limited by its solubility and necessity for intravenous administration. Other RTKIs are entering phase II testing in AML and other receptor-competent hematologic malignancies, including SU11248 (Sugen), PTK787/ZK222584 (Novartis), and AG13736 (Agouron). These orally bioavailable inhibitors demonstrate broad receptor specificity, inhibiting the VEGF receptors, c-kit, PDGF, bFGF, and c-FMS. Although these agents may have inherent antineoplastic effects in receptor-competent malignancies, they may be more effective when combined with traditional DNA-interactive antineoplastics. As a group, this class of agents presents enormous potential to impact survival signaling from cytokine receptors, selected constitutively active TKs, integrin activation state, and angiogenic response. The angiogenic inhibitors are one of the most exciting classes of therapeutics to enter clinical testing in hematologic malignancies. Their clinical potential, however, will not be realized for several years to come.

Bruce D. Cheson, MD*

*

NCI-CTEP, Executive Plaza, No.-741, Bethesda MD 20892

During the past few decades, results of clinical trials in the NHLs have failed to demonstrate a major impact on patient survival. Research strategies in the indolent NHL largely focused on comparisons of combinations and permutations of regimens including alkylating agents or nucleosides with or without anthracyclines or interferon. After decades of comparing various combination regimens in the aggressive NHL, CHOP (cyclophosphamide, doxorubicin, vincristine and prednisone) remained the standard. The recent availability of clinically active monoclonal antibodies has the potential to alter our approach to these patients. Nevertheless, there is considerable room for improvement in our current chemotherapy. The recent recognition of distinct molecular targets provides an opportunity to identify new drugs with unique mechanisms of action rather than relying on empiric combinations of nonspecific chemotherapy agents (Table 6 ). This section will present information on two such novel compounds as examples for the future development of chemotherapy agents.

Flavopiridol

Progression through the cell cycle is a tightly regulated process controlled by the phosphorylation and proteolysis of regulatory proteins including cyclins, cyclin-dependent kinases (cdk), and their inhibitors (Figure 7; see color page 551). Positive regulatory function of the cell cycle is afforded by p53, the retinoblastoma protein (Rb), and the p16 (INK4A) family of cyclin-dependent kinase inhibitors. Protein kinases play a central role in the function, proliferation, growth, transformation, and death of cells. The process of malignant transformation is associated with a progressive loss of cdk inhibitors and overexpression of cyclins, leading to a growth and proliferation advantage for the malignant cell. An important regulatory checkpoint occurs late in the G1 phase of the cell cycle where cyclin-dependent kinases are activated by cyclins to drive cells to the S phase.

The D-type cyclins, which associate with cdk4 and cdk6 to traverse G1 and with Rb and p53 to regulate G1/S progression, are often mutated in malignancy. Thus, blocking cell cycle progression with cdk inhibitors may lead to growth arrest and apoptosis.

Mantle cell lymphoma (MCL) is a distinct clinical, genetic and molecular entity. Clinically, it exhibits the worst features of the aggressive and indolent NHL and can behave in an aggressive manner. Like the indolent NHL, it is not a curable disease and has a median survival of only 2.5-3 years.1 At the molecular level, MCL is characterized by overexpression of the G1 cyclin, cyclin D1, generally in association with the chromosomal translocation t(11;14)(q13;q32).2 This translocation juxtaposes the enhancer element of the immunoglobulin heavy chain region on chromosome 14 to the bcl-1 or PRAD-1 protooncogene encoding the cyclin D1 protein on chromosome 11. As a consequence of the overexpression of cyclin D1, there is excess progression of cells from G1 through the S phase of the cell cycle. In addition, p53 mutations in patients with MCL are associated with a poor outcome and also with blastic morphology.3 The overexpression of cyclin D1 makes MCL a suitable tumor against which to study agents that target cyclin D1 (Figure 8; see color page 552).

Flavopiridol is a semisynthetic nonchlorinated flavone derivative that was first isolated from the plant alkaloid rohitukine from the leaves and stems of Amoora rohituka, and later from Dysoxylum binectariferum. Both plants are used in India as herbal medicines. It is the first active cyclin dependent kinase inhibitor to enter clinical trials in the US. Initial studies with a human breast cancer line showed that flavopiridol could inhibit in vitro growth with arrest of cells at G1 or G2 phases of the cell cycle. In vitro activity against cycling as well as noncycling cells has now also been demonstrated in a number of tumor cell lines. Flavopiridol inhibits a variety of protein kinases including cyclin D1, which is implicated in the pathogenesis of MCL, as well as cdk1, cdk23, and cdk4 through targeting of the ATP-binding site (Figure 9; see color page 552). It is active against all cdks at concentrations of 100-300 nM, which may be a major factor in its in vitro antitumor activity. In addition, it induces growth arrest, cytotoxic cell death and apoptotic changes in a variety of tumor types, including leukemias and lymphomas. The drug exhibits comparable activity against resting and proliferating cells.

Preclinical studies

Based on in vitro observations that flavopiridol can kill noncycling cells, studies were conducted to test the hypothesis that it might be advantageous to combine the drug with agents that inhibit cell cycle progression. Flavopiridol demonstrated sequence specific synergy in vitro with cell cycle active agents including fludarabine and cytarabine, as well as paclitaxel, topotecan, gemcitabine, irinotecan, doxorubicin, etoposide, cisplatin, and 5-FU.4 The greatest activity is observed when chemotherapy agents are administered prior to flavopiridol. The sequence specificity has been thought to reflect the arrest of cells in G1 and G2 during and 24 hours following flavopiridol administration. Although schedules that approximate continuous infusion have been active in a number of cell lines, intermittent bolus schedules have shown activity against lymphoma and solid tumor cell.

Flavopiridol induces apoptosis of B cell chronic lymphocytic leukemia (B-CLL) and lymphoma cells (Table 7 )58. Flavopiridol is cytotoxic to CLL cells at concentrations of 0.1-0.3 μM with only a 24-hour drug exposure, with no additional benefit from more prolonged exposure. Whether apoptosis occurs in association with a change in expression of bcl-2 or bax is controversial.57 Whereas König et al5 noted an association with bcl-2 expression, Byrd et al found apoptosis without bcl-2 modulation, even in the setting of decreased p53 expression, possibly related to cleavage of caspase 3.6,7 The mean attainable concentration in vivo was 2.4 times the concentration that resulted in apoptosis of 50% of human CLL cells in vitro.7 Unfortunately there was no apparent selectivity between the malignant lymphocytes and normal mononuclear cells. In the studies by Achenbach et al9 induction of apoptosis by flavopiridol appeared to be independent of bcl-2. In other studies, flavopiridol induced concentration dependent apoptosis of CLL cells with decreases in antiapoptotic proteins Mcl-1, X-linked inhibitor of apoptosis (XIAP), and BAG-1 in nearly all cases.10 Whether the disparities among these various studies regarding the role of bcl-2 reflects differences in doses or other factors is not clear. Flavopiridol may also have antiangiogenic properties.11 Arguello et al12 studied the toxicity and activity of flavopiridol in mice xenografts with either a leukemia or lymphoma cell line. There was little effect on most normal tissues; however, the spleens were smaller with reduced white pulp and absent follicle centers. The size of the thymus was decreased and markedly depleted of lymphocytes. Peripheral lymph nodes were also reduced in size and without follicle centers. Following administration of flavopiridol by bolus i.v. or intraperitoneal administration, 11 of 12 human HL-60 xenografts underwent complete regressions, and the animals remained disease-free for several months. Six of 8 animals with a lymphoma cell line (SUDHL-4) underwent either a major (n = 2) or complete (n = 4) regression, and the animals remained disease free for more than 60 days. This effect was shown to be both dose and schedule dependent.

Clinical trials

Flavopiridol is the first cdk inhibitor to be tested in clinical trials. The first two phase I trials with this agent used a 72-hour infusion schedule given every 2 weeks. This schedule resulted in blood levels that were inhibitory in vitro. The recommended phase II dose using this schedule was 78 mg/m2/d for 3 days. The dose limiting toxicity in the initial studies was secretory diarrhea. When antidiarrheal prophylaxis with cholestyramine and loperamide was used, hypotension became dose-limiting, along with a proinflammatory syndrome characterized by fever, fatigue, tumor pain, and alterations in acute phase reactants. Responses were observed in one patient each with renal cell carcinoma and colon cancer, and a minor response was reported in a patient with NHL. A number of patients also experience prolonged stable disease, suggesting the possibility of a cytostatic effect. Other schedules of administration, including a 1-hour daily bolus, have also been tested. The maximum tolerated dose (MTD) on the 1-, 3-, and 5-day schedules were 37.5, 50, and 62.5 mg/m2/d, respectively. Other trials are exploring schedules involving 1- and 3-hour infusions or a bolus followed by a continuous infusion in CLL and other hematologic diseases and solid tumors.

Because of its effect on cyclin D1, flavopiridol was considered early on as a potential agent for the treatment of patients with MCL. Two such trials have been conducted: the Dana-Farber Institute, using a 72-hour infusion, and in the National Cancer Institute (NCI)-Canada trial. In the latter flavopiridol was administered at a dose of 50 mg/m2 over 1 hour daily for 3 days every 3 weeks. Patients received from one to more than 6 cycles. Severe toxicities included diarrhea (n = 5, including 3 patients who did not receive appropriate prophylaxis) and fatigue (78% overall, < 5% grade IV). The diarrhea correlates inversely with systemic glucuronidation of flavopiridol.13 Hematologic toxicity has been mostly mild to moderate with a few cases of severe neutropenia. A number of patients also experienced thrombosis that was thought to be drug related. In the Dana-Farber trial, activity was minimal (M. Shipp, personal communication). In the NCI-Canada trial there were 3 partial responses and 11 patients with stable disease (E. Eisenhauer, J. Connors, personal communication). These disparate observations suggest a possible schedule-dependent difference in efficacy. Similar observations have been made in CLL trials (J. Byrd, personal communication). Thus, in phase II trials using a 1-hour schedule, activity has been observed both in MCL and CLL where no activity was noted using the longer infusion schedules.

Clearly, this agent has limited potential as monotherapy in the lymphoid malignancies studied thus far; however, it has potential in combination with a variety of chemotherapy agents. Multiagent strategies are being planned, primarily with the shorter infusion schedule in a collaborative effort between Aventis Pharmaceutical and the NCI.

Proteasome Inhibitors

The proteasome is a large, multicentric protease complex with a pivotal role in cellular protein regulation. It is composed of a two copies of a 19S regulatory protein that recognizes substrates adorned with ubiquitin chains. The two copies form the 20S core particle that contains the catalytic protease functions. The proteasome degrades proteins that have been conjugated to ubiquitin (the ubiquitin-proteasome pathway)14–,17 (Figure 10 ). The ubiquitin-proteasome pathway plays a critical role in the degradation of intracellular proteins involved in cell cycle control, transcription factor activation, apoptosis, cell trafficking, and tumor growth through an ATP-dependent process. Some of the proteins that undergo the degradation include the cyclins and cdk inhibitors. Many tumor cells depend on rapid cell cycling, which requires expression and degradation of numerous regulatory proteins. Cells accumulate in the G2-M phase of the cell cycle with a decrease of cells in G1. Unfortunately, the indolent lymphoid malignancies are not rapidly cycling and tend to be in G0. Nevertheless, proteasome inhibitors also induce apoptosis despite cellular accumulation of the cdk inhibitors p21 and p27 and irrespective of p53 status.

The proteasome is also required for activation of the nuclear transcription factor NF-κB, which plays a role in maintaining cell viability through the transcription of inhibitors of apoptosis in response to environmental stress or cytotoxic agents.18 

Based on these observations, targeting the proteasome has become a novel new approach to cancer therapy. Several naturally occurring and synthetic proteasome inhibitors have been identified including the boronic acid peptides such as PS-341; the natural product lactacystin, a streptomyces metabolite; MG0132; and a synthetic peptide aldehyde. PS-341 is the compound being most widely studied in the clinic. This dipeptidyl boronic acid is a specific and selective inhibitor of the 26S proteasome.17 PS-341 may also lead to an induction of apoptosis (Figure 11 ). Moreover, enhanced tumor efficacy in combination with other chemotherapy agents and radiation has been suggested. In vitro studies have demonstrated marked synergy between PS-341 and chemotherapy drugs such as the topoisomerase I inhibitor irinotecan (CPT-11). CPT-11 can increase NF-κB leading to greater transcription of factors that protect cells from apoptosis.19 PS-341 may block activation of NF-κB, thus potentiating the activity of the chemotherapy drug. Since NF-kB can induce drug resistance, this agent may make cells more chemosensitive.20 

PS-341 has demonstrated activity against many tumor types in vitro and is a poor substrate for the multi-drug resistance transporter. It is also active even in the setting of overexpression of bcl-2 and has shown a lack of acquired drug resistance. Using a prostate cancer cell line, Adams et al17 found that exposure of the cells to PS-341 results in accumulation in the G-M phase of the cell cycle. Intravenous treatment of animals with this agent led to a 60% decrease in tumor size. Another proteasome inhibitor, PSI, also induces apoptosis of myeloid leukemia cells to a greater degree than it affects normal hematopoietic progenitors.21 PSI has also shown activity against a murine model of Burkitt's lymphoma.22 PS-341 has also been shown to inhibit growth, induces apoptosis conferred by IL-6, and overcomes drug resistance in human myeloma cells.23 These effects have been shown to be additive with dexamethasone, a highly active agent in MM. Low levels of the cdk inhibitor p27 in MCL may result from increased proteasome-mediated degradation,24 making this tumor also a possible target for proteasome inhibitors.

Proteasome inhibitors may have a potential role in the therapy of patients with CLL. First, the ubiquitin proteasome-dependent protein processing may be altered in CLL cells.25 In addition, proteasome inhibition has been shown to induce apoptosis of CLL lymphocytes at concentrations that do not have that effect on normal cells. Masdehors et al25 studied cells from 50 patients with previously untreated CLL. They found that CLL lymphocytes were significantly more sensitive to the induction of apoptosis by the proteasome inhibitor lactacystin than were normal lymphocytes. Furthermore, lactacystin sensitized chemoresistant and radioresistant CLL cells to apoptosis induced by TNFα.26,27 Proteasome inhibition also induces DNA fragmentation and apoptosis of CLL lymphocytes, even those resistant to glucocorticoids.28,29 The changes were also associated with inhibition of NF-κB.

Toxicology studies in rodents and primates identified adverse effects including gastrointestinal toxicity, anorexia, emesis and diarrhea. Hematologic toxicity was not noted. Of interest was that there was lymphocyte depletion of the spleen and thymus.

PS-341 is currently in clinical trials for a variety of solid tumors, leukemias and lymphomas. A number of doses and schedules of administration are under investigation. In a phase I trial in patients with hematologic malignancies30 PS-341 was administered as a twice weekly bolus injection for 4 consecutive weeks followed by a 2-week rest period. Responses were assessed after each 6-week cycle. As of the preliminary publication, 3 patients were treated at each of the first three dose levels, most of whom were heavily pretreated. Included were 1 patient with refractory anemia with excess blasts (RAEB), 2 with Hodgkin's disease, 3 with NHL, and 3 with MM. Although the MTD had not yet been reached, two of the MM patients experienced responses with one major response with > 50% decrease in IgG and normal numbers of bone marrow plasma cells. Other schedules under investigation include twice weekly every other week, and weekly for 4 weeks.

The maximum pharmacodynamic effect of PS-341 on 20S proteasome activity in the peripheral blood is observed at 1-hour post dose, measured by an inhibition assay.31 The compound has been well tolerated at the levels of 20S inhibition of > 60% that are considered essential for its antitumor efficacy. Toxicities encountered to date including dose-limiting neuropathy as well as fatigue, gastrointestinal symptoms, and thrombocytopenia. An active development strategy is planned by Millennium Pharmaceuticals in collaboration with the NCI.

We are in an exciting period in the treatment of the NHL. The therapeutic paradigm for the treatment of indolent NHL is rapidly moving away from alkylating agent-based regimens to those that include purine analogs and, more recently, monoclonal antibodies. However, no single chemotherapeutic or biologic agent alone is likely to significantly prolong the survival of these patients. Future trials should develop rational combinations of chemotherapy and biological agents based on a better understanding of mechanisms of drug action and resistance, interactions among agents, and tumor biology. To expedite the identification of effective strategies, patients should be entered onto clinical trials addressing important research questions. Identification of the most promising of these drugs and regimens will be facilitated by recently published guidelines for assessment of treatment response.32 The large number of new, unique therapeutic agents available for clinical trials should provide optimism for improving the cure rate for patients with NHL.

Table 1.

Properties of BCL-2.

•Oncogenic protein 
•Anti-apoptotic 
•Mitochondrial, endoplasmic reticulum, nuclear membrane localizations 
•Homotypic and heterotypic dimerization within family 
•Membrane channel/pore function 
•Cytochrome C release from mitochondria via BCL-2 family channels regulates cell fate under stress 
•Oncogenic protein 
•Anti-apoptotic 
•Mitochondrial, endoplasmic reticulum, nuclear membrane localizations 
•Homotypic and heterotypic dimerization within family 
•Membrane channel/pore function 
•Cytochrome C release from mitochondria via BCL-2 family channels regulates cell fate under stress 
Table 2.

Phase I study of G3139 and cyclophosphamide in relapsed follicular lymphoma.

LevelG3139 mg/kg/day CIVI (day 1 to 14)Cyclophosphamide mg/m2/IV (day 8)
0.6 250 done 
1.2 250 done 
2.3 250 done 
2.3 500 done 
2.3 750 ongoing 
3.1 1000 
5.0 1000 
LevelG3139 mg/kg/day CIVI (day 1 to 14)Cyclophosphamide mg/m2/IV (day 8)
0.6 250 done 
1.2 250 done 
2.3 250 done 
2.3 500 done 
2.3 750 ongoing 
3.1 1000 
5.0 1000 
Table 3.

Endogenous stimulators and inhibitors of angiogenesis.

Angiogenic MoleculesNative Inhibitors
Abbreviations: VEGF, vascular endothelial growth factor; PIGF, placental growth factor; iNOS, inducible nitric oxide synthetase; MMP, matrix metalloproteases 
Growth factors VEGF (A-D), PIGF Interferon (IFN)-α,γ 
 Angiogenin  
 Angiotropin  
 Epidermal growth factor (EGF)  
 Fibroblast growth factor (acidic and basic; FGF) Prolactin (16 Kd fragment), IFN-α,γ 
 Hepatocyte growth factor/scatter factor (HGF, SF)  
 Platelet-derived growth factor (PDGF)  
 Tumor necrosis factor-α (TNFα)  
 Interleukin-8 (IL-8) Insulin-like growth factor-1 (IGF-1)  
Proteases and Protease Inhibitors Cathepsin Tissue inhibitor of metalloprotease (TIMP-1, TIMP-2) 
 Gelatinase-A, -B (MMP2,9)  
 Stromelysin  
 Urokinase-type plasminogen activator (uPA) Plasminogen activator- inhibitor-1 (PAI-1) 
Trace Elements Copper Zinc 
Oncogenes c-myc p53 Rb 
 ras  
 c-src  
 v-raf  
 c-jun  
Signal Transduction Enzymes Thymidine phosphorylase  
 RAS - arnesyl transferase  
 Geranylgeranyltransferase  
Cytokines Interleukin-1 Interleukin-10 
 Interleukin-6 Interleukin-12 
 Interleukin-8  
Endogenous αvβ3 integrin Angionpoietin-2 
 Angionpoietin-1(Ang-1) Angiotensin 
 Antionstatin II (AT1 receptor) Angiotensin II (AT2 receptor) 
 Endothelin (ETB receptor) Caveolin-1, caveolin-2 
 Erythropoietin Endostatin 
 Hypoxia Interferon-α 
 iNOS Isoflavones 
 Platelet-activating factor (PAF) Platelet factor-4 
 Prostaglandin E, COX-2  
 Thrombopoietin Thrombospondin 
  Troponin-1 
Angiogenic MoleculesNative Inhibitors
Abbreviations: VEGF, vascular endothelial growth factor; PIGF, placental growth factor; iNOS, inducible nitric oxide synthetase; MMP, matrix metalloproteases 
Growth factors VEGF (A-D), PIGF Interferon (IFN)-α,γ 
 Angiogenin  
 Angiotropin  
 Epidermal growth factor (EGF)  
 Fibroblast growth factor (acidic and basic; FGF) Prolactin (16 Kd fragment), IFN-α,γ 
 Hepatocyte growth factor/scatter factor (HGF, SF)  
 Platelet-derived growth factor (PDGF)  
 Tumor necrosis factor-α (TNFα)  
 Interleukin-8 (IL-8) Insulin-like growth factor-1 (IGF-1)  
Proteases and Protease Inhibitors Cathepsin Tissue inhibitor of metalloprotease (TIMP-1, TIMP-2) 
 Gelatinase-A, -B (MMP2,9)  
 Stromelysin  
 Urokinase-type plasminogen activator (uPA) Plasminogen activator- inhibitor-1 (PAI-1) 
Trace Elements Copper Zinc 
Oncogenes c-myc p53 Rb 
 ras  
 c-src  
 v-raf  
 c-jun  
Signal Transduction Enzymes Thymidine phosphorylase  
 RAS - arnesyl transferase  
 Geranylgeranyltransferase  
Cytokines Interleukin-1 Interleukin-10 
 Interleukin-6 Interleukin-12 
 Interleukin-8  
Endogenous αvβ3 integrin Angionpoietin-2 
 Angionpoietin-1(Ang-1) Angiotensin 
 Antionstatin II (AT1 receptor) Angiotensin II (AT2 receptor) 
 Endothelin (ETB receptor) Caveolin-1, caveolin-2 
 Erythropoietin Endostatin 
 Hypoxia Interferon-α 
 iNOS Isoflavones 
 Platelet-activating factor (PAF) Platelet factor-4 
 Prostaglandin E, COX-2  
 Thrombopoietin Thrombospondin 
  Troponin-1 
Table 4.

Angiogenic profile of hematologic malignancies.

Angiogenic Receptors
MalignancyPro-Angiogenic MoleculesMVDVEGFR-1VEGFR-2FGF-RClinical/Biological CorrelationReference
Abbreviations: CR, complete remission; s, serum or plasma; u, urinary; c, cellular; CRP, C-reactive protein; B2M, beta2-microglobulin; LN, lymph node; MMM, myelofibrosis with myeloid metaplasia; Unk, unknown; IL-6, interleukin-6; ALIP, abnormal localized immature precursors; AML, acute myelogenous leukemia; MDS, myelodysplastic syndrome; ALL, acute lymphoblastic leukemia; NHL, non-Hodgkin's lymphoma; CLL, chronic lymphocytic leukemia; CML, chronic myeloid leukemia; MVD, microvessel density; OS, overall survival; DFS, disease-free survival; BM, bone marrow; VEGF, vascular endothelial growth factor; FGF, fibroblast growth factor; bFGF, basic FGF; SDF-1, stromal-derived factor 1 
* Denotes independent prognostic factor. 
AML s,cVEGF* ↑BM ++ +/- Unk Leukocytosis; ↓CR*, OS*, DFS* (VEGF) 68,69,71,72,78, 94-96 
 cMMP 2,9       
 sFGF       
 sSDF-1       
MDS VEGF ↑BM ++ +/- Unk Survival (VEGF); 69,72,96-9 
 cMMP-2,cMMP-9     blast % (MVD);  
 sSDF-1     TNFα generation; ALIP  
 sIL-8       
MMM  ↑BM Unk Unk Unk Survival (MVD)* 99 
      Splenomegaly (MVD)  
Myeloma s,cVEGF ↑BM Overall survival (MVD) 20,74,100-102 
 sbFGF*     Trisomy 13, CRP, B2 
 cMMP-2     ↑ stromal IL-6  
ALL s,ubFGF ↑BM Unk Unk Unk  72,103,104 
 sSDF-1       
 cMMP-2,cMMP-9       
NHL sbFGF* ↑LN +/- Unk Survival* (VEGF, FGF) 75-77,105,106 
 s,cVEGF*     High grade histology (VEGF, FGF)  
      Advanced stage  
CLL s,cVEGF ↑BM Unk Unk Advanced stage (MVD, VEGF) 72,73,107-109 
 c,sbFGF     Time to progression (VEGF)  
 sSDF-1     Fludarabine resistance (cFGF)  
CML sVEGF ↑BM Unk Unk  72,73 
 c,sbFGF       
 sSDF-1       
Angiogenic Receptors
MalignancyPro-Angiogenic MoleculesMVDVEGFR-1VEGFR-2FGF-RClinical/Biological CorrelationReference
Abbreviations: CR, complete remission; s, serum or plasma; u, urinary; c, cellular; CRP, C-reactive protein; B2M, beta2-microglobulin; LN, lymph node; MMM, myelofibrosis with myeloid metaplasia; Unk, unknown; IL-6, interleukin-6; ALIP, abnormal localized immature precursors; AML, acute myelogenous leukemia; MDS, myelodysplastic syndrome; ALL, acute lymphoblastic leukemia; NHL, non-Hodgkin's lymphoma; CLL, chronic lymphocytic leukemia; CML, chronic myeloid leukemia; MVD, microvessel density; OS, overall survival; DFS, disease-free survival; BM, bone marrow; VEGF, vascular endothelial growth factor; FGF, fibroblast growth factor; bFGF, basic FGF; SDF-1, stromal-derived factor 1 
* Denotes independent prognostic factor. 
AML s,cVEGF* ↑BM ++ +/- Unk Leukocytosis; ↓CR*, OS*, DFS* (VEGF) 68,69,71,72,78, 94-96 
 cMMP 2,9       
 sFGF       
 sSDF-1       
MDS VEGF ↑BM ++ +/- Unk Survival (VEGF); 69,72,96-9 
 cMMP-2,cMMP-9     blast % (MVD);  
 sSDF-1     TNFα generation; ALIP  
 sIL-8       
MMM  ↑BM Unk Unk Unk Survival (MVD)* 99 
      Splenomegaly (MVD)  
Myeloma s,cVEGF ↑BM Overall survival (MVD) 20,74,100-102 
 sbFGF*     Trisomy 13, CRP, B2 
 cMMP-2     ↑ stromal IL-6  
ALL s,ubFGF ↑BM Unk Unk Unk  72,103,104 
 sSDF-1       
 cMMP-2,cMMP-9       
NHL sbFGF* ↑LN +/- Unk Survival* (VEGF, FGF) 75-77,105,106 
 s,cVEGF*     High grade histology (VEGF, FGF)  
      Advanced stage  
CLL s,cVEGF ↑BM Unk Unk Advanced stage (MVD, VEGF) 72,73,107-109 
 c,sbFGF     Time to progression (VEGF)  
 sSDF-1     Fludarabine resistance (cFGF)  
CML sVEGF ↑BM Unk Unk  72,73 
 c,sbFGF       
 sSDF-1       
Table 5.

Anti-angiogenic agents in clinical trials.

TargetAgent (Source)DescriptionActionRef.
Abbreviations: MoAb, monoclonal antibody; VEGF, vascular endothelial growth factor; EC, endothelial cell; ECM, extracellular matrix; MMP, matrix metalloprotease; TKI, tyrosine kinase inhibitor; IFN, interferon; FTI, farnesyl transferase inhibitor; i.v., intravenous; po, per oral; IL, interleukin; FGF, fibroblast growth factor; bFGF, basic FGF; TNFα, tumor necrosis factor-α 
Angiogenic Factors Bevacizumab (Genentech) humanized MoAb VEGF neutralization 110,111 
 IM862 (Cytran) small peptide ↓VEGF + bFGF production, ‐IL-12 112 
 IFNα (Schering, Roche) recombinant protein ↓VEGF + bFGF production, 113 
    ↓EC proliferation  
     
Angiongenic Receptors 2C3 murine MoAb VEGFR-2 antagonist 114 
 CD101 (ImClone) rat MoAb VEGFR-2 antagonist 115 
 Angiozyme (Ribozyme) VEGFR catalytic ribozyme VEGFR-1,2 mRNA inactivation 116 
 SU5416 (Sugen) small molecule, i.v. VEGFR1,2-TKI, c-kit, PDGFR 92,117 
 SU11248 (Sugen) small molecule, po VEGFR1,2-TKI, c-kit, PDGFR, FGF  
 PTK787/ZK222584 (Novartis) small molecule, po VEGFR1,2-TKI, c-kit, PDGFR, FGF, c-fms 118 
 AG13736 (Agouron) small molecule, po VEGFR1,2-TKI, c-kit, PDGFR, FGF  
     
Vascular Adhesion Molecules SCH221153 (Schering) peptidomimetic αvβ3 + αvβ5 antagonist, EC apoptosis 119 
 Vitaxin (Ixsys) humanized MoAb αvβ3antagonist, EC apoptosis 120 
     
Activated Endothelial Cells Endostatin (EntreMed) collagen XVIII fragment ↓EC proliferation, EC apoptosis 121,122 
 Vasostatin calreticulin fragment ↓EC proliferation 123 
 TNP-470 (TAP Pharma.) fumagillin analogue ↓EC proliferation 124 
 2-methoxyestradiol estrogen metabolite EC apoptosis 125 
 Thalidomide (Celgene) phthalimidoglutarimide ↓EC proliferation (bFGF,VEGF) ↓TNFα 85 
 CC-5013 (Celgene) phtaloyl analogue ↓EC proliferation, ↓TNFα, Th1 stimulator 86 
 Arsenic trioxide (CTI) trivalent arsenical EC apoptosis, tumor cytotoxin 126 
 SCH66336 (Schering) non-peptide, FTI ↓EC proliferation, ↓VEGF 127,128 
 R115777 (Janssen) non-peptide, FTO ↓EC proliferation, ↓VEGF 91 
     
ECM Remodelling Prinomastat (Agouron) non-peptide hydroxymate MMP inhibition 81 
TargetAgent (Source)DescriptionActionRef.
Abbreviations: MoAb, monoclonal antibody; VEGF, vascular endothelial growth factor; EC, endothelial cell; ECM, extracellular matrix; MMP, matrix metalloprotease; TKI, tyrosine kinase inhibitor; IFN, interferon; FTI, farnesyl transferase inhibitor; i.v., intravenous; po, per oral; IL, interleukin; FGF, fibroblast growth factor; bFGF, basic FGF; TNFα, tumor necrosis factor-α 
Angiogenic Factors Bevacizumab (Genentech) humanized MoAb VEGF neutralization 110,111 
 IM862 (Cytran) small peptide ↓VEGF + bFGF production, ‐IL-12 112 
 IFNα (Schering, Roche) recombinant protein ↓VEGF + bFGF production, 113 
    ↓EC proliferation  
     
Angiongenic Receptors 2C3 murine MoAb VEGFR-2 antagonist 114 
 CD101 (ImClone) rat MoAb VEGFR-2 antagonist 115 
 Angiozyme (Ribozyme) VEGFR catalytic ribozyme VEGFR-1,2 mRNA inactivation 116 
 SU5416 (Sugen) small molecule, i.v. VEGFR1,2-TKI, c-kit, PDGFR 92,117 
 SU11248 (Sugen) small molecule, po VEGFR1,2-TKI, c-kit, PDGFR, FGF  
 PTK787/ZK222584 (Novartis) small molecule, po VEGFR1,2-TKI, c-kit, PDGFR, FGF, c-fms 118 
 AG13736 (Agouron) small molecule, po VEGFR1,2-TKI, c-kit, PDGFR, FGF  
     
Vascular Adhesion Molecules SCH221153 (Schering) peptidomimetic αvβ3 + αvβ5 antagonist, EC apoptosis 119 
 Vitaxin (Ixsys) humanized MoAb αvβ3antagonist, EC apoptosis 120 
     
Activated Endothelial Cells Endostatin (EntreMed) collagen XVIII fragment ↓EC proliferation, EC apoptosis 121,122 
 Vasostatin calreticulin fragment ↓EC proliferation 123 
 TNP-470 (TAP Pharma.) fumagillin analogue ↓EC proliferation 124 
 2-methoxyestradiol estrogen metabolite EC apoptosis 125 
 Thalidomide (Celgene) phthalimidoglutarimide ↓EC proliferation (bFGF,VEGF) ↓TNFα 85 
 CC-5013 (Celgene) phtaloyl analogue ↓EC proliferation, ↓TNFα, Th1 stimulator 86 
 Arsenic trioxide (CTI) trivalent arsenical EC apoptosis, tumor cytotoxin 126 
 SCH66336 (Schering) non-peptide, FTI ↓EC proliferation, ↓VEGF 127,128 
 R115777 (Janssen) non-peptide, FTO ↓EC proliferation, ↓VEGF 91 
     
ECM Remodelling Prinomastat (Agouron) non-peptide hydroxymate MMP inhibition 81 
Table 6.

New chemotherapy drugs for non-Hodgkin's lymphomas.

Mechanism of ActionRepresentative Agents
Cytotoxic Oxaliplatin 
Apoptosis induction nucleosides, retinoids, arsenicals 
Protein kinase C inhibition Bryostatin 
Cyclin inhibition Flavopiridol, UCN-01, rapamycin 
Farnesyl transferase inhibitors R115777, BMS-214662 
Histone deacetylation Depsipeptide 
Antiangiogenesis Thalidomide, SU5416, SU6668 
Proteasome inhibition PS-134 
Antitutubulin agents Epophilone 
Mechanism of ActionRepresentative Agents
Cytotoxic Oxaliplatin 
Apoptosis induction nucleosides, retinoids, arsenicals 
Protein kinase C inhibition Bryostatin 
Cyclin inhibition Flavopiridol, UCN-01, rapamycin 
Farnesyl transferase inhibitors R115777, BMS-214662 
Histone deacetylation Depsipeptide 
Antiangiogenesis Thalidomide, SU5416, SU6668 
Proteasome inhibition PS-134 
Antitutubulin agents Epophilone 
Table 7.

Effects of flavopiridol on CLL cells.

Induction of apoptosis 
Independent of prior fludarabine 
Activation of caspase 3 
Cell cycle arrest 
Decreased expression of p53 protein 
Cytotoxicity independent of p53 status 
Down-regulation of Mcl-1 
Induction of apoptosis 
Independent of prior fludarabine 
Activation of caspase 3 
Cell cycle arrest 
Decreased expression of p53 protein 
Cytotoxicity independent of p53 status 
Down-regulation of Mcl-1 
Figure 2.

Phosphodiester substitutions in first generation antisense oligonucleotides.

Figure 2.

Phosphodiester substitutions in first generation antisense oligonucleotides.

Close modal
Figure 3.

Oligodeoxynucleotides targeting Bcl-2.

Figure 3.

Oligodeoxynucleotides targeting Bcl-2.

Close modal
Figure 4.

Targeting Bcl-2 may promote apoptosis following chemotherapy and irradiation.

Figure 4.

Targeting Bcl-2 may promote apoptosis following chemotherapy and irradiation.

Close modal
Figure 5.

Survival of cohorts of 6 mice treated with oligonucleotides alone (left panel) or with cyclophosphamide (CY) (right panel).

All six surviving animals treated with cyclophosphamide and G3139 sacrificed at 90 days with no molecular evidence of disease detected.

Saline=control animals ; AS=antisense oligonucleotide G3139 directed at bcl-2 ; SN=reverse sequence sense control ; MM= 2-base mismatch control.

Figure 5.

Survival of cohorts of 6 mice treated with oligonucleotides alone (left panel) or with cyclophosphamide (CY) (right panel).

All six surviving animals treated with cyclophosphamide and G3139 sacrificed at 90 days with no molecular evidence of disease detected.

Saline=control animals ; AS=antisense oligonucleotide G3139 directed at bcl-2 ; SN=reverse sequence sense control ; MM= 2-base mismatch control.

Close modal
Figure 6.

Interaction of vascular endothelial growth factor (VEGF) family members with their cognate receptors on endothelial cells.

The VEGF receptors harbor seven extracellular immunoglobulin (Ig)-homology domains. The fifth Ig domain of VEGFR-3 is cleaved after biosynthesis to yield a stable disulfide bond linking the Ig subunits. VEGFR-1 and VEGFR-2 initiate signals essential to the angiogenic response, whereas VEGFR-3 regulates lymph angiogenesis. NRP-1 (neuropilin-1) binds to the carboxy-terminal sequence of VEGF165 to enhance binding of the angiogenic molecule to the VEFGR-2 receptor.

Abbreviations: TK, tyrosine kinase; MMP, matrix metalloproteases; PIGF, placental growth factor.

Figure 6.

Interaction of vascular endothelial growth factor (VEGF) family members with their cognate receptors on endothelial cells.

The VEGF receptors harbor seven extracellular immunoglobulin (Ig)-homology domains. The fifth Ig domain of VEGFR-3 is cleaved after biosynthesis to yield a stable disulfide bond linking the Ig subunits. VEGFR-1 and VEGFR-2 initiate signals essential to the angiogenic response, whereas VEGFR-3 regulates lymph angiogenesis. NRP-1 (neuropilin-1) binds to the carboxy-terminal sequence of VEGF165 to enhance binding of the angiogenic molecule to the VEFGR-2 receptor.

Abbreviations: TK, tyrosine kinase; MMP, matrix metalloproteases; PIGF, placental growth factor.

Close modal
Figure 10.

The ubiquitin-proteasome pathway.

The proteasome is a multicentric protease complex that plays a pivotal role in cellular protein regulation. In order for a protein to be suitable for degradation, it must first be adorned with ubiquitin. The ubiquitin-proteasome pathway plays a critical role in the degradation of intracellular proteins involved in cell cycle control and tumor growth.

Figure 10.

The ubiquitin-proteasome pathway.

The proteasome is a multicentric protease complex that plays a pivotal role in cellular protein regulation. In order for a protein to be suitable for degradation, it must first be adorned with ubiquitin. The ubiquitin-proteasome pathway plays a critical role in the degradation of intracellular proteins involved in cell cycle control and tumor growth.

Close modal
Figure 11.

PS-341 is a specific and selective inhibitor of the 26S proteasome.

NF-κB is activated when the proteasome degrades the inhibitor protein 1κBa resulting in downregulation of multiple gene products. NF-κB plays a role in maintaining cell viability through the transcription of inhibitors of apoptosis. PS-341 can block activation of NF-κB, which may induce apoptosis and also make cells more sensitive to a variety of chemotherapy agents.

Figure 11.

PS-341 is a specific and selective inhibitor of the 26S proteasome.

NF-κB is activated when the proteasome degrades the inhibitor protein 1κBa resulting in downregulation of multiple gene products. NF-κB plays a role in maintaining cell viability through the transcription of inhibitors of apoptosis. PS-341 can block activation of NF-κB, which may induce apoptosis and also make cells more sensitive to a variety of chemotherapy agents.

Close modal
1
Alizadeh AA, Eisen MB, Davis RE, et al. Distinct types of diffuse large B-cell lymphoma identified by gene expression profiling.
Nature
.
2000
;
403
:
503
.
2
Ong ST, Le Beau MM. Chromosomal abnormalities and molecular genetics of non-Hodgkin's lymphoma.
Semin Oncol
.
1998
;
25
:
447
.
3
Gascoyne RD, Adomat SA, Krajewski S, et al. Prognostic significance of Bcl-2 protein expression and Bcl-2 gene rearrangement in diffuse aggressive non-Hodgkin's lymphoma.
Blood
.
1997
;
90
:
244
.
4
Reed JC. Bcl-2 family proteins: regulators of apoptosis and chemoresistance in hematologic malignancies.
Semin Hematol
.
1997
;
34
:
9
.
5
Yang E, Korsmeyer SJ. Molecular thanatopsis: a discourse on the BCL2 family and cell death.
Blood
.
1996
;
88
:
386
.
6
Reed JC. Double identity for proteins of the Bcl-2 family.
Nature
.
1997
;
387
:
773
.
7
Zamecnik PC, Stephenson ML. Inhibition of Rous sarcoma virus replication and cell transformation by a specific oligodeoxynucleotide.
Proc Natl Acad Sci U S A
.
1978
;
75
:
280
.
8
Stephenson ML, Zamecnik PC. Inhibition of Rous sarcoma viral RNA translation by a specific oligodeoxyribonucleotide.
Proc Natl Acad Sci U S A
.
1978
;
75
:
285
.
9
Mizuno T, Chou MY, Inouye M. A unique mechanism regulating gene expression: translational inhibition by a complementary RNA transcript (micRNA).
Proc Natl Acad Sci U S A
.
1984
;
81
:
1966
.
10
Izant JG, Weintraub H. Inhibition of thymidine kinase gene expression by anti-sense RNA: a molecular approach to genetic analysis.
Cell
.
1984
;
36
:
1007
.
11
Wintersberger U. Ribonucleases H of retroviral and cellular origin.
Pharmacology & Therapeutics
.
1990
;
48
:
259
.
12
Sohail M, Hochegger H, Klotzbucher A, Guellec RL, Hunt T, Southern EM. Antisense oligonucleotides selected by hybridisation to scanning arrays are effective reagents in vivo.
Nucleic Acids Res
.
2001
;
29
:
2041
.
13
Ho SP, Bao Y, Lesher T, et al. Mapping of RNA accessible sites for antisense experiments with oligonucleotide libraries.
Nat Biotechnol
.
1998
;
16
:
59
.
14
Ho SP, Britton DH, Stone BA, et al. Potent antisense oligonucleotides to the human multidrug resistance-1 mRNA are rationally selected by mapping RNA-accessible sites with oligonucleotide libraries.
Nucleic Acids Res
.
1996
;
24
:
1901
.
15
Agrawal S, Temsamani J, Tang JY. Pharmacokinetics, biodistribution, and stability of oligodeoxynucleotide phosphorothioates in mice.
Proc Natl Acad Sci U S A
.
1991
;
88
:
7595
.
16
Raynaud FI, Orr RM, Goddard PM, et al. Pharmacokinetics of G3139, a phosphorothioate oligodeoxynucleotide antisense to bcl-2, after intravenous administration or continuous subcutaneous infusion to mice.
J Pharmacol Exp Ther
.
1997
;
281
:
420
.
17
Jansen B, Schlagbauer-Wadl H, Brown BD, et al. bcl-2 antisense therapy chemosensitizes human melanoma in Scid mice.
Nat Med
.
1998
;
4
:
232
.
18
Monia BP, Johnston JF, Geiger T, Muller M, Fabbro D. Antitumor activity of a phosphorothioate antisense oligodeoxynucleotide targeted against C-raf kinase [see comments].
Nat Med
.
1996
;
2
:
668
.
19
Weiner GJ, Liu HM, Wooldridge JE, Dahle CE, Krieg AM. Immunostimulatory oligodeoxynucleotides containing the CpG motif are effective as immune adjuvants in tumor antigen immunization.
Proc Natl Acad Sci U S A
.
1997
;
94
:
10833
.
20
Wooldridge JE, Ballas Z, Krieg AM, Weiner GJ. Immunostimulatory oligodeoxynucleotides containing CpG motifs enhance the efficacy of monoclonal antibody therapy of lymphoma.
Blood
.
1997
;
89
:
2994
.
21
Ballas ZK, Rasmussen WL, Krieg AM. Induction of NK activity in murine and human cells by CpG motifs in oligodeoxynucleotides and bacterial DNA.
J Immunol
.
1996
;
157
:
1840
.
22
Krieg AM, Yi AK, Matson S, et al. CpG motifs in bacterial DNA trigger direct B-cell activation.
Nature
.
1995
;
374
:
546
.
23
Saijo Y, Uchiyama B, Abe T, Satoh K, Nukiwa T. Contiguous four-guanosine sequence in c-myc antisense phosphorothioate oligonucleotides inhibits cell growth on human lung cancer cells: possible involvement of cell adhesion inhibition.
Jpn J Cancer Res
.
1997
;
88
:
26
.
24
Buckheit RW Jr, Roberson JL, Lackman-Smith C, Wyatt JR, Vickers TA, Ecker DJ. Potent and specific inhibition of HIV envelope-mediated cell fusion and virus binding by G quartet-forming oligonucleotide (ISIS 5320).
AIDS Res Hum Retroviruses
.
1994
;
10
:
25
Cotter FE, Johnson P, Hall P, et al. Antisense oligonucleotides suppress B-cell lymphoma growth in a SCID-hu mouse model.
Oncogene
.
1994
;
9
:
3049
.
26
Cotter FE. Antisense therapy for lymphomas.
Hematol Oncol
.
1997
;
15
:
3
.
27
Reed JC, Miyashita T, Takayama S, et al. BCL-2 family proteins: regulators of cell death involved in the pathogenesis of cancer and resistance to therapy.
J Cell Biochem
.
1996
;
60
:
23
.
28
Reed JC. Bcl-2: prevention of apoptosis as a mechanism of drug resistance. Hematol Oncol.
Clin North Am
.
1995
;
9
:
451
.
29
Miyashita T, Reed JC. bcl-2 gene transfer increases relative resistance of S49.1 and Wehi7.2 lymphoid cells to cell death and DNA fragmentation induced by glucocorticoids and multiple chemotherapeutic drugs.
Cancer Res
.
1992
;
52
:
5407
.
30
Miyashita T, Reed JC. Bcl-2 oncoprotein blocks chemotherapy-induced apoptosis in a human leukemia cell line.
Blood
.
1993
;
81
:
151
.
31
Kamesaki S, Kamesaki H, Jorgensen TJ, Tanizawa A, Pommier Y, Cossman J. bcl-2 protein inhibits etoposide-induced apoptosis through its effects on events subsequent to topoisomerase II-induced DNA strand breaks and their repair [published erratum appears in Cancer Res 1994 Jun 1;54(11):3074].
Cancer Res
.
1993
;
53
:
4251
.
32
Walton MI, Whysong D, PM OC, Hockenbery D, Korsmeyer SJ, Kohn KW. Constitutive expression of human Bcl-2 modulates nitrogen mustard and camptothecin induced apoptosis.
Cancer Res
.
1993
;
53
:
1853
.
33
Reed JC, Kitada S, Takayama S, Miyashita T. Regulation of chemoresistance by the bcl-2 oncoprotein in non-Hodgkin's lymphoma and lymphocytic leukemia cell lines.
Ann Oncol
.
1994
;
5
:
61
.
34
Keith FJ, Bradbury DA, Zhu YM, Russell NH. Inhibition of bcl-2 with antisense oligonucleotides induces apoptosis and increases the sensitivity of AML blasts to Ara-C.
Leukemia
.
1995
;
9
:
131
.
35
Klasa RJ, Bally MB, Ng R, Goldie JH, Gascoyne RD, Wong FM. Eradication of human non-Hodgkin's lymphoma in SCID mice by BCL-2 antisense oligonucleotides combined with low-dose cyclophosphamide.
Clin Cancer Res
.
2000
;
6
:
2492
.
36
Dyer MJ, Lillington DM, Bastard C, et al. Concurrent activation of MYC and BCL2 in B cell non-Hodgkin lymphoma cell lines by translocation of both oncogenes to the same immunoglobulin heavy chain locus.
Leukemia
.
1996
;
10
:
1198
.
37
Macpherson N, Lesack D, Klasa R, et al. Small noncleaved, non-Burkitt's (Burkitt-Like) lymphoma: cytogenetics predict outcome and reflect clinical presentation.
J Clin Oncol
.
1999
;
17
:
1558
.
38
Webb A, Cunningham D, Cotter F, et al. BCL-2 antisense therapy in patients with non-Hodgkin lymphoma.
Lancet
.
1997
;
349
:
1137
.
39
Waters JS, Webb A, Cunningham D, et al. Phase I clinical and pharmacokinetic study of bcl-2 antisense oligonucleotide therapy in patients with non-Hodgkin's lymphoma.
J Clin Oncol
.
2000
;
18
:
1812
.
40
Jansen B, Wacheck V, Heere-Ress E, et al. Chemosensitisation of malignant melanoma by BCL2 antisense therapy.
Lancet
.
2000
;
356
:
1728
.
1
Hanahan D. Signaling vascular morphogenesis and maintenance.
Science
.
1997
;
277
:
48
–50.
2
Risau W. Mechanisms of angiogenesis.
Nature
.
1997
;
386
:
671
–674.
3
Liekens S, DeClercq E, Neyts J. Angiogenesis: regulators and clinical applications.
Biochemical Pharmacology
.
2001
;
61
:
253
–270.
4
Leung DW, Cachianes G, Kuang W-J, Goeddel DV, Ferrara N. Vascular endothelial growth factor is a secreted angiogenic mitogen.
Science
.
1989
;
246
:
1306
–1309.
5
Folkman J. Clinical applications of research on angiogenesis.
N Engl J Med
.
1995
;
333
:
1757
–1763.
6
Tischer E, Mitchell R, Hartman T, et al. The human gene for vascular endothelial growth factor. Multiple protein forms are encoded through alternative exon splicing.
J Biol Chem
.
1991
:
266
:
11947
–11954.
7
Gitay-Goren H, Soker S, Vlodavsky I, Neufeld G. The binding of vascular endothelial growth factor to its receptors is dependent on cell surface associated heparin-like molecules.
J Biol Chem
.
1992
;
267
:
6093
–6098.
8
Roberts R, Gallagher J, Spooncer E, Allen TD, Bloomfield F, Dexter RM. Heparan-sulfate bound growth factors: a mechanism for stromal cell mediated haemopoiesis.
Nature
.
1998
;
332
:
376
–378.
9
Keyt BA, Berleau LT, Nguyen HV. The carboxyl-terminal domain (111-165) of vascular endothelial growth factor is critical for its mitogenic potency.
J Biol Chem
.
1996
;
271
:
7788
–7795.
10
Soker S, Gollamudi-Payne S, Fidder H, Charmahelli H, Klagbrun M. Inhibition of vascular endothelial growth factor (VEGF)-induced endothelial cell proliferation by a peptide corresponding to the exon 7-encoded domain of VEGF165.
J Biol Chem
.
1997
;
272
:
31582
–31588.
11
Brogi E, Wu T, Namiki A, Isner JM. Indirect angiogenic cytokines upregulate VEGF and bFGF gene expression in vascular smooth muscle cells, whereas hypoxia upregulates VEGF expression only.
Circulation
.
1994
;
90
:
649
–652.
12
Namiki A, Brogi E, Kearney M, et al. Hypoxia induces vascular endothelial growth factor in cultured human endothelial cells.
J Biol Chem
.
1995
;
270
:
31189
–31195.
13
Waltenberger J, Mayr U, Pentz S, Hombach V. Functional upregulation of the vascular endothelial growth factor receptor KDR by hypoxia.
Circulation
.
1996
;
94
:
1647
–1654.
14
De Vries C, Escobeo JA, Ueno H, Houck K, Ferarra N, Williams LT. The fms-like tyrosine kinase receptor, a receptor for vascular endothelial growth factor.
Science
.
1992
;
255
:
989
–991.
15
Terman BL, Dougher-Vermazen M, Carrion ME, et al. Identification of the KDR tyrosine kinase as a receptor for vascular endothelial cell growth factor.
Biochem Biophys Res Commun
.
1992
:
187
:
1579
–1586.
16
Hamada K, Oike Y, Nobuyuki T, et al. VEGF-C signaling pathways through VEGFR-2 and VEGFR-3 in vasculoangiogenesis and hematopoiesis.
Blood
.
2000
;
96
:
3793
–3800.
17
Joukov V, Pajusola K, Kaipainen A, et al. A novel vascular endothelial growth factor, VEGF-C, is a ligand for the Flt4 (VEGFR-3) and KDR (VEGFR-2) receptor tyrosine kinases.
EMBO J
.
1996
:
15
:
290
–298.
18
Waltenberger J, Claesson-Welsh L, Siegbahn A, Shibuya M, Heldin C-H. Different signal transduction properties of KDR and Flt1, two receptors for vascular endothelial growth factor.
J Biol Chem
.
1994
;
269
:
26988
–26995.
19
Pajusola K, Aprelikova O, Pelicci G, Weich H, Claesson-Welsh L, Alitalo K. Signaling properties of FLT4, a proteolytically processed receptor tyrosine kinase related to two VEGF receptors.
Oncogene
.
1994
:
9
:
3545
–3555.
20
Bellamy WT, Richter L, Frutiger Y, Grogan TM. Expression of vascular endothelial growth factor (VEGF) and its receptors in hematopoietic malignancies.
Cancer Res
.
1999
;
59
:
728
–733.
21
Ziegler BL, Valtieri M, Porada GA, et al. KDR receptor: a key marker defining hematopoietic stem cells.
Science
.
1999
;
285
:
1553
–1558.
22
Barleon B, Sozzani S, Zhou D, Weich HA, Mantovani A, Marme D. Migration of human monocytes in response to vascular endothelial growth factor (VEGF) is mediated via the VEGF receptor flt-1.
Blood
.
1996
;
87
:
3336
–3343.
23
Sawano A, Iwai S, Sakurai Y, Ito M, Shitara K, Nakahata T, Shibuya M. Flt-1, vascular endothelial growth factor receptor 1, is a novel cell surface marker for the lineage of monocyte-macrophages in humans.
Blood
.
2001
:
97
:
785
–791.
24
Kaipainen A, Korhonen J. Mustonen T, et al. Expression of the fms-like tyrosine kinase 4 gene becomes restricted to lymphatic endothelium during development.
Proc Natl Acad Sci U S A
.
1995
;
92
:
3566
–3570.
25
Carmeliet P, Ferreira V, Breier G, et al. Abnormal blood vessel development and lethality in embryos lacking a single VEGF allele.
Nature
.
1996
;
380
:
435
–439.
26
Ferrara N, Carver-Moore K, Chen H, et al. Heterozygous embryonic lethality induced by targeted inactivation of the VEGF gene.
Nature
.
1996
;
380
:
439
–442.
27
Shalaby F, Rossant J, Yamaguchi TP, et al. Failure of blood island formation and vasculogenesis in Flk-1-deficient mice.
Nature
.
1995
;
376
:
62
–66.
28
Hiratsuka S, Minowa O, Kuno J, Noda T, Shibuya M. Flt-1 lacking the tyrosine kinase domain is sufficient for normal development and angiogenesis in mice.
Proc Natl Acad Sci U S A
.
1998
;
95
:
9349
–9354.
29
Kukk E, Lymboussaki A, Taira S, et al. VEGF-C receptor binding and pattern of expression with VEGFR-3 suggests a role in lymphatic vascular development.
Development
.
1996
;
122
:
3829
–3837.
30
Achen MG, Jeltsch M, Kukk E, et al. Vascular endothelial growth factor D (VEGF-D) is a ligand for the tyrosine kinases VEGF receptor 2 (FLK1) and VEGF receptor 3 (Flt4).
Proc Natl Acad Sci U S A
.
1998
;
95
:
548
–553.
31
Olofsson B, Korpelainen E, Pepper MS, et al. Vascular endothelial growth factor B (VEGF-B) binds to VEGF receptor-1 and regulates plasminogen activator activity in endothelial cells.
Proc Natl Acad Sci U S A
.
1998
;
95
:
11709
–11714.
32
Meyer M, Clauss M, Lepple-Wienhues A, et al. A novel vascular endothelial growth factor encoded by Orf virus, VEGF-E, mediates angiogenesis via signaling through VEGFR-2 (KDR) but not VEGFR-1 (Flt-1) receptor tyrosine kinases.
EMBO J
.
1999
;
18
:
363
–374.
33
Yamada Y, Takakura N, Yasue H, Ogawa H, Fujisawa H, Suda T. Exogenous clustered neuropilin 1 enhances vasculogenesis and angiogenesis.
Blood
.
2001
;
97
:
1671
–1678.
34
Veikkola T, Karkkainen M, Claesson-Welsh L, Alitalo K. Regulation of angiogenesis via vascular endothelial growth factor receptors.
Cancer Research
.
2000
;
60
:
203
–212.
35
Kroll J, Waltenberger J. The vascular endothelial growth factor receptor KDR activates multiple signal transduction pathways in porcine aortic endothelial cells.
J Biol Chem
.
1997
;
272
:
32521
–32527.
36
Igarashi K, Shigeta K, Isohara T, Yaman T, Uno I. Sck interacts with KDR and Flt1 via its SH2 domain.
Biochm Biophys Res Commun
.
1998
;
251
:
77
–82.
37
Takahashi T, Ueno H, Shibuya M. VEGF activates protein kinase C-dependent, but Ras-independent Raf-independent Rad-MEK-MAP kinase pathway for DNA synthesis in primary endothelial cells. Oncogene. 1999;18;2221-2230.
38
Igarashi K, Isohara T, Kato T, Shigeta K, Yamano T, Uno I. Tryrosine 1213 of Flt1 is a major binding site of Nck and SHP-2.
Biochem Biophys Res Commun
.
1998
;
246
:
95
–99.
39
Xia P, Aiello LP, Ishii H, et al. Characterization of vascular endothelial growth factor's effect on the activation of protein kinase C, its isoforms, and endothelial cell growth.
J Clin Investig
.
1996
;
98
:
2018
–2026.
40
Gerber H-P, McMurtrey A, Kowalski J, et al. Vascular endothelial growth factor regulates endothelial cell survival through the phosphatidylinositol 3'-kinase/Akt signal transduction pathway.
J Biol Chem
.
1998
;
273
:
30336
–30343.
41
Coffer PJ, Jin J, Woodgett JR. Protein kinase B (c-Akt): a multifunctional mediator of phosphatidylinositol 3-kinase activation.
Biochem J
.
1998
;
335
:
1
–13.
42
Liu ZY, Ganju RK, Wang JF, et al. Characterization of signal transduction pathways in human bone marrow endothelial cells.
Blood
.
1997
;
90
:
2253
–2259.
43
Byzova T, Goldman C, Pampori N, et al. A mechanism for modulation of cellular responses to VEGF: activation of the integrins.
Molecular Cell
.
2000
;
6
:
851
–860.
44
Soldi R, Mitola S, Strasly M, Defilippi P, Tarone G, Bussolino F. Role of αvβ3 in the activation of vascular endothelial growth factor receptor-2.
EMBO
.
1999
;
18
:
882
–892.
45
Carmeliet P, Lampugnani M-G, Moons L, et al. Targeted deficiency or cytosolic truncation of the VE-cadherin gene in mice impairs VEGF-mediated endothelial survival and angiogenesis.
Cell
.
1999
;
98
:
147
–157.
46
Semenza GL. Regulation of mammalian O2 homeostasis by hypoxia-inducible factor 1.
Ann Rev Cell Dev Biol
.
1999
;
15
:
551
–578.
47
Salceda S, Caro J. Hypoxia-inducible factor 1a (HIF-1a) protein is rapidly degraded by the ubiquitin-proteasome system under normoxic conditions. Its stabilization by hypoxia depends on redox-induced changes.
J Biol Chem
.
1997
;
272
:
22642
–22647.
48
Yu F, White S, Zhao Q, Lee F. Dynamic, site-specific interaction of hypoxia-inducible factor-1a with the von Hippel-Lindau tumor suppressor protein.
Cancer Res
.
2001
;
61
:
4136
–4142.
49
Kerbel RS, Viloria-Petit A, Okada F, Rak J. Establishing a link between oncogenes and tumor angiogenesis.
Mol Med
.
1998
;
4
:
286
–295.
50
Okada F, Rak JW, St Croix B, et al. Impact of oncogenes in tumor angiogenesis: mutant K-ras up- regulation of vascular endothelial growth factor/vascular permeability factor is necessary, but not sufficient for tumorigenicity of human colorectal carcinoma cells.
Proc Natl Acad Sci U S A
.
1998
;
95
:
3609
–3614.
51
Arbiser JL, Moses MA, Fernandez CA, et al. Oncogenic H-ras stimulates tumor angiogenesis by two distinct pathways.
Proc Natl Acad Sci U S A
.
1997
;
94
:
861
–866.
52
Ravi R, Mookerjee B, Bhujwalla ZM, et al. Regulation of tumor angiogenesis by p53-induced degradation of hypoxia-inducible factor 1a.
Genes Dev
.
2000
;
14
:
34
–44.
53
Guillemin K, Drasnow MA. The hypoxic response: Huffing and HIFing.
Cell
.
1997
;
89
:
9
–12.
54
Marshall C. The embryonic origins of human haematopoiesis [review].
Br J Haematol
.
2001
;
112
:
838
–850.
55
Eichmann A, Corbel C, Nataf V, Vaigot P, Breant C, Le Douarin NM. Ligand-dependent development of the endothelial and hemapoietic lineages from embryonic mesodermal cells expressing vascular endothelial growth factor receptor 2.
Proc Natl Acad Sci U S A
.
1997
;
94
:
5141
–5146.
56
Gunsilius E, Duba H-C, Petzer A, et al. Evidence from leukaemia model for maintenance of vascular endothelium by bone-marrow-derived endothelial cells.
Lancet
.
2000
;
355
:
1688
–1691.
57
Gao Z, McAlister V, Williams G. Repopulation of liver endothelium by bone-marrow-derived cells.
Lancet
.
2001
;
357
:
932
–933.
58
Ikpeazu C, Davidson M, Halteman D, Goodman S, Browning P, Brandt S. Donor origin of circulating endothelial progenitors after allogeneic bone marrow transplantation.
Biology of Blood and Marrow Transplantation
.
2000
;
6
:
301
–308.
59
Peichev M, Naiyer A, Pereira D, et al. Expression of VEGFR-2 and AC133 by circulating human CD34+ cells identifies a population of functional endothelial precursors.
Blood
.
2000
;
95
:
952
–958.
60
Gabrilovich D, Ishida T, Oyama T, et al. Vascular endothelial growth factor inhibits the development of dendritic cells and dramatically affects the differentiation of multiple hematopoietic lineages in vivo.
Blood
.
1998
;
92
:
4150
–4166.
61
Tordjman R, Delaire S, Plouet J, et al. Erythroblasts are a source of angiogenic factors.
Blood
.
2001
;
97
:
1968
–1974.
62
Broxmeyer HE, Cooper S, Li ZH, et al. Myeloid progenitor cell regulatory effects of vascular endothelial cell growth factor.
Int J Hematol
.
1995
;
62
:
203
–215.
63
Gerber H-P, Vu TH, Ryan AM, Dowalski J, Werb Z, Ferrara N. VEGF couples hypertrophic cartilage remodeling, ossification and angiogenesis during endochondral bone formation.
Nat Med
.
1999
;
5
:
623
–628.
64
Niida S, Kaku M, Amano H, et al. Vascular endothelial growth factor can substitute for macrophage colony-stimulating factor in the support of osteoclastic bone resorption.
J Exp Med
.
1999
;
190
:
293
–298.
65
Kuramoto K, Uesaka T, Kimura A, Kobayashi M, Watanabe H, Katoh O. ZK7, a novel zinc finger gene, is induced by vascular endothelial growth factor and inhibits apoptotic death in hematopoietic cells.
Cancer Research
.
2000
;
60
:
425
–430.
66
Katoh O, Takahashi T, Oguri T, et al. Vascular endothelial growth factor inhibits apoptotic death in hematopoietic cells after exposure to chemotherapeutic drugs by inducing MCL1 acting as an anti-apoptotic factor.
Cancer Research
.
1998
;
58
:
5565
–5569.
67
Katoh O, Tauchi H, Kawaishi K, Kimura A, Satow Y. Expression of the vascular endothelial growth factor (VEGF) receptor gene, KDR, in hematopoietic cells and inhibitory effect of VEGF on apoptotic cell death caused by ionizing radiation.
Cancer Research
.
1995
;
55
:
5687
–5692.
68
Fiedler W, Graeven U, Ergun S, et al. Vascular endothelial growth factor, a possible paracrine growth factor in human acute myeloid leukemia.
Blood
.
1997
;
89
:
1870
–1875.
69
Bellamy W, Richter L, Dirjani D, et al. Vascular endothelial cell growth factor is an autocrine promoter of abnormal localized immature myeloid precursors and leukemia progenitor formation in myelodysplastic syndromes.
Blood
.
2001
;
97
:
1427
–1434.
70
List AF, Glinsmann-Gibson B, Stadheim C, Meuillet E, Bellamy W, Powis G. VEGFR-1 (Flt-1) and VEGFR-2 (KDR) stimulate the proliferation of AML cells via the P13-kinase and Akt/protein kinase-b (PKB) signal pathway.
Blood
.
2000
;
96
:
301a
.
71
Dias S, Hattori K, Zhu Z, et al. Autocrine stimulation of VEGFR-2 activates human leukemic cell growth and migration.
J Clin Invest
2000
;
106
:
511
–521.
72
Aguayo A, Kantarjian H, Manshouri T, et al. Angiogenesis in acute and chronic leukemias and myelodysplastic syndromes.
Blood
.
2000
;
96
:
2240
–2245.
73
Krejci P, Dvorakova D, Krahulcova E. et al. FGF-2 abnormalities in B cell chronic lymphocytic and chronic myeloid leukemias.
Leukemia
.
2001
;
15
:
228
–237.
74
Plowright E, Li Z, Bergsagel L, et al. Ectopic expression of fibroblast growth factor receptor 3 promotes myeloma cell proliferation and prevents apoptosis.
Blood
.
2000
;
95
:
992
–998.
75
Salven P, Orpana A, Teerenhovi L, Joensuu H. Simultaneous elevation in the serum concentrations of the angiogenic growth factors VEGF and bFGF is an independent predictor of poor prognosis in non-Hodgkin lymphoma: a single-institution study of 200 patients.
Blood
.
2000
;
96
:
3712
–3718.
76
Foss HD, Araujo A, Demel G, Klotzbach H, Hummel M, Stein H. Expression of vascular endothelial growth factor in lymphomas and Castleman's disease.
J Pathol
.
1997
:
183
:
44
–50.
77
Bellamy WT, Richter L, Frutiger Y, et al. Expression of vascular endothelial cell growth factor (VEGF) and its receptors in non-Hodgkin's lymphoma (NHL).
Proc Amer Assoc Cancer Res
.
1999
;
40
:
227
.
78
Aguayo A, Estey E, Kantarjian H, et al. Cellular vascular endothelial growth factor is a predictor of outcome in patients with acute myeloid leukemia.
Blood
.
1999
;
94
:
3717
–3721.
79
Fusetti L, Pruneri G, Gobbi A, et al. Human myeloid and lymphoid malignancies in the non-obese diabetic/severe combined immunodeficiency mouse model: frequency of apoptotic cells in solid tumors and efficiency and speed of engraftment correlate with vascular endothelial growth factor production.
Cancer Res
.
2000
;
60
:
2527
–2534.
80
Singhal S, Mehta J, Desikan R, et al. Antitumor activity of thalidomide in refractory multiple myeloma.
N Engl J Med
.
1999
;
341
:
1565
–1571.
81
Price A, Shi Q, Morris D, et al. Marked inhibition of tumor growth in a malignant glioma tumor model by a novel synthetic matrix metalloproteinase inhibitor AG3340.
Clinical Cancer Research
.
1999
;
5
:
845
–854.
82
Gearing AJH, Beckett P, Christodoulou M, et al. Processing of tumor necrosis factor-α precursor by metalloproteinases.
Nature
.
1994
;
370
:
555
–557.
83
Black RA, Rauch CT, Kozlosky CJ, et al. A metalloproteinase disintegrin that releases tumor necrosis factor-alpha from cells.
Nature
.
1997
;
385
:
729
–733.
84
Kayagaki N, Kawasaki A, Ebata T, et al. Metalloproteinase-mediated release of human Fas ligand.
J Exp Med
.
1995
;
182
:
1777
–1783.
85
D'Amato RJ, Loughman MS, Flynn E, Folkman J. Thalidomide is an inhibitor of angiogenesis.
Proc Natl Acad Sci U S A
.
1994
;
91
:
4082
–4085.
86
Corral LG, Haslett PAJ, Muller GW, et al. Differential cytokine modulation and T cell activation by two distinct classes of thalidomide analogs that are potent inhibitors of TNF-α.
J Immunol
.
1999
;
163
:
380
–386.
87
Hideshima T, Chauhan D, Shima Y, et al. Thalidomide and its analogs overcome drug resistance of human multiple myeloma cells to conventional therapy.
Blood
.
2000
;
96
:
2943
–2950.
88
Davies FE, Raje N, Hideshima T, et al. Thalidomide and immunomodulatory derivatives augment natural killer cytotoxicity in multiple myeloma. Blood. 2001;98:210-216,
89
Raza A, Meyer P, Dutt T, et al. Thalidomide produces transfusion independence in patients with long-standing refractory anemias and myelodysplastic syndromes.
Blood
2001
;
9
:
958
–965.
90
Shi YP, Ferrara N. Oncogenic ras fails to restore an in vivo tumorigenic phenotype in embryonic stem cells lacking vascular endothelial growth factor (VEGF).
Biochem Biophys Res Commun
.
1999
;
254
:
480
–483.
91
Karp JE, Lancet JE, Kaufmann SH, et al. Clinical and biologic activity of the farnesyltransferase inhibitor R115777 in adults with refractory and relapsed acute leukemias: a phase I clinical-laboratory correlative trial.
Blood
.
2001
;
97
:
3361
–3369.
92
Smolich BD, Yuen HA, West KA, Giles FJ, Albitar M, Cherrington JM. The antiangiogenic protein kinase inhibitors SU5416 and SU6668 inhibit the SCF receptor (c-kit) in a human myeloid leukemia cell line and in acute myeloid leukemia blasts.
Blood
.
2001
;
97
:
1413
–1421.
93
Mesters RM, Padro T, Bieker R, et al. Stable remission after administration of the receptor tyrosine kinase inhibitor SU5416 in a patient with refractory acute myeloid leukemia.
Blood
.
2001
;
98
:
241
–243.
94
Hussong JW, Rodgers GM, Shami PJ. Evidence of increased angiogenesis in patients with acute myeloid leukemia.
Blood
.
2000
;
95
:
309
–313.
95
Padro T, Ruiz S, Bieker R, et al. Increased angiogenesis in the bone marrow of patients with acute myeloid leukemia.
Blood
.
2000
;
95
:
2637
–2644.
96
Janawska-Wieczorek A, Marquuez LA, Matsuzaki A, et al. Expression of matrix metalloproteinases (MMP-2 and –9) and tissue inhibitors of metalloproteinases (TIMP-1 and –2) in acute myelogenous leukaemia blasts: comparison with normal bone marrow cells.
Br J Hematol
.
1999
;
105
:
402
–411.
97
Pruneri G, Bertolini F, Soligo D, et al. Angiogenesis in myelodysplastic syndromes.
Br J Cancer
.
1999
;
81
:
1398
–1401.
98
Arimura K, Atima N, Ohtsubo H, et al.
Matrix metalloproteinase inhibitor inhibits apoptosis induction of MDS bone marrow cells
.
1999
;
94
:
105a
.
99
Mesa RA, Hanson CA, Rajkumar V, Schroeder G, Tefferi A. Evaluation and clinical correlations of bone marrow angiogenesis in myelofibrosis with myeloid metaplasia.
Blood
.
2000
;
96
:
3374
–9980.
100
Rajkumar SV, Leong T, Roche PC, et al. Prognostic value of bone marrow angiogenesis in multiple myeloma.
Clinical Cancer Research
.
2000
;
6
:
3111
–3116.
101
Vacca A, Ribatti D, Presta M, et al. Bone marrow neovascularization, plasma cell angiogenic potential, and matrix metalloproteinase-2 secretion parallel progression of human multiple myeloma.
Blood
.
1999
;
93
:
3064
–3073.
102
Dankbar B, Padro T, Leo R, et al. Vascular endothelial growth factor and interleukin-6 in paracrine trumor-stromal cell interactions in multiple myeloma.
Blood
.
2000
;
95
:
2630
–2636.
103
Kuittinen O, Savolainen E-R, Koistinen P, Mottonen M, Turpeenniemi-Hujanen T. MMP-2 and MMP-9 expression in adult and childhood acute lymphatic leukemia (ALL).
Leukemia Research
.
2001
;
25
:
125
–132.
104
Perez-Atayde AR, Sallan SE, Tedrow U, Connors S, Allred E, Folkman J. Spectrum of tumor angiogenesis in the bone marrow of children with acute lymphoblastic leukemia.
Am J Pathol
.
1997
;
150
:
815
–821.
105
Bertolini F, Paolucci M, Peccatori F, et al. Angiogenic growth factors and Endostatin in non-Hodgkin's lymphoma.
Br J Hematol
.
1999
;
106
:
504
–509.
106
Vacca A, Ribatti D, Ruco L, et al. Angiogenesis extent and macrophage density increase simultaneously with pathological progression in B-cell non-Hodgkin's lymphoma.
Br J Cancer
.
1999
;
79
:
965
–970.
107
Kini AR, Kay NE, Peterson LC. Increased bone marrow angiogenesis in B cell lymphocytic leukemia.
Leukemia
.
2000
;
14
:
1414
–1418.
108
Molica S, Vitelli G, Levato D, Gandolfo GM, Liso V. Increased serum levels of vascular endothelial growth factor predict risk of progression in early B-cell chronic lymphocytic leukaemia.
Br J Heamatol
.
1999
;
107
:
605
–610.
109
Chen H, Treweeke AT, West DC, et al. In vitro and in vivo production of vascular endothelial growth factor by chronic lymphocytic leukemia cells.
Blood
.
2000
;
96
:
3181
–3187.
110
Kim KJ, Li B, Winer J, et al. Inhibition of vascular endothelial growth factor-induced angiogenesis suppresses tumor growth in vivo.
Nature
.
1999
;
362
:
841
–844.
111
Gordon MS, Margolin K, Talpaz M, et al. Phase I safety and pharmacokinetic study of recombinant human ant-vascular endothelial growth factor in patients with advanced cancer.
J Clin Oncology
.
2001
;
19
:
843
–850.
112
Tulple A, Scadden DT, Espina BM, et al. Results of a randomized study of IM862 nasal solution in the treatment of AIDS-related Kaposi's sarcoma.
J Clin Oncol
.
2000
;
18
:
716
–723.
113
Dinney CPN, Bielenberg DR, Perrotte P, et al. Inhibition of basic fibroblast growth factor expression, angiogenesis, and growth of human bladder carcinoma in mice by systemic interferon-κ adminstration.
Cancer Res
.
1998
;
58
:
808
–814.
114
Brekken RA, Overholser JP, Stastny VA, Waltenberger J, Minna JD, Thorpe PE. Selective inhibition of vascular endothelial growth factor (VEGF) receptor 2 (KDR/Flk-1) activity by a monoclonal anti-VEGF antibody blocks tumor growth in mice.
Cancer Res
.
2000
;
60
:
5117
–5124.
115
Prewett M, Huber J, Li Y, et al. Antivascular endothelial growth factor receptor (fetal liver kinase 1) monoclonal antibody inhibits tumor angiogenesis and growth of several mouse and human tumors.
Cancer Res
.
1999
;
59
:
5209
–5218.
116
Sandberg JA, Parker VP, Blanchard KS, et al. Pharmacokinetics and tolerability of an antiangiogenic ribozyme (ANGIOZYME) in healthy volunteers.
J Clin Pharm
.
2000
;
40
:
1462
–1469.
117
Fong TA, Shawver LK, Sun L, et al. SU5416 is a potent and selective inhibitor of the vascular endothelial growth factor receptor (Flk-1/KDR) that inhibits tyrosine kinase catalysis, tumor vascularization, and growth of multiple tumor types.
Cancer Res
.
1999
;
59
:
99
–106.
118
Drevs J, Hofmann I, Hugenschmidt H, et al. Effects of PTK787/ZK 222584, a specific inhibitor of vascular endothelial growth factor receptor tyrosine kinases, on primary tumor, metastasis, vessel density, and blood flow in a murine renal cell carcinoma model.
Cancer Res
.
2000
;
60
:
4819
–4824.
119
Kumar CC, Malkowski M, Yin Z, et al. Inhibition of angiogenesis and tumor growth by SCH221153, a dual αvβ3 and αvβ5 integrin receptor antagonist.
Cancer Res
.
2001
;
61
:
2232
–2238.
120
Kerr JS, Slee MA, Mousa SA. Small molecule av integrin antagonists: novel anticancer agents.
Expert Opin Investig Drugs
.
2000
;
9
:
1271
–1279.
121
Kim Y-M, Jang J-W, Lee O-H, et al. Endostatin inhibits endothelial and tumor cellular invasion by blocking the activation and catalytic activity of matrix metalloproteinase 2.
Cancer Res
.
2000
;
60
:
5410
–5413.
122
O'Reilly MS, Boehm T, Shing Y, et al. Endostatin: an endogenous inhibitor of angiogenesis and tumor growth.
Cell
.
1997
;
88
:
277
–285.
123
Pike SE, Yao L, Setsuda J, et al. Vasostatin, a calreticulin fragment, inhibits angiogenesis and suppresses tumor growth.
J Exp Med
.
1998
;
188
:
2349
–2356.
124
Logothetis CJ, Wu KK, Finn LD, et al. Phase I trial of the angiogenesis inhibitor TNP-470 for progressive androgen-independent prostate cancer.
Clinical Cancer Research
.
2001
;
7
:
1198
–1203.
125
Fotsis T, Zhang Y, Pepper MS, et al. The endogenous oestrogen metabolite 2-methoxyoestradiol inhibits angiogenesis and suppresses tumour growth.
Nature
.
1994
;
368
:
237
–239.
126
Lew YS, Brown SL, Griffin RJ, Song CW, Kim JH. Arsenic trioxide causes selective necrosis in solid murine tumors by vascular shutdown.
Cancer Res
.
1999
;
59
:
6033
–6037.
127
Reichert A, Heisterkamp N, Daley GQ, Groffen J. Treatment of Bcr/Abl-positive acute lymphoblastic leukemia in P190 transgenic mice with the farnesyl transferase inhibitor SCH66336.
Blood
.
2001
;
97
:
1399
–1403.
128
Peters DG, Hoover RR, Gerlach MJ, 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
–1412.
1
Argatoff LH, Connors JM, Klasa RJ, Horsman DE, Gascoyne RD. Mantle cell lymphoma: a clinicopathologic study of 80 cases.
Blood
.
1997
;
89
:
2067
–2078.
2
Rimokh R, Berger F, Delsol G, et al. Rearrangement and overexpression of the BCL1/PRAD1 gene in intermediate lymphocytic lymphomas and in t(11q13)bearing leukemias.
Blood
.
1993
;
81
:
3063
–3067.
3
Greiner TC, Moynihan MJ, Chan WC, et al. p53 mutations in mantle cell lymphomas are associated with variant cytology and predict a poor prognosis.
Blood
.
1996
;
87
:
4302
–4310.
4
Bible KC, Kaurmann SH. Cytotoxic synergy between flavopiridol (NSC 649890, L868275) and various antineoplastic agents: the importance of sequence of administration.
Cancer Res
.
1997
;
57
:
3375
–3380.
5
König A, Schwartz GK, Mohammed RM, Al-Katib A, Gabrilove JL. The novel cyclin-dependent kinase inhibitor flavopiridol downregulates Bcl2 and induces growth arrest and apoptosis in chronic B-cell leukemia cell lines.
Blood
.
1997
;
90
:
4307
–4312.
6
Byrd JC, Shinn CA, Bedi A, et al. Flavopiridol has marked in vitro activity against B-chronic lymphocytic leukemia (BCLL) and induces apoptosis independent of p53 status. Blood. 1997;90(suppl 1):531a (abstr 2366).
7
Byrd JC, Shinn C, Waselenko JK, et al. Flavopiridol induces apoptosis in chronic lymphocytic leukemia cells via activation of caspase3 without evidence of bcl2 modulation or dependence on functional p53.
Blood
.
1998
;
92
:
3804
–3816.
8
Parker BW, Kaur G, Nieves-Niera W, et al. Early induction of apoptosis in hematopoietic cell lines after exposure to flavopiridol.
Blood
.
1998
;
91
:
458
–465.
9
Achenbach TV, Müller R, Slater EP. BCL2 independence of flavopiridolinduced apoptosis.
J Biol Chem
.
2000
;
275
:
32089
–32097.
10
Kitada S, Zapata JM, Andreeff M, Reed JC. Protein kinase inhibitors flavopiridol and 7-hydroxy-staurosporine downregulate antiapoptosis proteins in B cell chronic lymphocytic leukemia.
Blood
.
2000
;
96
:
393
–397.
11
Melillo G, Sausville EA, Cloud K, Lahusen T, Varesio L, Senderowicz AM. Flavopiridol, a protein kinase inhibitor, downregulates hypoxic induction of vascular endothelial growth factor expression in human monocytes.
Cancer Res
.
1999
;
59
:
5433
–5437.
12
Arguello F, Alexander M, Sterry JA, et al. Flavopiridol induces apoptosis of normal lymphoid cells, causes immunosuppression, and has potent antitumor activity in vivo against human leukemia and lymphoma xenografts.
Blood
.
1998
;
91
:
2482
–2490.
13
Innocenti F, Stadler W, Iyer L, Ramírez J, Vokes EE, Ratain MJ. Flavopiridol metabolism in cancer patients is associated with the occurrence of diarrhea.
Clin Cancer Res
.
2000
;
6
:
3400
–3405.
14
Maki C, Huibregtse J, Howley P. In vivo ubiquination and proteasome-mediated degradation of p53.
Cancer Res
.
1996
;
56
:
2649
–2654.
15
Yoshida H, Kitamura K, Tanaka K, et al. Accelerated degradation of PML-retinoic acid receptor a (PML RARα by all trans retinoic acid in acute promyelocytic leukemia: possible role of the proteasome pathway.
Cancer Res
.
1996
;
56
:
2945
–2948.
16
Adams J, Stein R. Novel inhibitors of the proteasome and their therapeutic use in inflammation.
Ann Rep Med Chem
.
1996
;
31
:
279
–288.
17
Adams J, Palombella VJ, Sausville EA, et al. Proteasome inhibitors: a novel class of potent and effective antitumor agents.
Cancer Res
.
1999
;
59
:
2615
–2622.
18
Beg AA, Baltimore D. An essential role for NF-κB in preventing TNFa induced cell death.
Science
.
1996
;
274
:
782
–784.
19
Wang CVY, Cusack JC, Liu R. Control of inducible chemoresistance: enhanced antitumor therapy through increased apoptosis by inhibition of NFκB.
Nature
.
1999
;
5
:
412
–417.
20
Lin ZP, Boller YC, Amer SM, et al. Prevention of brefeldin-induced resistance to teniposide by the proteasome inhibitor MG132: involvement of NF-κB activation in drug resistance.
Cancer Res
.
1998
;
58
:
3059
–3065.
21
Soglio D, Servida F, Delia D, et al. The apoptogenic response of human myeloid leukaemia cells and of normal and malignant haematopoietic progenitor cells to the proteasome inhibitor PSI.
Br J Haematol
.
2001
;
113
:
126
–136.
22
Orlowski RZ, Eswara JR, Larond-Walker A, Grever MR, Orlowski M, Dang CV. Tumor growth inhibition induced in a murine model of human Burkitt's lymphoma by a proteasome inhibitor. Cancer Res. 1998;58:4342-4348,
23
Hideshima T, Richardson P, Chauhan D, et al. The proteasome 6-inhibitor PS-341 inhibits growth, induces apoptosis, and overcomes drug resistance in human multiple myeloma cells.
Cancer Res
.
2001
;
61
:
3071
–3076.
24
Chiarle R, Budel LM, Skolnik J, et al. Increased proteasome degradation of cyclin-dependent kinase inhibitor p27 is associated with a decreased overall survival in mantle cell lymphoma.
Blood
.
2000
;
95
:
619
–626.
25
Masdehors P, Omura S, MerleBeral H, et al. Increased sensitivity of CLL-derived lymphocytes to apoptotic death activation by the proteasime-specific inhibitor lactacystin.
Br J Haemtaol
.
1999
;
105
:
752
–757.
26
Delic J, Masdehors P, Omura S, et al. The proteasome inhibitor lactacystin induces apoptosis and sensitizes chemo and radioresistant human chronic lymphocytic leukaemia lymphocytes to TNF α-initiated apoptosis.
Br J Cancer
.
1999
;
77
:
375
–376.
27
Delic J, Masdehors P, Omura S, et al. The proteasome inhibitor lactacystin induces apoptosis and sensitizes chemo- and radioresistant human CLL lymphocytes to TNF α-induced apoptosis.
Br J Cancer
.
1998
;
77
:
1103
–1107.
28
Masdehors P, MerleVBeral H, Magdelenat H, Delic J. Ubiquitin proteasome system and increased susceptibility of BCLL lymphocytes to apoptotic death activation.
Leuk Lymphoma
.
2000
;
38
(56):
499
–504.
29
Chandra J, Niemer I, Gilbreath J, et al. Proteasome inhibitors induce apoptosis in glucocorticoid resistant chronic lymphocytic leukemia lymphocytes.
Blood
.
1998
;
92
:
42204229
.
30
Stinchcombe TE, Mitchell BS, Depcik V, et al. PS341 is active in multiple myeloma: preliminary report of a phase I trial of the proteasome inhibitor PS-341 in patients with hematologic malignancies. Blood. 2000;96(suppl 1):516a (abstr 2219).
31
Nix D, Pien C, Newman R, et al. Clinical development of a proteasome inhibitor, PS341, for the treatment of cancer. Proc ASCO. 2001;20:86 (abstr 339)
32
Cheson BD, Horning SJ, Coiffier B, et al. Report of an International Workshop to standardize response criteria for non-Hodgkin's lymphomas.
Clin Oncol
.
1999
;
17
:
1244
–1253.