Mouse models that recapitulate human malignancy are valuable tools for the elucidation of the underlying pathogenetic mechanisms and for preclinical studies. Several genetically engineered mouse models have been generated, either mimicking genetic aberrations or deregulated gene expression in chronic lymphocytic leukemia (CLL). The usefulness of such models in the study of the human disease may potentially be hampered by species-specific biological differences in the target cell of the oncogenic transformation. Specifically, do the genetic lesions or the deregulated expression of leukemia-associated genes faithfully recapitulate the spectrum of lymphoproliferations in humans? Do the CLL-like lymphoproliferations in the mouse have the phenotypic, histological, genetic, and clinical features of the human disease? Here we compare the various CLL mouse models with regard to disease phenotype, penetrance, and severity. We discuss similarities and differences of the murine lymphoproliferations compared with human CLL. We propose that the Eμ-TCL1 transgenic and 13q14-deletion models that have been comprehensively studied at the levels of leukemia phenotype, antigen-receptor repertoire, and disease course show close resemblance to the human disease. We conclude that modeling CLL-associated genetic dysregulations in mice can provide important insights into the molecular mechanisms of disease pathogenesis and generate valuable tools for the development of novel therapies.

Chronic lymphocytic leukemia (CLL) originates from the clonal expansion of mature B cells, which show features of antigenic stimulation and express the CD5 cell surface antigen. CLL is a complex disease in which genetic abnormalities cooperate with microenvironmental factors in the malignant transformation of the tumor-cell precursor and in leukemia progression.1  Compared with most other subtypes of non-Hodgkin lymphoma (NHL), CLL shows a lower frequency of genetic mutations per case and a different spectrum of genetic aberrations, which mostly comprise chromosomal deletions (13q14, ATM, and TP53) or amplifications (trisomy of chromosome 12).2,3  More recently, next-generation sequencing (NGS) analyses have identified several novel recurrent mutations in CLL, including those that target the NOTCH1, MYD88, SF3B1, and BIRC3 genes.4-11  A number of genes are overexpressed in CLL tumor cells compared with normal lymphocytes, presumably as a direct consequence of the genetic aberrations (eg, BCL2 and MCL1 due to deletion of mir-15a/16-112 ) or through as yet unknown mechanisms (eg, ROR113,14  or TCL115 ). Finally, genome-wide association studies have identified several susceptibility loci for familial CLL,16  including a single nucleotide polymorphism in the IRF4 gene, a known regulator of B-cell developmental processes.17 

Several mouse models mimicking genetic lesions found in CLL (13q14 deletion), transgenic for genes that are overexpressed in the disease (including TCL1, APRIL, BCL2 × traf2dn, ROR1), or driven by ectopic oncogene expression (IgH.T and IgH.TEμ) have been generated (Table 1).

Table 1

Mouse models of CLL

Mouse modelKey findings
mir-15a/16-1−/− and mir-15a/16-1floxedCD19-Cre mice26  Germ-line mutations interfering with the normal expression of mir-15a/16-1 are observed in a small fraction of CLL patients.115  The first indication that a disruption of the physiological expression of mir-15a/16-1 favors CLL development stemmed from the analysis of the NZB strain which is characterized by a germ-line mutation in the 3′ flanking region of pre-mir-16-1, resulting in decreased expression of mature mir-16-1.116  The targeted genetic inactivation of mir-15a/16-1 in mice provided conclusive evidence for a tumor-suppressor role of these microRNAs in CLL development,26  as proposed.25  25-30% of mir-15a/16-1−/− and mir-15a/16fl/–CD19-Cre mice developed, late in life, MBL and CLL, and less frequently CD5-negative NHLs in a B-cell autonomous fashion.26  The loss of the mir-15a/16-1 cluster led to an earlier entry into the cell cycle and an up-regulation of BCL2 protein levels compared with wild-type B cells,26  as previously described.117  
14qC3 minimal deleted region (MDR)−/− and MDRfloxedCD19-Cre mice26  MDR-deleted mice that in addition to mir-15a/16-1 lack the dleu2 and dleu5 genes show a higher penetrance of the phenotype and a more aggressive disease course compared with mir-15a/16-1-deleted mice,26  although the pathogenetic mechanism remains to be elucidated. 40-45% of MDR−/− and MDRfl/–CD19-Cre mice developed MBL, CLL and CD5-negative NHLs in a B-cell autonomous fashion. MDR−/− mice, in contrast to mir-15a/16-1−/− mice, succumbed to the lymphoproliferations earlier than their wild-type littermates. 
14qC3 common deleted region (CDR)floxedCD19-Cre mice27  Homozygous deletion of the CDR in the germ-line causes embryonic lethality.27  Mice with deletion of the CDR in B cells develop lymphoproliferations at a similar frequency (40-45%) as MDR-deleted mice.27  However, the mice are characterized by a different spectrum of lymphoproliferations, as they mostly develop CLL with rare instances of MBL and NHLs. In addition, CDR-deleted mice presented with tumor cell infiltrates in spleen, bone marrow and non-lymphoid organs that were generally larger than those observed in the mir-15a/16-1 and MDR-deleted mice presenting with CLL. CDR+/− mice developed mainly CD5-positive lymphoproliferations at a penetrance of about 25%,27  which is similar to that observed in MDR+/− mice.26  While both CDR+/− and MDR+/− mice have a similar disease onset, once the lymphoproliferations develop, deletion of the CDR leads to a more aggressive disease.27  DLEU7 has been suggested to be a negative regulator of NF-κB activity.118  
Eμ-TCL1 transgenic mice28,29  Exogenous expression of the human TCL1 gene under the control of the IGHV promoter and IGH enhancer () in vivo (Eμ-TCL1) results in the clonal expansion of CD5+IgM+ B cells.28,29  Between 13 and 18 mo of age, virtually all -TCL1 mice develop an overt leukemia and massive infiltrations of monoclonal CD5+ B cells in both lymphoid and non-lymphoid tissues. TCL1 is a co-activator of the serine/threonine kinase AKT,38,39  activates the NF-κB pathway in CLL cells,40  and inhibits DNMT3A and DNMT3B activity.41  Leukemia development is at least partially dependent on enhanced AKT activity.90  
APRIL transgenic mice31,32  Based on the finding of elevated levels of tumor necrosis factor (TNF) family member APRIL in sera of CLL patients,119 APRIL transgenic mice have been generated that accumulate increased levels of the molecule in the sera.31,32  APRIL induces proliferation of B cells. APRIL transgenic mice develop clonal lymphoproliferations originating from peritoneal CD5+ B1 cells between 9 and 12 mo of age with a penetrance of 40%.32  
BCL2 × traf2dn transgenic mice42  Mice transgenic for the anti-apoptotic gene BCL2,120  which is expressed at high levels in human CLL cells, and a dominant negative form of the adaptor protein TNF receptor-associated factor 2 (traf2dn), the latter being structurally similar to TRAF1 which is overexpressed in CLL,33  have been generated to study the synergistic effects of these molecules in CLL pathogenesis.42  By 14 mo, ∼80% of BCL2 × traf2dn double-transgenic mice had died of a CLL-like disease showing lymphocytosis, lymphoadenopathy, spleen enlargement and involvement of bone marrow and non-lymphoid organs. 
ROR1 transgenic mice34  Receptor tyrosine kinase-like orphan receptor 1 (ROR1) is highly expressed on CLL cells in humans.13,14  This oncoembryonic antigen is considered a potential target for CLL therapy as it is virtually absent from adult tissues. Transgenic mice with the human ROR1 gene which is controlled by the murine Ig enhancer/promoter to ascertain B cell-restricted expression develop, late in life (≥ 15 mo), clonal lymphoproliferations resembling CLL at a very low penetrance (∼5%).34  However, ROR1 x TCL1 double-transgenic mice succumbed to CLL more rapidly than single Eμ-TCL1 or ROR1-tg mice,34  demonstrating that ROR1 expression accelerates disease progression in Eμ-TCL1 mice. 
Eμ-mir-29 transgenic mice35  Eμ-mir-29 transgenic mice that overexpress the microRNA cluster miR-29a/b in B cells present with expanded CD5+ B-cell populations and late in life develop an indolent CLL-like leukemia at a penetrance of ∼20%.35  Overexpression of miR-29 in myeloid cells promotes leukemogenesis.121  By analogy, it is possible that miR-29 may have oncogenic functions in the precursor cells of murine CLL. However, since in human CLL, miR-29 expression is downregulated and thought to exert tumor-suppressor functions,122  the implications of the findings from the Eμ-mir-29 model regarding a role of this microRNA cluster in human CLL development are presently unclear.123  
Vh11 × irf4−/− mice36  To elucidate the proposed role of reduced IRF4 expression in CLL development16  in a mouse model, IRF4-deficient mice were crossed to mice transgenic for the Vh11 heavy chain gene which develop expansions of peritoneal CD5+ B cells.36 Vh11 × irf4−/− mice presents with a CLL-like disease in the majority of mice by 12 mo. Since irf4−/− mice are severely immunodeficient due to critical functions of IRF4 in various immune cell types,17  the conclusive determination of a B cell-intrinsic role of IRF4 deficiency in CLL pathogenesis needs to await results from Vh11 × irf4 conditional mice crossed to a B cell-specific deletor mouse. With regard to a possible leukemogenic function of altered IRF4 levels, a recent study demonstrated that reduced IRF4 expression altered the migration properties of B cells, most likely by upregulating NOTCH2 activity.43  
IgH.T and IgH.TEμ mice37  The SV40 T antigen was suggested to exert an oncogenic function in B-cell malignancies.124  Sporadic SV40 T antigen expression in mature B cells has been achieved by insertion of a SV40 T antigen gene in opposite transcriptional orientation in the IGH chain locus between the D and JH segments, in presence (IgH.TEμ) or absence (IgH.T) of an extra copy of the enhancer. Virtually all aging IgH.TEμ and 13% of IgH.T mice developed an expansion of CD5+ B cells carrying either unmutated IGHV genes with preferential usage of the Vh11 family, or highly mutated IGHV genes. 
Mouse modelKey findings
mir-15a/16-1−/− and mir-15a/16-1floxedCD19-Cre mice26  Germ-line mutations interfering with the normal expression of mir-15a/16-1 are observed in a small fraction of CLL patients.115  The first indication that a disruption of the physiological expression of mir-15a/16-1 favors CLL development stemmed from the analysis of the NZB strain which is characterized by a germ-line mutation in the 3′ flanking region of pre-mir-16-1, resulting in decreased expression of mature mir-16-1.116  The targeted genetic inactivation of mir-15a/16-1 in mice provided conclusive evidence for a tumor-suppressor role of these microRNAs in CLL development,26  as proposed.25  25-30% of mir-15a/16-1−/− and mir-15a/16fl/–CD19-Cre mice developed, late in life, MBL and CLL, and less frequently CD5-negative NHLs in a B-cell autonomous fashion.26  The loss of the mir-15a/16-1 cluster led to an earlier entry into the cell cycle and an up-regulation of BCL2 protein levels compared with wild-type B cells,26  as previously described.117  
14qC3 minimal deleted region (MDR)−/− and MDRfloxedCD19-Cre mice26  MDR-deleted mice that in addition to mir-15a/16-1 lack the dleu2 and dleu5 genes show a higher penetrance of the phenotype and a more aggressive disease course compared with mir-15a/16-1-deleted mice,26  although the pathogenetic mechanism remains to be elucidated. 40-45% of MDR−/− and MDRfl/–CD19-Cre mice developed MBL, CLL and CD5-negative NHLs in a B-cell autonomous fashion. MDR−/− mice, in contrast to mir-15a/16-1−/− mice, succumbed to the lymphoproliferations earlier than their wild-type littermates. 
14qC3 common deleted region (CDR)floxedCD19-Cre mice27  Homozygous deletion of the CDR in the germ-line causes embryonic lethality.27  Mice with deletion of the CDR in B cells develop lymphoproliferations at a similar frequency (40-45%) as MDR-deleted mice.27  However, the mice are characterized by a different spectrum of lymphoproliferations, as they mostly develop CLL with rare instances of MBL and NHLs. In addition, CDR-deleted mice presented with tumor cell infiltrates in spleen, bone marrow and non-lymphoid organs that were generally larger than those observed in the mir-15a/16-1 and MDR-deleted mice presenting with CLL. CDR+/− mice developed mainly CD5-positive lymphoproliferations at a penetrance of about 25%,27  which is similar to that observed in MDR+/− mice.26  While both CDR+/− and MDR+/− mice have a similar disease onset, once the lymphoproliferations develop, deletion of the CDR leads to a more aggressive disease.27  DLEU7 has been suggested to be a negative regulator of NF-κB activity.118  
Eμ-TCL1 transgenic mice28,29  Exogenous expression of the human TCL1 gene under the control of the IGHV promoter and IGH enhancer () in vivo (Eμ-TCL1) results in the clonal expansion of CD5+IgM+ B cells.28,29  Between 13 and 18 mo of age, virtually all -TCL1 mice develop an overt leukemia and massive infiltrations of monoclonal CD5+ B cells in both lymphoid and non-lymphoid tissues. TCL1 is a co-activator of the serine/threonine kinase AKT,38,39  activates the NF-κB pathway in CLL cells,40  and inhibits DNMT3A and DNMT3B activity.41  Leukemia development is at least partially dependent on enhanced AKT activity.90  
APRIL transgenic mice31,32  Based on the finding of elevated levels of tumor necrosis factor (TNF) family member APRIL in sera of CLL patients,119 APRIL transgenic mice have been generated that accumulate increased levels of the molecule in the sera.31,32  APRIL induces proliferation of B cells. APRIL transgenic mice develop clonal lymphoproliferations originating from peritoneal CD5+ B1 cells between 9 and 12 mo of age with a penetrance of 40%.32  
BCL2 × traf2dn transgenic mice42  Mice transgenic for the anti-apoptotic gene BCL2,120  which is expressed at high levels in human CLL cells, and a dominant negative form of the adaptor protein TNF receptor-associated factor 2 (traf2dn), the latter being structurally similar to TRAF1 which is overexpressed in CLL,33  have been generated to study the synergistic effects of these molecules in CLL pathogenesis.42  By 14 mo, ∼80% of BCL2 × traf2dn double-transgenic mice had died of a CLL-like disease showing lymphocytosis, lymphoadenopathy, spleen enlargement and involvement of bone marrow and non-lymphoid organs. 
ROR1 transgenic mice34  Receptor tyrosine kinase-like orphan receptor 1 (ROR1) is highly expressed on CLL cells in humans.13,14  This oncoembryonic antigen is considered a potential target for CLL therapy as it is virtually absent from adult tissues. Transgenic mice with the human ROR1 gene which is controlled by the murine Ig enhancer/promoter to ascertain B cell-restricted expression develop, late in life (≥ 15 mo), clonal lymphoproliferations resembling CLL at a very low penetrance (∼5%).34  However, ROR1 x TCL1 double-transgenic mice succumbed to CLL more rapidly than single Eμ-TCL1 or ROR1-tg mice,34  demonstrating that ROR1 expression accelerates disease progression in Eμ-TCL1 mice. 
Eμ-mir-29 transgenic mice35  Eμ-mir-29 transgenic mice that overexpress the microRNA cluster miR-29a/b in B cells present with expanded CD5+ B-cell populations and late in life develop an indolent CLL-like leukemia at a penetrance of ∼20%.35  Overexpression of miR-29 in myeloid cells promotes leukemogenesis.121  By analogy, it is possible that miR-29 may have oncogenic functions in the precursor cells of murine CLL. However, since in human CLL, miR-29 expression is downregulated and thought to exert tumor-suppressor functions,122  the implications of the findings from the Eμ-mir-29 model regarding a role of this microRNA cluster in human CLL development are presently unclear.123  
Vh11 × irf4−/− mice36  To elucidate the proposed role of reduced IRF4 expression in CLL development16  in a mouse model, IRF4-deficient mice were crossed to mice transgenic for the Vh11 heavy chain gene which develop expansions of peritoneal CD5+ B cells.36 Vh11 × irf4−/− mice presents with a CLL-like disease in the majority of mice by 12 mo. Since irf4−/− mice are severely immunodeficient due to critical functions of IRF4 in various immune cell types,17  the conclusive determination of a B cell-intrinsic role of IRF4 deficiency in CLL pathogenesis needs to await results from Vh11 × irf4 conditional mice crossed to a B cell-specific deletor mouse. With regard to a possible leukemogenic function of altered IRF4 levels, a recent study demonstrated that reduced IRF4 expression altered the migration properties of B cells, most likely by upregulating NOTCH2 activity.43  
IgH.T and IgH.TEμ mice37  The SV40 T antigen was suggested to exert an oncogenic function in B-cell malignancies.124  Sporadic SV40 T antigen expression in mature B cells has been achieved by insertion of a SV40 T antigen gene in opposite transcriptional orientation in the IGH chain locus between the D and JH segments, in presence (IgH.TEμ) or absence (IgH.T) of an extra copy of the enhancer. Virtually all aging IgH.TEμ and 13% of IgH.T mice developed an expansion of CD5+ B cells carrying either unmutated IGHV genes with preferential usage of the Vh11 family, or highly mutated IGHV genes. 

AKT, protein kinase B/v-AKT murine thymoma viral oncogene; DNMT, DNA methyltransferase; NF-κB, nuclear factor κB; SV40, simian virus 40; TNF, tumor necrosis factor.

Mouse models mimicking the spectrum of deletions of chromosomal region 13q14

Deletion of 13q14 is the most frequent genetic lesion in CLL that occurs in >50% of cases and is clinically associated with an indolent disease with no or delayed need for therapy.3,18  The deletion is detected also in monoclonal B-cell lymphocytosis (MBL),19  an expansion of CD5+ B lymphocytes in the peripheral blood (PB) of otherwise healthy individuals19,20  that is thought to precede CLL,21  and is present at a lower frequency in other NHL subtypes.22 

The 13q14 region corresponds to the murine 14qC3 locus and encodes several genes that are highly conserved among human and mouse (Figure 1). A minimal deleted region (MDR) includes the first exon of the DLEU1 sterile RNA, the deleted in leukemia (DLEU) 2 gene, encoding a sterile transcript and the miR-15a/16-1 cluster, located in an intron of DLEU2.23-25  A larger 13q14 deletion is found in a sizable number of CLL cases and is hence named the common deleted region (CDR). The functional dissection of the 13q14 tumor suppressor locus by using transgenic mouse models demonstrated that the size of 13q14 deletions influences the phenotype of the developing lymphoproliferations and the severity of disease (Table 1),26,27  suggesting a tumor suppressor function for multiple genetic elements encoded in the deleted region. Specifically, the results provided evidence for a causal role of the microRNAs mir-15a/16-1 in CLL pathogenesis and demonstrated that the additional deletions of dleu2 and/or dleu5 caused a higher disease penetrance and a more severe disease course.26  In both models, CD5-negative NHLs developed along with CD5-positive lymphoproliferations that encompassed MBL and CLL. Importantly, CDR-deleted mice were characterized by a different spectrum of lymphoproliferations compared with MDR and mir-15a/16-1–deleted mice, as they mostly developed CLL (rarely MBL or NHLs), and a more aggressive disease course.27  A commonality among all 13q14 mouse models was the expression of unmutated stereotypic antigen receptors,26,27  indicating a role of antigen in the expansion of the CLL clone.

Figure 1

Human 13q14 and murine 14qC3 locus. Indicated are the MDR and the CDR.

Figure 1

Human 13q14 and murine 14qC3 locus. Indicated are the MDR and the CDR.

Close modal

Mouse models mimicking the deregulated expression of genes in human CLL

Several mouse models have been generated that are transgenic for genes overexpressed in the disease (Table 1). The first transgenic mouse that developed a CLL-like disease and that since has been widely used in the study of CLL pathogenesis and therapy is the Eμ-TCL1 transgenic mouse.28,29  The TCL1 gene is expressed in the vast majority of human CLL cases.15  Strong TCL1 expression is associated with markers of poor prognosis, including unmutated IGHV genes.15 Eμ-TCL1 mice developed a clonal expansion of CD5+IgM+ B cells initially in the peritoneal cavity and on progression in the PB, spleen, and bone marrow,28,29  eventually giving rise to a monoclonal B-cell population in virtually 100% of mice (Table 1).28  The B-cell receptors (BCRs) expressed by lymphoproliferations in Eμ-TCL1 mice harbored unmutated IGHV gene rearrangements and exhibited stereotypy in IGHV, IGKV, and IGLV gene rearrangements.30 

The characteristics of the other mouse models that develop a CLL-like disease, including APRIL transgenic mice,31,32 BCL2 × traf2dn transgenic mice,33 ROR1 transgenic mice,34 Eμ-mir-29 transgenic mice,35 Vh11 × irf4−/− mice,36  and IgH.T and IgH.TEμ mice37  are summarized in Table 1.

Commonalities and disparities among CLL mouse models

In all mouse models, the lymphoproliferations develop late in life and resemble the indolent disease course of CLL. Also, the target cell of the malignant transformation to CLL seems to be the peritoneal B1a cell, as suggested by the expression of CD5 and of unmutated IGHV genes, high levels of IgM, and low levels of IgD and CD23, at least in the models where these parameters were analyzed. Noticeably, whereas virtually all Eμ-TCL1 mice developed CLL, a sizable fraction of B-cell lymphoproliferations in the 13q14 deletion models represented CD5-negative NHLs. The latter observation genuinely reflects the heterogeneous spectrum of B-cell malignancies with the 13q14 deletion occurring in humans.

The most notable difference among the CLL mouse models is the penetrance of the phenotype, which is highest in the Eμ-TCL1 mice (∼100%), intermediate in the 13q14-MDR and CDR deletion models and the APRIL transgenic mice (40-50%), and lowest in the mir-15/16-1 deletion model (∼25%) and the ROR1 transgenic mice (5%) (Table 2). The different tumor incidence among the CLL mouse models suggests that the various genetic aberrations predispose to, but are not sufficient per se, for the malignant transformation of B cells, a situation similar to that encountered by 13q deletions in low-count MBL, which virtually never progresses into frank leukemia.20  Thus, additional alterations are required to induce leukemia. The heterogeneity of the lymphoproliferation phenotypes observed in mir-15a/16-1−/− and MDR−/− mice (CLL, MBL, and CD5 NHL) may reflect the specific nature of the secondary hits and/or the different B-cell differentiation stages in which the malignant transformation occurs. In particular, the observation that CDR-deleted mice predominantly develop a CLL-like disease27  suggests the existence of tumor suppressor genes in the h13q14/m14qC3 locus in addition to the DLEU2/miR-15a/16-1 cluster. The almost complete disease penetrance in the Eμ-TCL1 model is likely due to a strong oncogenic function of the hTCL1 gene, which affects multiple pathways, including coactivation of the serine/threonine kinase protein kinase B/v-AKT murine thymoma viral oncogene (AKT),38,39  activation of the nuclear factor κB pathway,40  and inhibition of the activity of DNA methyltransferase (DNMT)3A and DNMT3B.41  In the APRIL transgenic model, the sustained microenvironmental stimulation likely facilitates the expansion of B1a cells, which transform in 40% of cases.32  Similarly, prosurvival signals delivered through the ROR1 receptor may favor the expansion of a preleukemic B-cell population that gives rise to a CLL-like disease in 5% of animals.34  Conversely, an increased proliferative capacity seems to account for leukemia development in 20% of Eμ-mir-29 transgenic mice.35  Overexpression of BCL2 synergizes with a dominant negative form of the adaptor protein tumor necrosis factor (TNF) receptor-associated factor 2 (traf2dn) in the malignant transformation of B cells, resulting in the development of a CLL-like disease in ∼80% of BCL2 × traf2dn transgenic mice, most likely due to resistance to apoptosis and altered expression of adhesion molecules.42  Apoptosis resistance has also been suggested to account for leukemia development in the Vh11 × irf4−/− model,36  although a recent study suggests that the possible leukemogenic function of reduced IRF4 expression may be the alteration of the migration properties of B cells,43  presumably through up-regulation of NOTCH2 activity in IRF4-deficient cells.43  Finally, expression of the simian virus 40 (SV40) T antigen in B cells has been associated with the induction of genomic instability and deregulation of genes involved in proliferation, DNA repair, and apoptosis.37  The almost complete disease penetrance in the latter 2 mouse models is likely to be due to a skewed IGHV gene repertoire (the Vh11 × irf4−/−) and the ectopic expression of a potent oncogene (SV40 T antigen), respectively, thus introducing exogenous factors that accelerate B-cell transformation. Therefore, among the models that mimic genetic dysregulations found in human CLL, overexpression of the hTCL1 gene in mouse B cells accounts for the highest disease penetrance.

Table 2

Features of genetically engineered mouse models of CLL

Mouse modelDisease penetranceTime of appearance of circulating leukemic cellsAge of deathIG gene rearrangements
mir-15a/16-1−/− and mir-15a/16-1floxed CD19-Cre ∼20% CLL 12-18 mo 15-18 mo Unmutated and stereotypic IGHV genes 
∼8% MBL 
∼2% CD5 NHL 
14qC3-MDR−/− and MDRfloxedCD19-Cre ∼22% CLL 6-18 mo 12-18 mo Unmutated and stereotypic IGHV genes 
∼12% MBL 
∼6% CD5 NHL 
14qC3-CDRfloxedCD19-Cre ∼50% CLL 6-18 mo 12-18 mo Unmutated and stereotypic IGHV genes 
∼3% MBL 
Eμ-TCL1 tg 100% CLL 6 mo 12-18 mo Unmutated IGHV genes, stereotypic IGHV and IGLV genes 
APRIL tg 40% CLL Not analyzed 12 to >15 mo Not analyzed 
BCL2 × traf2dn tg 80% CLL 9-15 mo >80% dead at 14 mo Clonal IGHV rearrangements 
ROR1 tg 5% CLL >15 mo >15 mo Clonal IGHV rearrangements 
Eμ-mir-29 tg 20% CLL 12-24 mo 24-26 mo Clonal IGHV rearrangements 
Vh11 × irf4−/− 100% CLL (preceded by MBL in >40% cases) 5-10 mo >9 mo Clonal IGHV rearrangements (determined by FACS) 
IgH.TEμ 100% CLL <5 mo Killed at 2-10 mo Unmutated IGHV genes (preferentially Vh11) and some highly mutated 
Mouse modelDisease penetranceTime of appearance of circulating leukemic cellsAge of deathIG gene rearrangements
mir-15a/16-1−/− and mir-15a/16-1floxed CD19-Cre ∼20% CLL 12-18 mo 15-18 mo Unmutated and stereotypic IGHV genes 
∼8% MBL 
∼2% CD5 NHL 
14qC3-MDR−/− and MDRfloxedCD19-Cre ∼22% CLL 6-18 mo 12-18 mo Unmutated and stereotypic IGHV genes 
∼12% MBL 
∼6% CD5 NHL 
14qC3-CDRfloxedCD19-Cre ∼50% CLL 6-18 mo 12-18 mo Unmutated and stereotypic IGHV genes 
∼3% MBL 
Eμ-TCL1 tg 100% CLL 6 mo 12-18 mo Unmutated IGHV genes, stereotypic IGHV and IGLV genes 
APRIL tg 40% CLL Not analyzed 12 to >15 mo Not analyzed 
BCL2 × traf2dn tg 80% CLL 9-15 mo >80% dead at 14 mo Clonal IGHV rearrangements 
ROR1 tg 5% CLL >15 mo >15 mo Clonal IGHV rearrangements 
Eμ-mir-29 tg 20% CLL 12-24 mo 24-26 mo Clonal IGHV rearrangements 
Vh11 × irf4−/− 100% CLL (preceded by MBL in >40% cases) 5-10 mo >9 mo Clonal IGHV rearrangements (determined by FACS) 
IgH.TEμ 100% CLL <5 mo Killed at 2-10 mo Unmutated IGHV genes (preferentially Vh11) and some highly mutated 

There is also the perception that the disease developing in Eμ-TCL1 mice is more aggressive compared with the one in the 13q14 deletion models. Indeed, based on the occurrence of unmutated BCRs in Eμ-TCL1 leukemic mice that resemble those in human CLL and are structurally similar to autoantibodies or antibodies recognizing microbial antigens, it was proposed that CLL developing in those mice correspond to the aggressive, treatment-resistant subtype in humans. However, the IGHV genes of the leukemic clones developing in Eμ-TCL1 and 13q14 deletion models were surprisingly similar. A sizable number of rearrangements obtained from leukemias in both of these models could be assigned to 5 stereotype sets defined by Chiorazzi et al.30  Even more intriguing, CLLs with identical IGHV-IGHD-IGHJ junctions were found among the Eμ-TCL1 and 13q14-deleted mouse models.27  These observations suggest that the same antigens are involved in the transformation and progression of B-cell clones developing in Eμ-TCL1 transgenic and 13q14-deleted animals. Moreover, these models show a similar disease phenotype that is characterized by the infiltration of the leukemic cells in lymphoid and nonlymphoid organs. Together with evidence that supports a common cell of origin (see below), this argues against a difference in disease aggressiveness among the Eμ-TCL1 transgenic and 13q14-deleted models. The shorter latency of the appearance of a B1a cell expansion (4-5 months in Eμ-TCL1 mice vs 9-12 months in the 13q14 deletion models) and higher penetrance of the phenotype suggest that the hTCL1 oncogene causes the expansion of a larger pool of CLL precursor cells that are targets for additional transforming events in Eμ-TCL1 compared with 13q14-deleted mice. Interestingly, relative to the mir-15a/16-1−/− and MDR−/− models, there is no obvious MBL stage in Eμ-TCL1 transgenic and CDR-deleted mice that precedes CLL development, which may be indicative of a faster disease progression. However, it seems that after a malignant clone has appeared, Eμ-TCL1 and 13q14-deleted animals show the same disease course.

The usefulness of CLL mouse models for the study of the human disease depends on the grade of relatedness of the disease characteristics between animal models and patients. There are 2 issues to consider. First, with respect to the involvement of the genetic lesions or the deregulated expression of leukemia-associated genes, do the models faithfully recapitulate the spectrum of lymphoproliferations observed in the human disease? Second, do the CLL-like lymphoproliferations in the mouse have the phenotypic, histological, genetic, and clinical features of the disease developing in humans? Evidently, a major factor in this is the phenotypic and functional relatedness of the cells targeted by the oncogenic transformation in humans and mice. In the following, we will discuss the commonalities among CLL in mice and humans with respect to disease origin, repertoire of antigen receptors, and phenotypic and epigenetic characteristics of the tumor cells.

Disease origin

Possibly all CD5-positive lymphoproliferations developing in CLL mouse models derive from the transformation of peritoneal, self-replenishing CD5+ B1a cells.44  Clonal populations of B1a cells develop spontaneously within the peritoneal cavity of aging mice and eventually disseminate to and expand in other lymphoid organs.45,46  This suggests that over time, B1a cells become susceptible to malignant transformation as the results of genetic abnormalities that accumulate during their self-replenishment, eventually giving rise to a CLL-like disease. Protumorigenic forces such as deletion of 13q14, overexpression of TCL1, or high serum levels of APRIL that confer a proliferation and/or survival advantage may accelerate the disease onset, increase the penetrance of the phenotype, and impact disease course and severity (Table 2).

What is the relationship between the CLL precursor cells in the human and mouse? The longstanding discussion about the putative cell of origin of CLL in humans has recently been revisited by Chiorazzi and Ferrarini,47  who provide conceivable arguments for the not mutually exclusive possibilities that the oncogenic transformation may occur in a marginal zone B cell, a transitional B cell, or a human B1-like cell,47  and independent evidence suggests that CLL cases with unmutated IGHV genes may originate from transformed naïve B cells.48  The notion of a B1a-like cell of origin of CLL recently gained traction through the identification of a previously unrecognized CD5-expressing B-cell population in the adult.49  These cells comprise only about 1% of PB B cells, but their gene expression profile is closely related to that of CLL cells.49  One notable aspect of this new hypothesis is that it would be consistent with the fact that the CD5 antigen is invariably expressed on human and murine CLL cases. However, because the function of these newly identified CD5+ B cells is unknown, it is presently impossible to draw a functional relationship between these cells and murine B1a cells.

Stereotypic antigen receptors

A recent study provided a fascinating example that the expansion of the CLL clone can be driven by antigen-independent cell autonomous signaling.50  Nevertheless, there is clear evidence for the role of antigen-BCR stimulation via autoantigens or external antigens in CLL pathogenesis.51-54  A characteristic of CLL is the expression of stereotypic antigen receptors among unrelated CLL cases that show highly similar heavy-chain complementarity-determining (HCDR3) regions,55  presumably as the result of strong selection forces because the normal HCDR3 repertoire is extremely diverse due to IGHV-IGHD-IGHJ combinatorial and junctional diversity. Sequence analyses of antigen-receptor rearrangements from Eμ-TCL1 transgenic and 13q14 deletion models demonstrated a skewed IGHV-IGHD-IGHJ rearrangement repertoire in the leukemic clones derived from these mice,26,27,30  which is reflected by the expression of highly similar HCDR3 regions that are reminiscent of human stereotypic antigen receptors. Specifically, 44% of the sequences from the -TCL1, miR-15a/16-1, MDR, and CDR cohorts could be assigned to 8 sets of stereotypic HCDR3 regions, which are defined by ≥80% homology at the protein level.27  This finding clearly suggests that, similarly to the human disease, discrete antigens or classes of structurally related epitopes select and activate the leukemic clones in the corresponding mouse models and provides additional molecular evidence for the similarity of the disease developing in humans and the CLL mouse models.

Phenotypic and epigenetic characteristics

The ZAP70 tyrosine kinase is recruited to the cell surface of CLL cells on stimulation through the BCR and promotes activation of the downstream signaling pathway.56,57  In addition, ZAP70+ CLL cells display increased migration properties.58  ZAP70 expression is up-regulated in splenic B cells of Eμ-TCL1 mice59  already in the early stage of the disease (ref. 59 and G.S., M.T.S.B, and P.G., unpublished data, 2008), suggesting an involvement of the kinase during the clonal expansion. Deleting zap70 in the B-cell compartment of Eμ-TCL1 mice may provide new insights into the role of ZAP70 in the dynamics of disease onset, progression, and dissemination.

Recent studies suggest that CLL cells have the functional capacity to express and secrete interleukin (IL)-10,60  a property that characterizes a subset of B cells with immune regulatory functions (B-10 cells).61  Based on the observation in Eμ-TCL1 mice that the expansion of IL-10competent malignant B cells precedes the development of overt leukemia, IL-10 production was suggested to participate in the immunoregulatory capacity of the malignant cells and therefore influence disease progression, outcome, and response to treatment,60  possibly through suppression of the antitumor response.

As observed in CLL patients, leukemic Eμ-TCL1 mice display multiple T-cell alterations, including the shift from a naïve to a memory T-cell subtype62  and defective signal transduction at the immune synapse.63  Also, similar to human CLL,64  leukemic mice present with an increased number of T cells in the PB.62  Interestingly, transfer of leukemic B cells into disease-free -TCL1 mice caused the development of T-cell dysfunctions that allowed the malignant clone to escape from the host immune system and expand in the recipient mice.63  Together, these findings suggest that the Eμ-TCL1 model represents a suitable tool to study the interplay between CLL cells and T cells in disease development and progression.

CLL cells in humans show multiple epigenetic alterations that have been detected also in the -TCL1 model.65  Aberrant methylation of promoter sequences was observed in B cells of -TCL1 mice starting at 3 months of age, and the level of DNA methylation increased with time, suggesting a causative role in the development and progression of CLL.66  In addition, splenic B cells from -TCL1 mice showed an elevated number of hypomethylated regions compared with wild-type B cells.41 

Eμ-TCL1 transgenic mouse model as a tool to investigate pathogenic mechanisms

Due to the similarity of the disease characteristics of CLL in humans and mice and the complete penetrance of the model, Eμ-TCL1 mice have been extensively used in the dissection of the pathogenic mechanisms leading to CLL. Several transgenic and knockout mouse models have been crossed with Eμ-TCL1 mice to elucidate the functional role of specific molecules in the onset and progression of CLL in vivo (Figure 2; supplemental Table 1), providing important new insights into the pathogenic role of those genes in the dysregulation of signaling, proliferation, and apoptosis, in mediating altered trafficking and homing, and in the aberrant cross-talk with the microenvironment.

Figure 2

Study of novel pathogenic mechanisms of CLL in the TCL1-driven leukemia model. Overexpression (transgenic [tg]) or deficiency of different molecules in the Eμ-TCL1 transgenic mouse model variably affects the disease phenotype (BM, bone marrow; TAM, tumor-associated macrophages). The following mouse models have been crossed with Eμ-TCL1 mice: xid,91 pkcβ−/− (or pkcβ+/−),70 xbp1fl/flCD19-Cre,73 tir8−/−,72 dnrag1-tg,71 hs1−/−,79 rhoh−/−,59,80 p53−/−,75 id4−/−,76 miR29a/b-tg,35 fzd6−/−,89 cd44−/−,87 mif−/−,86 ROR1-tg,34 APRIL-tg,85  and baff-tg.84 

Figure 2

Study of novel pathogenic mechanisms of CLL in the TCL1-driven leukemia model. Overexpression (transgenic [tg]) or deficiency of different molecules in the Eμ-TCL1 transgenic mouse model variably affects the disease phenotype (BM, bone marrow; TAM, tumor-associated macrophages). The following mouse models have been crossed with Eμ-TCL1 mice: xid,91 pkcβ−/− (or pkcβ+/−),70 xbp1fl/flCD19-Cre,73 tir8−/−,72 dnrag1-tg,71 hs1−/−,79 rhoh−/−,59,80 p53−/−,75 id4−/−,76 miR29a/b-tg,35 fzd6−/−,89 cd44−/−,87 mif−/−,86 ROR1-tg,34 APRIL-tg,85  and baff-tg.84 

Close modal

As predicted from the study of CLL in humans,67,68  the establishment of suitable Eμ-TCL1 compound mouse models could validate the critical role of BCR signaling in CLL pathogenesis,69,70  provided evidence for the importance of continuous autoantigenic stimulation in CLL development,71  and suggested a proleukemogenic function of Toll-like receptor signaling.72  Interestingly, a recent study found that B cell-specific inhibition of the endoplasmic reticulum stress response, which is coupled to defective spleen tyrosine kinase and Bruton agammaglobulinemia tyrosine kinase activation on BCR stimulation, reduces leukemia development in Eμ-TCL1 mice.73 

As expected from the known tumor suppressor function of TP53, genetic ablation of tp53 in mice, thus mimicking 17p deletion,3  the most deleterious genetic abnormalities in CLL,74  accelerated leukemia onset and progression in the Eμ-TCL1 model.75  Similar observations were made for a putative CLL tumor suppressor gene: inhibitor of DNA binding protein 4 (id4).76 

The notion that aberrant cytoskeletal organization and cell trafficking contribute to CLL pathogenesis potentially through a disrupted interaction of the malignant cells with the bone marrow microenvironment77,78  was experimentally confirmed in vivo in hematopoietic cell-specific Lyn substrate (hs1)−/−Eμ-TCL1 mice.79  In addition, another Eμ-TCL1 compound mouse model provided evidence that cellular motility plays an important role in bone marrow infiltration of CLL cells.59,80 

In CLL patients, endothelial, nurse-like, and antigen-presenting cells release the TNF family members B-cell activating factor (BAFF) and APRIL, which activate the transformed B cells.81-83  In accordance, ectopic expression of the corresponding transgenes in Eμ-TCL1 mice revealed that BAFF and APRIL protect leukemic cells from apoptosis in vivo, resulting in a rapid expansion of the malignant clone.84,85  The importance of microenvironmental interactions in the maintenance of the CLL clone could further be demonstrated by 2 independent Eμ-TCL1 compound mouse models (mif−/−Eμ-TCL1 and cd44−/−Eμ-TCL1 mice).86,87  Moreover, the potential role of the tumor microenvironment in signaling to malignant B cells through wingless-type MMTV integration site family (WNT) receptors88  was addressed by genetically ablating the WNT receptor frizzled 6 (fzd6) in Eμ-TCL1 mice. The results suggested that dysregulated FZD6 expression modulates the disease course as it delayed but did not abrogate tumor growth.89 

In summary, Eμ-TCL1 compound mouse models have proved extremely useful in determining the pathogenic functions of CLL-associated genes in an in vivo context. It will be interesting to explore the leukemogenic role of the newly identified oncogenic mutations in the NOTCH1, MYD88, SF3B1, and BIRC3 genes4-11  by modeling these mutations on the background of Eμ-TCL1, as well as other CLL mouse models.

Eμ-TCL1 transgenic mouse model as a preclinical model for novel therapeutics

Mouse models represent a useful tool to evaluate efficacy, pitfalls, and potential side effects of novel therapeutic strategies. The -TCL1 mouse has been validated as a suitable preclinical model by demonstrating that the response of leukemic mice to fludarabine, a cytotoxic drug used in first-line treatment of CLL, resembles that observed in human patients.29  The disadvantage of the long disease latency was overcome by transferring splenic leukocytes or B cells from leukemic mice into syngeneic or immunodeficient recipient mice.34,84,87,90-99  Generally, transplanted animals developed a frank leukemia within a few weeks or months. Together with xenograft models of CLL,100  such transfer experiments represent reproducible systems for the elucidation of the efficacy and mechanism of action of novel therapies. As such, preclinical studies with -TCL1 mice were instrumental in understanding the pathogenic relevance of molecular targets in disease biology (supplemental Table 2), as summarized in the following.

Inhibition of signaling

Based on the known function of the TCL1 oncogene as an AKT coactivator,38,39  the dependence of TCL1-driven leukemia on AKT signaling has been investigated by inhibiting the mTOR effector with rapamycin in a transfer model.90  The study demonstrated a critical role of the AKT pathway in the disease onset rather than progression.

The fact that CLL cells are chronically stimulated and in lymphoid tissues show a gene expression signature of lymphocytes activated through the BCR101  provided a rationale for the development of strategies based on BCR signaling inhibitors. In clinical practice, the Bruton agammaglobulinemia tyrosine kinase inhibitor ibrutinib (PCI-32765), the spleen tyrosine kinase inhibitor fostamatinib, and the phosphatidylinositol 3-kinase-δ inhibitor idelalisib (GS-1101) caused rapid lymph node shrinkage and transient lymphocytosis in patients102,103  by impairing the retention of CLL cells in the tissue microenvironments.104  Importantly, a transient increase in PB B-cell counts has also been reported early after treatment of mice transplanted with Eμ-TCL1–derived leukemias with fostamatinib and ibrutinib.92,93 

Inhibition of cytoskeletal functions and cell trafficking

The findings that BCR inhibitors cause rapid tissue mobilization of CLL cells established a new paradigm in the CLL field, which identifies the disruption of the homing properties of leukemic cells as a therapeutic approach. This notion is supported by results from HS1 or RHOH-deficient Eμ-TCL1 compound mouse models (supplemental Table 1),79,80  by recent phase 3 studies,105  and by the Food and Drug Administration approval of ibrutinib for the treatment of CLL. In line with these assertions, interfering with the proper cytoskeleton function of CLL cells via the dual v-src avian sarcoma (Schmidt-Ruppin A-2) viral oncogene homolog (SRC) kinase/BCR-Abelson murine leukemia viral oncogene homolog 1 (ABL) inhibitor dasatinib may be a conceivable therapeutic option for the subset of patients with an active v-yes-1 yamaguchi sarcoma viral related oncogene homolog (LYN)/HS1 axis.98  Indeed, inhibition of LYN/HS1 signaling impaired leukemia progression and lymphoid organ infiltration in a transfer model of TCL1-driven leukemia. Moreover, an indirect effect on leukemic cell trafficking may contribute to the efficacy of lenalidomide, an immunomodulatory agent that is currently under investigation as salvage therapy for patients with relapsed or refractory CLL106  and as a consolidation strategy in progressive cases.107  Lenalidomide treatment reduced the chemokine response of CLL cells in vitro and down-modulated the expression of RHOH,80  which, in the Eμ-TCL1 model, is indispensable for the homing of malignant B cells to the bone marrow.80 

Inhibition of tumor-microenvironmental interactions

Several novel therapeutics target the cross-talk between the malignant cells and the tumor microenvironment through neutralization of protumorigenic factors, inhibition of surface receptors, or nuclear transport. Treatment with the anti-CD44 antibody IM7 that exerts a proapoptotic function prevented the accumulation of CD44-expressing leukemic B cells in the PB of an adoptive transfer model of TCL1-driven leukemia early after injection.87  Similarly, administration of the anti-ROR1 monoclonal antibody D10 inhibited the engraftment of leukemic cells from ROR1 × Eμ-TCL1 transgenic mice into ROR1 transgenic recipients, which was accompanied by a strong reduction of PB and splenic B-cell counts.34  A reduction in the number of circulating malignant B cells was also observed on neutralizing the prosurvival effect of BAFF with CD268-Fc and CD269-Fc decoy receptors in a model obtained by intraperitoneal injection of CD5+ leukemic B cells from baff × Eμ-TCL1 transgenic mice into baff transgenic recipients.84  Finally, marked antileukemic effects have recently been reported for the inhibitor of nuclear export KPT-185 and the antibody drug conjugate IMGN529.95,97  KPT-185 acts by blocking the activity of the nuclear exporter chromosome region maintenance 1 protein, or exportin1 (CRM1/XPO1), which is a downstream mediator of microenvironmental signals sustaining CLL cell survival and growth.97  Administration of the small molecule to SCID mice engrafted with leukemic B cells from Eμ-TCL1 mice improved survival of those mice. A benefit in terms of overall survival has also been observed in CD37 transgenic mice engrafted with CD37 × Eμ-TCL1 leukemia on treatment with IMGN529, which consists of the cytotoxic anti-CD37 antibody and the antimicrotubule agent DM1 that exerts an antiproliferative effect.95 

Forced induction of apoptosis

A major pathogenic mechanism in CLL development constitutes the resistance of the tumor cells to apoptosis. Therefore, therapeutics that target the antiapoptotic pathway represent valuable strategies for CLL treatment. Preclinical studies of actinomycin D108  and silvestrol (a plant-derived cyclopenta[b]benzofuran)109  that, respectively, inhibit transcription and translation, and of 17-dimethylaminoethylamino-17-demethoxygeldanamycin (17-DMAG) (a derivative of the heat shock protein 90 inhibitor geldanamycin)94  showed promising results and were indeed associated with a reduction in the levels of antiapoptotic proteins. The studies may provide a rationale for the treatment of CLL cases carrying a deletion of the 17p locus (that are resistant to standard chemotherapy including fludarabine)110  with actinomycin D or silvestrol, which can induce apoptosis in a P53-independent fashion.108,109 

Epigenome targeting

Finally, CLL cells display epigenetic alterations, and evidence suggests that overexpression of histone deacetylase enzymes contributes to CLL pathogenesis.65,111  The histone deacetylase inhibitor MGCD0103 induced apoptosis112  and autophagy suppression113  in CLL cells in vitro and showed antitumor activity in a phase 2 clinical trial.114  Moreover, the class I and II deacetylase (DAC) inhibitor, AR-42, prolonged the overall survival of mice transplanted with TCL1-driven leukemia.99  Class I and II inhibitors such as AR-42 that target both histone and nonhistone substrates may be promising therapeutics for a multitarget therapy.

Several genetically engineered mouse models of CLL have been generated that mimic the human disease. The 13q14 deletion models provided the first in vivo evidence of the tumor-suppressor function of a CLL-associated genetic lesion. As numerous studies revealed similarities between CLL developing in humans and Eμ-TCL1 mice, this transgenic model has become a suitable tool in the investigation of pathogenic mechanisms of CLL and represents a convenient preclinical model, particularly due to its complete disease penetrance.

Based on the stereotypic antigen receptors in the Eμ-TCL1 mice that recall those in some aggressive forms of human CLL, and the association of the 13q14-deletion in humans with CLL bearing somatically mutated IGHV genes that have a more favorable disease course, it has been suggested that the corresponding mouse models mimic the aggressive and indolent forms of CLL, respectively. However, this view may have to be revisited due to the circumstance that the CD5+ lymphoproliferations developing in the Eμ-TCL1 transgenic and 13q14 deletion models are similar regarding their phenotype and their expression of unmutated IGHV genes and of stereotypic antigen receptors, indicating that the target cell of the oncogenic transformation may be the same in the 2 models. Two implications ensue from this: first, in lieu of knowledge on the precise cell of origin of CLL in humans, it is impossible to confidently assign the CLL developing in the mouse models to a particular human CLL subtype. Second, the virtual absence of somatically mutated IGHV genes in CLLs developing in the mouse models may suggest that the normal cellular counterpart(s) of CLL differ among mice and humans, eg, that a somatically mutated CD5+ subset may simply not exist in the mouse. In support of the latter assumption is the observation that, similar to humans, CD5-negative NHL with somatically mutated IGHV genes do develop in 13q14 deletion models.

Regardless of these unresolved issues, genetically engineered CLL mouse models, together with xenograft models, have emerged as valuable tools for the first-line testing of new therapies. A downside of the transgenic CLL models in their use as preclinical models is the long latency until disease develops. The presence of stereotypic receptors and active BCR signaling in the leukemic cells of those mice may provide a rationale for the development of a CLL mouse model with potentially shorter latency by combining an antigen or BCR-driven mechanism with the commonly known, or newly identified, CLL-associated genetic aberrations.

The online version of this article contains a data supplement.

The authors thank Riccardo Dalla-Favera and Federico Caligaris-Cappio for discussions and support.

This project was supported by Associazione Italiana per la Ricerca sul Cancro (Investigator Grant and Special Program Molecular Clinical Oncology–5 per mille #9965) and Ricerca Finalizzata 2010–Ministero della Salute, Roma, and by a grant from the CLL Global Research Foundation.

Contribution: G.S., M.T.S.B., P.G., and U.K. wrote the paper.

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

The current affiliation for G.S. is Department of Experimental, Diagnostic and Specialty Medicine, University of Bologna, Bologna, Italy.

Correspondence: Ulf Klein, Herbert Irving Comprehensive Cancer Center, Columbia University, 1130 St. Nicholas Ave, New York, NY 10032; e-mail: uk30@columbia.edu; or Paolo Ghia, Università Vita-Salute San Raffaele, Via Olgettina 58, 20132 Milan, Italy; e-mail: ghia.paolo@hsr.it.

1
Zenz
 
T
Mertens
 
D
Küppers
 
R
Döhner
 
H
Stilgenbauer
 
S
From pathogenesis to treatment of chronic lymphocytic leukaemia.
Nat Rev Cancer
2010
, vol. 
10
 
1
(pg. 
37
-
50
)
2
Gaidano
 
G
Foà
 
R
Dalla-Favera
 
R
Molecular pathogenesis of chronic lymphocytic leukemia.
J Clin Invest
2012
, vol. 
122
 
10
(pg. 
3432
-
3438
)
3
Döhner
 
H
Stilgenbauer
 
S
Benner
 
A
, et al. 
Genomic aberrations and survival in chronic lymphocytic leukemia.
N Engl J Med
2000
, vol. 
343
 
26
(pg. 
1910
-
1916
)
4
Fabbri
 
G
Rasi
 
S
Rossi
 
D
, et al. 
Analysis of the chronic lymphocytic leukemia coding genome: role of NOTCH1 mutational activation.
J Exp Med
2011
, vol. 
208
 
7
(pg. 
1389
-
1401
)
5
Puente
 
XS
Pinyol
 
M
Quesada
 
V
, et al. 
Whole-genome sequencing identifies recurrent mutations in chronic lymphocytic leukaemia.
Nature
2011
, vol. 
475
 
7354
(pg. 
101
-
105
)
6
Rossi
 
D
Bruscaggin
 
A
Spina
 
V
, et al. 
Mutations of the SF3B1 splicing factor in chronic lymphocytic leukemia: association with progression and fludarabine-refractoriness.
Blood
2011
, vol. 
118
 
26
(pg. 
6904
-
6908
)
7
Wang
 
L
Lawrence
 
MS
Wan
 
Y
, et al. 
SF3B1 and other novel cancer genes in chronic lymphocytic leukemia.
N Engl J Med
2011
, vol. 
365
 
26
(pg. 
2497
-
2506
)
8
Quesada
 
V
Conde
 
L
Villamor
 
N
, et al. 
Exome sequencing identifies recurrent mutations of the splicing factor SF3B1 gene in chronic lymphocytic leukemia.
Nat Genet
2012
, vol. 
44
 
1
(pg. 
47
-
52
)
9
Rossi
 
D
Fangazio
 
M
Rasi
 
S
, et al. 
Disruption of BIRC3 associates with fludarabine chemorefractoriness in TP53 wild-type chronic lymphocytic leukemia.
Blood
2012
, vol. 
119
 
12
(pg. 
2854
-
2862
)
10
Balatti
 
V
Bottoni
 
A
Palamarchuk
 
A
, et al. 
NOTCH1 mutations in CLL associated with trisomy 12.
Blood
2012
, vol. 
119
 
2
(pg. 
329
-
331
)
11
Landau
 
DA
Carter
 
SL
Stojanov
 
P
, et al. 
Evolution and impact of subclonal mutations in chronic lymphocytic leukemia.
Cell
2013
, vol. 
152
 
4
(pg. 
714
-
726
)
12
Calin
 
GA
Cimmino
 
A
Fabbri
 
M
, et al. 
MiR-15a and miR-16-1 cluster functions in human leukemia.
Proc Natl Acad Sci USA
2008
, vol. 
105
 
13
(pg. 
5166
-
5171
)
13
Fukuda
 
T
Chen
 
L
Endo
 
T
, et al. 
Antisera induced by infusions of autologous Ad-CD154-leukemia B cells identify ROR1 as an oncofetal antigen and receptor for Wnt5a.
Proc Natl Acad Sci USA
2008
, vol. 
105
 
8
(pg. 
3047
-
3052
)
14
Daneshmanesh
 
AH
Mikaelsson
 
E
Jeddi-Tehrani
 
M
, et al. 
Ror1, a cell surface receptor tyrosine kinase is expressed in chronic lymphocytic leukemia and may serve as a putative target for therapy.
Int J Cancer
2008
, vol. 
123
 
5
(pg. 
1190
-
1195
)
15
Herling
 
M
Patel
 
KA
Khalili
 
J
, et al. 
TCL1 shows a regulated expression pattern in chronic lymphocytic leukemia that correlates with molecular subtypes and proliferative state.
Leukemia
2006
, vol. 
20
 
2
(pg. 
280
-
285
)
16
Di Bernardo
 
MC
Crowther-Swanepoel
 
D
Broderick
 
P
, et al. 
A genome-wide association study identifies six susceptibility loci for chronic lymphocytic leukemia.
Nat Genet
2008
, vol. 
40
 
10
(pg. 
1204
-
1210
)
17
De Silva
 
NS
Simonetti
 
G
Heise
 
N
Klein
 
U
The diverse roles of IRF4 in late germinal center B-cell differentiation.
Immunol Rev
2012
, vol. 
247
 
1
(pg. 
73
-
92
)
18
Chiorazzi
 
N
Ferrarini
 
M
B cell chronic lymphocytic leukemia: lessons learned from studies of the B cell antigen receptor.
Annu Rev Immunol
2003
, vol. 
21
 (pg. 
841
-
894
)
19
Rawstron
 
AC
Bennett
 
FL
O’Connor
 
SJ
, et al. 
Monoclonal B-cell lymphocytosis and chronic lymphocytic leukemia.
N Engl J Med
2008
, vol. 
359
 
6
(pg. 
575
-
583
)
20
Fazi
 
C
Scarfò
 
L
Pecciarini
 
L
, et al. 
General population low-count CLL-like MBL persists over time without clinical progression, although carrying the same cytogenetic abnormalities of CLL.
Blood
2011
, vol. 
118
 
25
(pg. 
6618
-
6625
)
21
Landgren
 
O
Albitar
 
M
Ma
 
W
, et al. 
B-cell clones as early markers for chronic lymphocytic leukemia.
N Engl J Med
2009
, vol. 
360
 
7
(pg. 
659
-
667
)
22
Liu
 
Y
Hermanson
 
M
Grandér
 
D
, et al. 
13q deletions in lymphoid malignancies.
Blood
1995
, vol. 
86
 
5
(pg. 
1911
-
1915
)
23
Liu
 
Y
Corcoran
 
M
Rasool
 
O
, et al. 
Cloning of two candidate tumor suppressor genes within a 10 kb region on chromosome 13q14, frequently deleted in chronic lymphocytic leukemia.
Oncogene
1997
, vol. 
15
 
20
(pg. 
2463
-
2473
)
24
Migliazza
 
A
Bosch
 
F
Komatsu
 
H
, et al. 
Nucleotide sequence, transcription map, and mutation analysis of the 13q14 chromosomal region deleted in B-cell chronic lymphocytic leukemia.
Blood
2001
, vol. 
97
 
7
(pg. 
2098
-
2104
)
25
Calin
 
GA
Dumitru
 
CD
Shimizu
 
M
, et al. 
Frequent deletions and down-regulation of micro- RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia.
Proc Natl Acad Sci USA
2002
, vol. 
99
 
24
(pg. 
15524
-
15529
)
26
Klein
 
U
Lia
 
M
Crespo
 
M
, et al. 
The DLEU2/miR-15a/16-1 cluster controls B cell proliferation and its deletion leads to chronic lymphocytic leukemia.
Cancer Cell
2010
, vol. 
17
 
1
(pg. 
28
-
40
)
27
Lia
 
M
Carette
 
A
Tang
 
H
, et al. 
Functional dissection of the chromosome 13q14 tumor-suppressor locus using transgenic mouse lines.
Blood
2012
, vol. 
119
 
13
(pg. 
2981
-
2990
)
28
Bichi
 
R
Shinton
 
SA
Martin
 
ES
, et al. 
Human chronic lymphocytic leukemia modeled in mouse by targeted TCL1 expression.
Proc Natl Acad Sci USA
2002
, vol. 
99
 
10
(pg. 
6955
-
6960
)
29
Johnson
 
AJ
Lucas
 
DM
Muthusamy
 
N
, et al. 
Characterization of the TCL-1 transgenic mouse as a preclinical drug development tool for human chronic lymphocytic leukemia.
Blood
2006
, vol. 
108
 
4
(pg. 
1334
-
1338
)
30
Yan
 
XJ
Albesiano
 
E
Zanesi
 
N
, et al. 
B cell receptors in TCL1 transgenic mice resemble those of aggressive, treatment-resistant human chronic lymphocytic leukemia.
Proc Natl Acad Sci USA
2006
, vol. 
103
 
31
(pg. 
11713
-
11718
)
31
Stein
 
JV
López-Fraga
 
M
Elustondo
 
FA
, et al. 
APRIL modulates B and T cell immunity.
J Clin Invest
2002
, vol. 
109
 
12
(pg. 
1587
-
1598
)
32
Planelles
 
L
Carvalho-Pinto
 
CE
Hardenberg
 
G
, et al. 
APRIL promotes B-1 cell-associated neoplasm.
Cancer Cell
2004
, vol. 
6
 
4
(pg. 
399
-
408
)
33
Zapata
 
JM
Krajewska
 
M
Krajewski
 
S
, et al. 
TNFR-associated factor family protein expression in normal tissues and lymphoid malignancies.
J Immunol
2000
, vol. 
165
 
9
(pg. 
5084
-
5096
)
34
Widhopf
 
GF
Cui
 
B
Ghia
 
EM
, et al. 
ROR1 can interact with TCL1 and enhance leukemogenesis in Eμ-TCL1 transgenic mice.
Proc Natl Acad Sci USA
2014
, vol. 
111
 
2
(pg. 
793
-
798
)
35
Santanam
 
U
Zanesi
 
N
Efanov
 
A
, et al. 
Chronic lymphocytic leukemia modeled in mouse by targeted miR-29 expression.
Proc Natl Acad Sci USA
2010
, vol. 
107
 
27
(pg. 
12210
-
12215
)
36
Shukla
 
V
Ma
 
S
Hardy
 
RR
Joshi
 
SS
Lu
 
R
A role for IRF4 in the development of CLL.
Blood
2013
, vol. 
122
 
16
(pg. 
2848
-
2855
)
37
ter Brugge
 
PJ
Ta
 
VB
de Bruijn
 
MJ
, et al. 
A mouse model for chronic lymphocytic leukemia based on expression of the SV40 large T antigen.
Blood
2009
, vol. 
114
 
1
(pg. 
119
-
127
)
38
Pekarsky
 
Y
Koval
 
A
Hallas
 
C
, et al. 
Tcl1 enhances Akt kinase activity and mediates its nuclear translocation.
Proc Natl Acad Sci USA
2000
, vol. 
97
 
7
(pg. 
3028
-
3033
)
39
Laine
 
J
Künstle
 
G
Obata
 
T
Sha
 
M
Noguchi
 
M
The protooncogene TCL1 is an Akt kinase coactivator.
Mol Cell
2000
, vol. 
6
 
2
(pg. 
395
-
407
)
40
Pekarsky
 
Y
Palamarchuk
 
A
Maximov
 
V
, et al. 
Tcl1 functions as a transcriptional regulator and is directly involved in the pathogenesis of CLL.
Proc Natl Acad Sci USA
2008
, vol. 
105
 
50
(pg. 
19643
-
19648
)
41
Palamarchuk
 
A
Yan
 
PS
Zanesi
 
N
, et al. 
Tcl1 protein functions as an inhibitor of de novo DNA methylation in B-cell chronic lymphocytic leukemia (CLL).
Proc Natl Acad Sci USA
2012
, vol. 
109
 
7
(pg. 
2555
-
2560
)
42
Zapata
 
JM
Krajewska
 
M
Morse
 
HC
Choi
 
Y
Reed
 
JC
TNF receptor-associated factor (TRAF) domain and Bcl-2 cooperate to induce small B cell lymphoma/chronic lymphocytic leukemia in transgenic mice.
Proc Natl Acad Sci USA
2004
, vol. 
101
 
47
(pg. 
16600
-
16605
)
43
Simonetti
 
G
Carette
 
A
Silva
 
K
, et al. 
IRF4 controls the positioning of mature B cells in the lymphoid microenvironments by regulating NOTCH2 expression and activity.
J Exp Med
2013
, vol. 
210
 
13
(pg. 
2887
-
2902
)
44
Berland
 
R
Wortis
 
HH
Origins and functions of B-1 cells with notes on the role of CD5.
Annu Rev Immunol
2002
, vol. 
20
 (pg. 
253
-
300
)
45
Stall
 
AM
Fariñas
 
MC
Tarlinton
 
DM
, et al. 
Ly-1 B-cell clones similar to human chronic lymphocytic leukemias routinely develop in older normal mice and young autoimmune (New Zealand Black-related) animals.
Proc Natl Acad Sci USA
1988
, vol. 
85
 
19
(pg. 
7312
-
7316
)
46
LeMaoult
 
J
Delassus
 
S
Dyall
 
R
Nikolić-Zugić
 
J
Kourilsky
 
P
Weksler
 
ME
Clonal expansions of B lymphocytes in old mice.
J Immunol
1997
, vol. 
159
 
8
(pg. 
3866
-
3874
)
47
Chiorazzi
 
N
Ferrarini
 
M
Cellular origin(s) of chronic lymphocytic leukemia: cautionary notes and additional considerations and possibilities.
Blood
2011
, vol. 
117
 
6
(pg. 
1781
-
1791
)
48
Forconi
 
F
Potter
 
KN
Wheatley
 
I
, et al. 
The normal IGHV1-69-derived B-cell repertoire contains stereotypic patterns characteristic of unmutated CLL.
Blood
2010
, vol. 
115
 
1
(pg. 
71
-
77
)
49
Seifert
 
M
Sellmann
 
L
Bloehdorn
 
J
, et al. 
Cellular origin and pathophysiology of chronic lymphocytic leukemia.
J Exp Med
2012
, vol. 
209
 
12
(pg. 
2183
-
2198
)
50
Dühren-von Minden
 
M
Übelhart
 
R
Schneider
 
D
, et al. 
Chronic lymphocytic leukaemia is driven by antigen-independent cell-autonomous signalling.
Nature
2012
, vol. 
489
 
7415
(pg. 
309
-
312
)
51
Chu
 
CC
Catera
 
R
Hatzi
 
K
, et al. 
Chronic lymphocytic leukemia antibodies with a common stereotypic rearrangement recognize nonmuscle myosin heavy chain IIA.
Blood
2008
, vol. 
112
 
13
(pg. 
5122
-
5129
)
52
Catera
 
R
Silverman
 
GJ
Hatzi
 
K
, et al. 
Chronic lymphocytic leukemia cells recognize conserved epitopes associated with apoptosis and oxidation.
Mol Med
2008
, vol. 
14
 
11-12
(pg. 
665
-
674
)
53
Lanemo Myhrinder
 
A
Hellqvist
 
E
Sidorova
 
E
, et al. 
A new perspective: molecular motifs on oxidized LDL, apoptotic cells, and bacteria are targets for chronic lymphocytic leukemia antibodies.
Blood
2008
, vol. 
111
 
7
(pg. 
3838
-
3848
)
54
Hoogeboom
 
R
van Kessel
 
KP
Hochstenbach
 
F
, et al. 
A mutated B cell chronic lymphocytic leukemia subset that recognizes and responds to fungi.
J Exp Med
2013
, vol. 
210
 
1
(pg. 
59
-
70
)
55
Stamatopoulos
 
K
Belessi
 
C
Moreno
 
C
, et al. 
Over 20% of patients with chronic lymphocytic leukemia carry stereotyped receptors: Pathogenetic implications and clinical correlations.
Blood
2007
, vol. 
109
 
1
(pg. 
259
-
270
)
56
Chen
 
L
Widhopf
 
G
Huynh
 
L
, et al. 
Expression of ZAP-70 is associated with increased B-cell receptor signaling in chronic lymphocytic leukemia.
Blood
2002
, vol. 
100
 
13
(pg. 
4609
-
4614
)
57
Calpe
 
E
Codony
 
C
Baptista
 
MJ
, et al. 
ZAP-70 enhances migration of malignant B lymphocytes toward CCL21 by inducing CCR7 expression via IgM-ERK1/2 activation.
Blood
2011
, vol. 
118
 
16
(pg. 
4401
-
4410
)
58
Deaglio
 
S
Vaisitti
 
T
Aydin
 
S
, et al. 
CD38 and ZAP-70 are functionally linked and mark CLL cells with high migratory potential.
Blood
2007
, vol. 
110
 
12
(pg. 
4012
-
4021
)
59
Sanchez-Aguilera
 
A
Rattmann
 
I
Drew
 
DZ
, et al. 
Involvement of RhoH GTPase in the development of B-cell chronic lymphocytic leukemia.
Leukemia
2010
, vol. 
24
 
1
(pg. 
97
-
104
)
60
DiLillo
 
DJ
Weinberg
 
JB
Yoshizaki
 
A
, et al. 
Chronic lymphocytic leukemia and regulatory B cells share IL-10 competence and immunosuppressive function.
Leukemia
2013
, vol. 
27
 
1
(pg. 
170
-
182
)
61
Mauri
 
C
Bosma
 
A
Immune regulatory function of B cells.
Annu Rev Immunol
2012
, vol. 
30
 (pg. 
221
-
241
)
62
Hofbauer
 
JP
Heyder
 
C
Denk
 
U
, et al. 
Development of CLL in the TCL1 transgenic mouse model is associated with severe skewing of the T-cell compartment homologous to human CLL.
Leukemia
2011
, vol. 
25
 
9
(pg. 
1452
-
1458
)
63
Gorgun
 
G
Ramsay
 
AG
Holderried
 
TA
, et al. 
E(mu)-TCL1 mice represent a model for immunotherapeutic reversal of chronic lymphocytic leukemia-induced T-cell dysfunction.
Proc Natl Acad Sci USA
2009
, vol. 
106
 
15
(pg. 
6250
-
6255
)
64
Kimby
 
E
Mellstedt
 
H
Nilsson
 
B
Björkholm
 
M
Holm
 
G
T lymphocyte subpopulations in chronic lymphocytic leukemia of B cell type in relation to immunoglobulin isotype(s) on the leukemic clone and to clinical features.
Eur J Haematol
1987
, vol. 
38
 
3
(pg. 
261
-
267
)
65
Chen
 
SS
Sherman
 
MH
Hertlein
 
E
, et al. 
Epigenetic alterations in a murine model for chronic lymphocytic leukemia.
Cell Cycle
2009
, vol. 
8
 
22
(pg. 
3663
-
3667
)
66
Chen
 
SS
Raval
 
A
Johnson
 
AJ
, et al. 
Epigenetic changes during disease progression in a murine model of human chronic lymphocytic leukemia.
Proc Natl Acad Sci USA
2009
, vol. 
106
 
32
(pg. 
13433
-
13438
)
67
Stevenson
 
FK
Krysov
 
S
Davies
 
AJ
Steele
 
AJ
Packham
 
G
B-cell receptor signaling in chronic lymphocytic leukemia.
Blood
2011
, vol. 
118
 
16
(pg. 
4313
-
4320
)
68
Burger
 
JA
Chiorazzi
 
N
B cell receptor signaling in chronic lymphocytic leukemia.
Trends Immunol
2013
, vol. 
34
 
12
(pg. 
592
-
601
)
69
Woyach
 
JA
Bojnik
 
E
Ruppert
 
AS
, et al. 
Bruton’s tyrosine kinase (BTK) function is important to the development and expansion of chronic lymphocytic leukemia (CLL).
Blood
2014
, vol. 
123
 
8
(pg. 
1207
-
1213
)
70
Holler
 
C
Piñón
 
JD
Denk
 
U
, et al. 
PKCbeta is essential for the development of chronic lymphocytic leukemia in the TCL1 transgenic mouse model: validation of PKCbeta as a therapeutic target in chronic lymphocytic leukemia.
Blood
2009
, vol. 
113
 
12
(pg. 
2791
-
2794
)
71
Nganga
 
VK
Palmer
 
VL
Naushad
 
H
, et al. 
Accelerated progression of chronic lymphocytic leukemia in Eμ-TCL1 mice expressing catalytically inactive RAG1.
Blood
2013
, vol. 
121
 
19
(pg. 
3855
-
3866, S1-S16
)
72
Bertilaccio
 
MT
Simonetti
 
G
Dagklis
 
A
, et al. 
Lack of TIR8/SIGIRR triggers progression of chronic lymphocytic leukemia in mouse models.
Blood
2011
, vol. 
118
 
3
(pg. 
660
-
669
)
73
Tang
 
CH
Ranatunga
 
S
Kriss
 
CL
, et al. 
Inhibition of ER stress-associated IRE-1/XBP-1 pathway reduces leukemic cell survival.
J Clin Invest
2014
, vol. 
124
 
6
(pg. 
2585
-
2598
)
74
Dufour
 
A
Palermo
 
G
Zellmeier
 
E
, et al. 
Inactivation of TP53 correlates with disease progression and low miR-34a expression in previously treated chronic lymphocytic leukemia patients.
Blood
2013
, vol. 
121
 
18
(pg. 
3650
-
3657
)
75
Liu
 
J
Chen
 
G
Feng
 
L
, et al. 
Loss of p53 and altered miR15-a/16-1-MCL-1 pathway in CLL: insights from TCL1-Tg:p53(-/-) mouse model and primary human leukemia cells.
Leukemia
2014
, vol. 
28
 
1
(pg. 
118
-
128
)
76
Chen
 
SS
Claus
 
R
Lucas
 
DM
, et al. 
Silencing of the inhibitor of DNA binding protein 4 (ID4) contributes to the pathogenesis of mouse and human CLL.
Blood
2011
, vol. 
117
 
3
(pg. 
862
-
871
)
77
Scielzo
 
C
Ghia
 
P
Conti
 
A
, et al. 
HS1 protein is differentially expressed in chronic lymphocytic leukemia patient subsets with good or poor prognoses.
J Clin Invest
2005
, vol. 
115
 
6
(pg. 
1644
-
1650
)
78
Muzio
 
M
Scielzo
 
C
Frenquelli
 
M
, et al. 
HS1 complexes with cytoskeleton adapters in normal and malignant chronic lymphocytic leukemia B cells.
Leukemia
2007
, vol. 
21
 
9
(pg. 
2067
-
2070
)
79
Scielzo
 
C
Bertilaccio
 
MT
Simonetti
 
G
, et al. 
HS1 has a central role in the trafficking and homing of leukemic B cells.
Blood
2010
, vol. 
116
 
18
(pg. 
3537
-
3546
)
80
Troeger
 
A
Johnson
 
AJ
Wood
 
J
, et al. 
RhoH is critical for cell-microenvironment interactions in chronic lymphocytic leukemia in mice and humans.
Blood
2012
, vol. 
119
 
20
(pg. 
4708
-
4718
)
81
Kern
 
C
Cornuel
 
JF
Billard
 
C
, et al. 
Involvement of BAFF and APRIL in the resistance to apoptosis of B-CLL through an autocrine pathway.
Blood
2004
, vol. 
103
 
2
(pg. 
679
-
688
)
82
Nishio
 
M
Endo
 
T
Tsukada
 
N
, et al. 
Nurselike cells express BAFF and APRIL, which can promote survival of chronic lymphocytic leukemia cells via a paracrine pathway distinct from that of SDF-1alpha.
Blood
2005
, vol. 
106
 
3
(pg. 
1012
-
1020
)
83
Cols
 
M
Barra
 
CM
He
 
B
, et al. 
Stromal endothelial cells establish a bidirectional crosstalk with chronic lymphocytic leukemia cells through the TNF-related factors BAFF, APRIL, and CD40L.
J Immunol
2012
, vol. 
188
 
12
(pg. 
6071
-
6083
)
84
Enzler
 
T
Kater
 
AP
Zhang
 
W
, et al. 
Chronic lymphocytic leukemia of Emu-TCL1 transgenic mice undergoes rapid cell turnover that can be offset by extrinsic CD257 to accelerate disease progression.
Blood
2009
, vol. 
114
 
20
(pg. 
4469
-
4476
)
85
Lascano
 
V
Guadagnoli
 
M
Schot
 
JG
, et al. 
Chronic lymphocytic leukemia disease progression is accelerated by APRIL-TACI interaction in the TCL1 transgenic mouse model.
Blood
2013
, vol. 
122
 
24
(pg. 
3960
-
3963
)
86
Reinart
 
N
Nguyen
 
PH
Boucas
 
J
, et al. 
Delayed development of chronic lymphocytic leukemia in the absence of macrophage migration inhibitory factor.
Blood
2013
, vol. 
121
 
5
(pg. 
812
-
821
)
87
Fedorchenko
 
O
Stiefelhagen
 
M
Peer-Zada
 
AA
, et al. 
CD44 regulates the apoptotic response and promotes disease development in chronic lymphocytic leukemia.
Blood
2013
, vol. 
121
 
20
(pg. 
4126
-
4136
)
88
Lu
 
D
Zhao
 
Y
Tawatao
 
R
, et al. 
Activation of the Wnt signaling pathway in chronic lymphocytic leukemia.
Proc Natl Acad Sci USA
2004
, vol. 
101
 
9
(pg. 
3118
-
3123
)
89
Wu
 
QL
Zierold
 
C
Ranheim
 
EA
Dysregulation of Frizzled 6 is a critical component of B-cell leukemogenesis in a mouse model of chronic lymphocytic leukemia.
Blood
2009
, vol. 
113
 
13
(pg. 
3031
-
3039
)
90
Zanesi
 
N
Aqeilan
 
R
Drusco
 
A
, et al. 
Effect of rapamycin on mouse chronic lymphocytic leukemia and the development of nonhematopoietic malignancies in Emu-TCL1 transgenic mice.
Cancer Res
2006
, vol. 
66
 
2
(pg. 
915
-
920
)
91
Woyach
 
JA
Bojnik
 
E
Ruppert
 
AS
, et al. 
Bruton’s tyrosine kinase (BTK) function is important to the development and expansion of chronic lymphocytic leukemia (CLL).
Blood
2014
, vol. 
123
 
8
(pg. 
1207
-
1213
)
92
Ponader
 
S
Chen
 
SS
Buggy
 
JJ
, et al. 
The Bruton tyrosine kinase inhibitor PCI-32765 thwarts chronic lymphocytic leukemia cell survival and tissue homing in vitro and in vivo.
Blood
2012
, vol. 
119
 
5
(pg. 
1182
-
1189
)
93
Suljagic
 
M
Longo
 
PG
Bennardo
 
S
, et al. 
The Syk inhibitor fostamatinib disodium (R788) inhibits tumor growth in the Eμ- TCL1 transgenic mouse model of CLL by blocking antigen-dependent B-cell receptor signaling.
Blood
2010
, vol. 
116
 
23
(pg. 
4894
-
4905
)
94
Hertlein
 
E
Wagner
 
AJ
Jones
 
J
, et al. 
17-DMAG targets the nuclear factor-kappaB family of proteins to induce apoptosis in chronic lymphocytic leukemia: clinical implications of HSP90 inhibition.
Blood
2010
, vol. 
116
 
1
(pg. 
45
-
53
)
95
Beckwith
 
KA
Frissora
 
FW
Stefanovski
 
MR
, et al. 
The CD37-targeted antibody-drug conjugate IMGN529 is highly active against human CLL and in a novel CD37 transgenic murine leukemia model [published online ahead of print January 21, 2014].
Leukemia
96
Wu
 
QL
Buhtoiarov
 
IN
Sondel
 
PM
Rakhmilevich
 
AL
Ranheim
 
EA
Tumoricidal effects of activated macrophages in a mouse model of chronic lymphocytic leukemia.
J Immunol
2009
, vol. 
182
 
11
(pg. 
6771
-
6778
)
97
Lapalombella
 
R
Sun
 
Q
Williams
 
K
, et al. 
Selective inhibitors of nuclear export show that CRM1/XPO1 is a target in chronic lymphocytic leukemia.
Blood
2012
, vol. 
120
 
23
(pg. 
4621
-
4634
)
98
ten Hacken
 
E
Scielzo
 
C
Bertilaccio
 
MT
, et al. 
Targeting the LYN/HS1 signaling axis in chronic lymphocytic leukemia.
Blood
2013
, vol. 
121
 
12
(pg. 
2264
-
2273
)
99
Lucas
 
DM
Alinari
 
L
West
 
DA
, et al. 
The novel deacetylase inhibitor AR-42 demonstrates pre-clinical activity in B-cell malignancies in vitro and in vivo.
PLoS ONE
2010
, vol. 
5
 
6
pg. 
e10941
 
100
Bertilaccio
 
MT
Scielzo
 
C
Simonetti
 
G
, et al. 
Xenograft models of chronic lymphocytic leukemia: problems, pitfalls and future directions.
Leukemia
2013
, vol. 
27
 
3
(pg. 
534
-
540
)
101
Herishanu
 
Y
Pérez-Galán
 
P
Liu
 
D
, et al. 
The lymph node microenvironment promotes B-cell receptor signaling, NF-kappaB activation, and tumor proliferation in chronic lymphocytic leukemia.
Blood
2011
, vol. 
117
 
2
(pg. 
563
-
574
)
102
Burger
 
JA
Buggy
 
JJ
Emerging drug profiles: Bruton tyrosine kinase (BTK) inhibitor ibrutinib (PCI-32765).
Leuk Lymphoma
2013
, vol. 
54
 
11
(pg. 
2385
-
2391
)
103
Robak
 
T
Robak
 
P
BCR signaling in chronic lymphocytic leukemia and related inhibitors currently in clinical studies.
Int Rev Immunol
2013
, vol. 
32
 
4
(pg. 
358
-
376
)
104
Burger
 
JA
Inhibiting B-cell receptor signaling pathways in chronic lymphocytic leukemia.
Curr Hematol Malig Rep
2012
, vol. 
7
 
1
(pg. 
26
-
33
)
105
Jones
 
JA
Byrd
 
JC
How will B-cell-receptor-targeted therapies change future CLL therapy?
Blood
2014
, vol. 
123
 
10
(pg. 
1455
-
1460
)
106
Badoux
 
XC
Keating
 
MJ
Wen
 
S
, et al. 
Phase II study of lenalidomide and rituximab as salvage therapy for patients with relapsed or refractory chronic lymphocytic leukemia.
J Clin Oncol
2013
, vol. 
31
 
5
(pg. 
584
-
591
)
107
Shanafelt
 
TD
Ramsay
 
AG
Zent
 
CS
, et al. 
Long-term repair of T-cell synapse activity in a phase II trial of chemoimmunotherapy followed by lenalidomide consolidation in previously untreated chronic lymphocytic leukemia (CLL).
Blood
2013
, vol. 
121
 
20
(pg. 
4137
-
4141
)
108
Merkel
 
O
Wacht
 
N
Sifft
 
E
, et al. 
Actinomycin D induces p53-independent cell death and prolongs survival in high-risk chronic lymphocytic leukemia.
Leukemia
2012
, vol. 
26
 
12
(pg. 
2508
-
2516
)
109
Lucas
 
DM
Edwards
 
RB
Lozanski
 
G
, et al. 
The novel plant-derived agent silvestrol has B-cell selective activity in chronic lymphocytic leukemia and acute lymphoblastic leukemia in vitro and in vivo.
Blood
2009
, vol. 
113
 
19
(pg. 
4656
-
4666
)
110
Döhner
 
H
Fischer
 
K
Bentz
 
M
, et al. 
p53 gene deletion predicts for poor survival and non-response to therapy with purine analogs in chronic B-cell leukemias.
Blood
1995
, vol. 
85
 
6
(pg. 
1580
-
1589
)
111
Sampath
 
D
Liu
 
C
Vasan
 
K
, et al. 
Histone deacetylases mediate the silencing of miR-15a, miR-16, and miR-29b in chronic lymphocytic leukemia.
Blood
2012
, vol. 
119
 
5
(pg. 
1162
-
1172
)
112
El-Khoury
 
V
Moussay
 
E
Janji
 
B
, et al. 
The histone deacetylase inhibitor MGCD0103 induces apoptosis in B-cell chronic lymphocytic leukemia cells through a mitochondria-mediated caspase activation cascade.
Mol Cancer Ther
2010
, vol. 
9
 
5
(pg. 
1349
-
1360
)
113
El-Khoury
 
V
Pierson
 
S
Szwarcbart
 
E
, et al. 
Disruption of autophagy by the histone deacetylase inhibitor MGCD0103 and its therapeutic implication in B-cell chronic lymphocytic leukemia [published online ahead of pritn January 14, 2014].
Leukemia
114
Blum
 
KA
Advani
 
A
Fernandez
 
L
, et al. 
Phase II study of the histone deacetylase inhibitor MGCD0103 in patients with previously treated chronic lymphocytic leukaemia.
Br J Haematol
2009
, vol. 
147
 
4
(pg. 
507
-
514
)
115
Calin
 
GA
Ferracin
 
M
Cimmino
 
A
, et al. 
A MicroRNA signature associated with prognosis and progression in chronic lymphocytic leukemia.
N Engl J Med
2005
, vol. 
353
 
17
(pg. 
1793
-
1801
)
116
Raveche
 
ES
Salerno
 
E
Scaglione
 
BJ
, et al. 
Abnormal microRNA-16 locus with synteny to human 13q14 linked to CLL in NZB mice.
Blood
2007
, vol. 
109
 
12
(pg. 
5079
-
5086
)
117
Cimmino
 
A
Calin
 
GA
Fabbri
 
M
, et al. 
miR-15 and miR-16 induce apoptosis by targeting BCL2.
Proc Natl Acad Sci USA
2005
, vol. 
102
 
39
(pg. 
13944
-
13949
)
118
Palamarchuk
 
A
Efanov
 
A
Nazaryan
 
N
, et al. 
13q14 deletions in CLL involve cooperating tumor suppressors.
Blood
2010
, vol. 
115
 
19
(pg. 
3916
-
3922
)
119
Planelles
 
L
Castillo-Gutiérrez
 
S
Medema
 
JP
Morales-Luque
 
A
Merle-Béral
 
H
Hahne
 
M
APRIL but not BLyS serum levels are increased in chronic lymphocytic leukemia: prognostic relevance of APRIL for survival.
Haematologica
2007
, vol. 
92
 
9
(pg. 
1284
-
1285
)
120
Katsumata
 
M
Siegel
 
RM
Louie
 
DC
, et al. 
Differential effects of Bcl-2 on T and B cells in transgenic mice.
Proc Natl Acad Sci USA
1992
, vol. 
89
 
23
(pg. 
11376
-
11380
)
121
Han
 
YC
Park
 
CY
Bhagat
 
G
, et al. 
microRNA-29a induces aberrant self-renewal capacity in hematopoietic progenitors, biased myeloid development, and acute myeloid leukemia.
J Exp Med
2010
, vol. 
207
 
3
(pg. 
475
-
489
)
122
Pekarsky
 
Y
Santanam
 
U
Cimmino
 
A
, et al. 
Tcl1 expression in chronic lymphocytic leukemia is regulated by miR-29 and miR-181.
Cancer Res
2006
, vol. 
66
 
24
(pg. 
11590
-
11593
)
123
Pekarsky
 
Y
Croce
 
CM
Is miR-29 an oncogene or tumor suppressor in CLL?
Oncotarget
2010
, vol. 
1
 
3
(pg. 
224
-
227
)
124
Gazdar
 
AF
Butel
 
JS
Carbone
 
M
SV40 and human tumours: myth, association or causality?
Nat Rev Cancer
2002
, vol. 
2
 
12
(pg. 
957
-
964
)

Supplemental data

Sign in via your Institution