72
Views
5
CrossRef citations to date
0
Altmetric
Review

Waldenström macroglobulinemia: biology, genetics, and therapy

&
Pages 49-58 | Published online: 26 Jul 2016

Abstract

Waldenström macroglobulinemia (WM) is a distinct clinicopathologic entity characterized by the presence of a lymphoplasmacytic lymphoma, a non-Hodgkin lymphoma, and IgM monoclonal gammopathy. WM is an indolent, uncommon malignancy mostly affecting the elderly. Patient outcomes have modestly improved since the introduction of rituximab to conventional cytotoxic chemotherapy more than 20 years ago. However, the pivotal discovery of the somatic MYD88 L265P mutation, harbored by most patients with WM, and the somatic CXCR4 WHIM mutations, similar to germline CXCR4 mutations seen in the warts, hypogammaglobulinemia, immunodeficiency, and myelokathexis (WHIM) syndrome, present in approximately one-third of patients with WM, has fundamentally changed our understanding of this disease and expanded the potential therapeutic targets. Within this new paradigm, ibrutinib emerged as a promising new drug. Ibrutinib targets Bruton’s tyrosine kinase, a downstream protein in the B-cell receptor pathway that is overactivated by the MYD88 L265P mutation. A seminal Phase II trial of ibrutinib in previously treated WM patients showed impressive response rates and confirmed the effects of MYD88 L265P and CXCR4 WHIM mutations in response to therapy. Ibrutinib is the first and only US Food and Drug Administration–approved drug specifically for the treatment of WM. However, before ibrutinib can be established as the standard of care for WM, long-term data regarding efficacy and safety are required. Further research to address ibrutinib resistance and cost-effectiveness is also imperative before ibrutinib can gain widespread acceptance. This review will cover the present pathophysiologic understanding of WM in light of the recent MYD88 and CXCR4 discovery, as well as current and emergent treatment regimens with focus on ibrutinib.

Introduction

Waldenström macroglobulinemia (WM) was described by Jan Waldenström, a Swedish physician, in 1944.Citation1 The World Health Organization categorizes WM as a subset of low-grade non-Hodgkin lymphoma, characterized by the presence of IgM monoclonal gammopathy and infiltration of the bone marrow by lymphoplasmacytic lymphoma (LPL).Citation2

The clonal lymphoplasmacytic cells in WM usually express global B-cell markers, such as CD19, CD20, CD22, and CD79a, but do not express CD5 (as in mantle cell lymphoma [MCL] and chronic lymphocytic leukemia [CLL]), CD10 (as in follicular lymphoma), or CD56 (as in the majority of multiple myeloma [MM] cases).Citation3Citation5 In rare instances, WM lymphocytes may express CD5, but not as much as in MCL or CLL.Citation6 The phenotype of LPL cells suggests that they are in a late stage of B-cell differentiation, which probably resembles an IgM+ B-cell that has undergone somatic hypermutation but not isotype class switching.Citation7 compares the features of various IgM monoclonal gammopathies.

Table 1 Comparison of IgM monoclonal gammopathiesTable Footnotea

Features required for the diagnosis of WM include 1) presence of an IgM monoclonal gammopathy of any size in serum, 2) bone marrow biopsy findings showing LPL infiltration, and 3) exclusion of other lymphoproliferative disorders, including CLL and MCL. The Mayo Clinic criteria for WM diagnosis require at least 10% marrow involvement by clonal lymphoplasmacytic cells in asymptomatic patients,Citation8 whereas the Second International Workshop on WM eliminated the requirement for a minimum amount of marrow involvement.Citation6

Clinical features

WM represents only 2% of all hematologic malignancies, with an age-adjusted incidence of 3.8 cases per million persons per year.Citation9 WM is predominantly a disease of elderly persons, with a median age at diagnosis ranging from 63 to 73 years in various studies, and a 95-fold higher incidence among octogenarians than among those <50 years.Citation10 Although the estimated median disease-specific survival is 11 years,Citation11 the median overall survival (OS) of Waldenström’s patients is approximately 7 years and reflects associated comorbidities seen in these elderly population.Citation12 Young patients with WM (<50 years) have a longer median OS (13–14.8 years).Citation12,Citation13

Approximately 25% of patients with WM are asymptomatic at the time of diagnosis, but symptoms will develop in up to 70% of them within 10 years of diagnosis.Citation14 The initial presentation of WM is usually nonspecific constitutional symptoms such as fatigue, weight loss, fever, and night sweats. Ultimately, symptoms develop secondary to hematopoietic tissue infiltration by LPL, IgM paraprotein deposition, and/or IgM autoimmune activity.Citation15 The most common clinical features are lymphadenopathy, thrombocytopenia, splenomegaly, anemia, bleeding due to hyperviscosity, and peripheral neuropathy. shows the clinical features of WM.

Table 2 Characteristics of IgM or direct Waldenström macroglobulinemia cell infiltration–related morbidities

In patients with a history of IgM monoclonal gammopathy of undetermined significance (MGUS), a known precursor of WM, the risk of WM is increased by 46-fold.Citation16 IgM MGUS patients carrying the MYD88 L265P mutation have higher risk of progressing to WM than those with wild-type MYD88 (odds ratio 4.7, 95% confidence interval [CI] 0.8–48.7).Citation17 WM has been known to show familial clustering; about 20% of patients with WM have at least one first degree relative with WM or other B-cell disorder.Citation18 There is abundant evidence to support a genetic predisposition in familial cases of WM.Citation19 In a genome-wide analysis of families at high risk for WM, the authors found strong evidence for involved genes on 1q and 4q, and evidence suggestive of gene linkage on 3q and 6q.Citation20 The pattern of affected patients in familial WM, which is usually seen in multiple generations and equally distributed between men and women, favors autosomal dominant or codominant inheritance.

Patients with WM are at increased risk for secondary malignancies. The most common sex-adjusted malignancies are prostate (2%–9.4%), breast (0.9%–8%), nonmelanoma skin cancer (7.1%), melanoma (2.2%), and lung (1.4%–2.1%).Citation19,Citation21 The cumulative incidence of solid cancer was shown to be 6%–9.5% at 5 years, 11%–16.1% at 10 years, and 17% at 15 years in one study.Citation21,Citation22 Hematologic malignancies, mainly diffuse large B-cell lymphoma and myelodysplastic syndrome/acute myeloid leukemia, were seen in 3.7%–5% of patients with WM, but these were mostly in patients previously treated with alkylating agents and nucleoside analogs.Citation21,Citation22 Second malignancies were seen more frequently in sporadic WM than in familial WM. Interestingly, five times more cases of lung cancer were seen in familial WM, and two times more cases of prostate cancer were seen in sporadic WM.Citation19 First-degree relatives of patients with WM are at significantly increased risk for MGUS, WM, or other non-Hodgkin lymphoma.Citation23

Genetic mutations associated with WM

WM has been associated with numerous chromosomal abnormalities and somatic mutations, but no causal relationship has been shown. An abnormality on cytogenetic analysis or fluorescence in situ hybridization is found in about half of patients with WM; the most common are 6q deletion (22%–46%), 13q14 deletion (13%–15%), trisomy 18 (11%–23%), trisomy 4 (4%–12%), and p53 deletion (4%–23%). A rearrangement of the IgH genes as part of a translocation is uncommon in WM, unlike in MM.Citation24,Citation25 Deletion of 6q is common but not specific to WM (it is found in other B-cell disorders), but trisomy 4 is unique to WM and occasionally is the only abnormality seen in these patients.Citation26 No cytogenetic abnormality has been shown to be diagnostic, predictive, or prognostic in WM patients.

The WM tumoral clone shows wide morphologic heterogeneity, from clonal B lymphocytes (BLs) to clonal plasma cells (PCs), and includes a lymphoplasmacytic population. WM-PC shows a pattern of cytogenetic abnormalities and gene rearrangement similar to that in WM-BL, which supports the idea that the WM clonal PC results from an incomplete maturation process of a WM clonal BL. Furthermore, the gene expression profile of PCs and BLs in WM shows complex clones that are similar to their counterparts in CLL and MM, but each still represents a singular entity. Interestingly, the PAX5 gene is significantly upregulated in WM-PC compared with MM-PC; PAX5 expression must be repressed to allow final PC differentiation. The BLIMP1 gene, which is also needed for differentiation from BLs to PCs, is down-regulated in WM-PC.Citation27 This evidence indicates that WM clonal cells are in a later stage of B-cell differentiation into PCs that express IgM, but have not undergone final isotype class switching. In contrast to MM-PC, WM clonal PCs do not express DKK1 and FRZB, which cause bone disease in MM, and may explain the lack of bone lesions in WM.Citation28

The protein MYD88 is an adaptor in the Toll-like receptor and interleukin-1 receptor signaling pathway. The pivotal discovery of the L265P mutation in MYD88 considerably improved our understanding of WM pathogenesis.Citation29 The gain-of-function MYD88 L265P mutation strongly promotes WM cell growth and survival through downstream activation of the above-mentioned pathways, with the transcription of NF-κB.Citation15 The MYD88 L265P mutation has been detected in 86% of sporadic WM cases and in all familial WM cases.Citation29 Even though the MYD88 L265P mutation was seen uniformly in familial WM, a recent study did not support the presence of germline MYD88 mutations in familial WM or IgM MGUS.Citation30

The MYD88 L265P mutation is not exclusive to WM. It can be identified in up to 87% of patients with IgM MGUS and in smaller proportions of patients with activated B-cell type diffuse large B-cell lymphomas (29%), mucosa-associated lymphoid tissue lymphoma (9%), splenic marginal zone lymphoma (10%–21%), and CLL (4%–10%).Citation31Citation33 The role of MYD88 L265P mutation in the development of WM is not clear; it may be a driver mutation in IgM MGUS transformation to WM or may only indicate a subset of patients with IgM MGUS who are predisposed to WM.Citation4,Citation29

In addition to the L265P mutation of MYD88, mutation of the C-X-C chemokine receptor type 4 (CXCR4) was found in 29% of patients with WM.Citation34 The somatic CXCR4 mutations found in WM are similar to germline mutations in CXCR4 described in the rare warts, hypogammaglobulinemia, immunodeficiency, and myelokathexis (WHIM) syndrome. CXCR4 WHIM mutations lead to permanent activation of CXCR4 by its ligand stromal cell-derived factor 1/C-X-C motif chemokine ligand 12 (CXCL12), stimulate the proliferation, migration, and homing of WM cells to bone marrow niches, and therefore promote cell survival.Citation35 Even though CXCR4 WHIM mutations have been associated with more aggressive disease features, such as higher IgM levels and bone marrow involvement, their presence has not affected the OS.Citation36 The CXCR4 WHIM mutations seem to mediate drug resistance and have been shown to affect response to novel treatments, such as Bruton’s tyrosine kinase (BTK), mammalian target of rapamycin, and phosphoinositide 3-kinase (PI3K) inhibitors, as described below.Citation37,Citation38

Risk stratification and prognosis of WM

In patients with IgM MGUS or smoldering WM, which by definition have no disease-related symptoms, the OS is similar to that in the general population of comparable age and sex,Citation39 but these patients carry an increased risk of disease progression to WM. The overall average risk of progression in patients with IgM MGUS is approximately 1.5% per year,Citation40 whereas patients with smoldering WM have an average risk of progression to symptomatic WM of 12% per year for the first 5 years, after which the risk decreases to rates similar to those for IgM MGUS.Citation14 The rate of progression is affected by IgM levels, hemoglobin levels, and degree of bone marrow involvement by LPL. Most patients with smoldering WM with more than 50% marrow involvement will progress to WM within 5 years.Citation14

The International Prognostic Scoring System for WM (IPSSWM) is widely used for stratification of patients with WM. Five adverse parameters are included in the IPSSWM score (age >65 years, hemoglobin ≤11.5 g/dL, platelets ≤100×109/L, β2-microglobulin >3 mg/L, and IgM level >7 g/dL); age >65 years is assigned two points because of its higher prognostic value, and the others are assigned one point. The total score at initiation of therapy stratifies patients into low- (score ≤1), intermediate- (score 2), or high-risk (score ≥3) categories. Five-year survival rates are 87%, 68%, and 36% for low, intermediate, and high risk, respectively.Citation41 The IPSSWM score was independently validated by a large Greek study.Citation42 Increased lactate dehydrogenase level (>250 IU/L) has been shown to further stratify patients with a high-risk IPSSWM score – an increased value was associated with a median OS of 37 months, compared with 104 months for a lower value.Citation42 The presence of MYD88 wild-type was also associated with an increased risk of death in WM patients (hazard ratio 10.6, 95% CI 2.4–46.2).Citation43 In the era of BTK inhibitors, further prognostication can be derived from MYD88 and CXCR4 mutation status as predictors of response to therapy, as discussed below.

Current and emerging treatments

Since WM remains an incurable disease, the aims of treatment are to relieve symptoms and decrease the risk of end-organ damage. Treatment is usually reserved for symptomatic patients and those with severe cytopenias (hemoglobin <10 −g/dL or platelet count <100×109/L). When possible, clinical trials should be considered for patients with newly diagnosed WM or in the relapsed/refractory setting. shows a summary of selected clinical trials in WM.

Table 3 Selected trials in Waldenström macroglobulinemia

Rituximab, a monoclonal anti-CD20 antibody, is widely used in the treatment of WM, either in combination with other agents or as monotherapy. A recent population-based study showed single-agent rituximab as the frontline choice in approximately 55% of WM patients.Citation44 The low toxicity profile and substantial response rates as a single agent make rituximab monotherapy suitable for minimally symptomatic WM patients or those with IgM-related neuropathy, hemolytic anemia, or mild cytopenias. An IgM “flare” is seen in approximately half of patients after starting rituximab.Citation45 Physicians should be aware of this phenomenon as it does not indicate treatment failure, but could aggravate IgM-mediated symptoms/hyperviscosity. Plasma exchange should be considered in patients with symptomatic hyperviscosity before starting rituximab. Single-agent rituximab has consistently shown overall response rates (ORRs) from 50% to 60% in treatment-naïve and relapsed/refractory patients, with a median progression-free survival (PFS) approaching 2 years.Citation46Citation48 The role of rituximab maintenance therapy was assessed in a retrospective study of 238 rituximab-naïve WM patients of whom 68 patients received maintenance therapy. Rituximab maintenance therapy was associated not only with improved median PFS (56 months vs 29 months, P=0.0001) and OS (not reached vs 116 months, P=0.0095) but also with increased infectious complications.Citation49 The superiority of rituximab maintenance therapy has not been evaluated in randomized, prospective studies, therefore not recommended outside clinical trials.

The combination of bendamustine, an alkylating agent, with rituximab in treatment-naïve patients with WM demonstrated an ORR of 100% and complete response (CR) of 53% in a Phase II trial.Citation50 The primary adverse effect was myelosuppression (leukopenia Grade ≥3: 16%; thrombocytopenia Grade ≥3: 3%). A Phase III trial comparing bendamustine plus rituximab (BR) with R-CHOP (rituximab, cyclophosphamide, doxorubicin, vincristine, and prednisone) in patients with indolent lymphomas, including 41 with WM, showed a similar ORR (95%) in both treatment arms. BR, however, had a significantly longer median PFS (70 months vs 28 months for R-CHOP; hazard ratio 0.33; P=0.003), fewer relapses and adverse events.Citation51 Because of these results, BR became a suitable first-line regimen for WM, particularly for patients with bulky disease in need of rapid disease control.

A Phase II trial of dexamethasone, rituximab, and cyclophosphamide (DRC) in treatment-naïve patients with WM demonstrated an ORR of 83%, including a CR rate of 7%. DRC was well tolerated, with a less than 9% rate of Grade 3 or greater adverse reactions.Citation52 The recently reported long-term outcomes of this Phase II trial showed a median PFS of approximately 3 years after DRC and an estimated 10-year OS of 53%.Citation53 DRC remains an active, well-tolerated, and non-stem-cell toxic treatment option in WM.

Proteasome inhibitors have also been shown to be effective against WM. Bortezomib has been studied as monotherapy, in combination with rituximab, and as a triplet regimen with rituximab and dexamethasone. Single-agent bortezomib in treatment of relapsed/refractory WM achieved an ORR of 85%. Responses occurred at a median of 1.4 months, but no CRs were seen.Citation54 The addition of rituximab to weekly bortezomib has been investigated in the treatment-naïveCitation55 and relapsed/refractoryCitation56 settings, showing ORRs of 89% and 81%, as well as CR rates of 4% and 3%, respectively. Median PFS was 16 months in the relapsed/refractory setting. Neuropathy Grade 3 or higher was seen in 5% of patients with this weekly regimen of bortezomib and rituximab. The combination of bortezomib/rituximab plus dexamethasone in treatment-naïve WM patients has yielded ORRs ranging from 85% to 96%, with CR rates as high as 13%. This triplet combination showed a median PFS of 42 months and 7%–30% incidence of Grade 3 or higher neuropathy.Citation57,Citation58

A recent Phase II trial of idelalisib, a selective oral inhibitor of PI3Kδ, in patients with indolent lymphomas (including ten with WM) refractory to alkylating agents and rituximab, showed promising results. The subset of WM patients achieved an ORR of 80% with this oral agent. The most common Grade 3 or higher adverse events were neutropenia (27%), increased liver enzyme values (21%), and diarrhea (13%).Citation59

A Phase I study showed activity of ABT-199, a small-molecule BCL-2 inhibitor recently named venetoclax, in patients with relapsed/refractory non-Hodgkin lymphoma, including four patients with WM who achieved a partial response.Citation60 Preclinical data showed a synergistic effect of ABT-199 with ibrutinib or idelalisib in WM cell models, independent of CXCR4 mutation statusCitation61 and may support its further investigation in combination strategies.

The role of autologous stem cell transplant (ASCT) in WM is not well established, but it was deemed safe and effective in multiple retrospective studies.Citation62Citation65 The largest cohort of patients with WM who underwent ASCT included 158 patients and showed a nonrelapse mortality of only 3.8% at 1 year.Citation64 ASCT achieved an ORR of approximately 95% and a major response rate (MRR) of 78%–80%, with an estimated 5-year OS of approximately 69% and 5-year time-to-next-therapy of 48%.Citation64,Citation65 Survival was affected by number of prior treatment regimens and disease chemosensitivity at the time of transplant.Citation64 Potential ASCT-eligible patients with otherwise long life expectancy could have stem cells harvested and stored early in the disease course before they are exposed to multiple stem-cell toxic treatments.Citation13,Citation64 In contrast to ASCT, allogeneic stem cell transplant is not recommended in patients with WM outside clinical trials due to the prohibitively high transplant-related mortality (23%–44% at 1 year) in this indolent disease.Citation63,Citation66

Comparative safety, efficacy, and tolerability of ibrutinib

BTK, a downstream cytoplasmic enzyme in the MYD88 signaling pathway, has been targeted by several drugs, most notably ibrutinib. The gain-of-function MYD88 L265P mutation, almost ubiquitous in WM patients, is thought to be the main driver of BTK activation in WM, which leads to increased cell survival and proliferation through NF-κB pathways.Citation67

Ibrutinib is an oral, small-molecule, selective, irreversible inhibitor of BTK that triggers apoptosis in WM cells with the MYD88 L265P mutation.Citation68 A Phase I trial of ibrutinib in 56 patients with relapsed/refractory B-cell lymphomas, including four patients with WM, showed an objective response in 60% of the patients and a median PFS of 16 months.Citation69 The encouraging results and acceptable tolerability led to a pivotal Phase II trial of ibrutinib in previously treated WM patients.Citation70 A total of 63 consecutive patients received daily ibrutinib, 420 mg, until disease progression or unacceptable adverse effects. The ORR was 90%, with a median treatment duration of 19 months.

Patients with WM who had the MYD88 L265P mutation and wild-type CXCR4 had the most benefit from ibrutinib, with an ORR of 100% and MRR of 91%; patients with MYD88 L265P and CXCR4 WHIM had less benefit (ORR of 86% and MRR of 62%), and even less benefit was seen for those with wild-type MYD88 and CXCR4 (ORR of 71% and MRR of 28%). These findings highlight BTK’s activity in proliferation of WM cells and survival in the mutated MYD88 L265P signaling pathway and show the drug resistance, possibly the result of bone marrow homing of WM cells, increased by CXCR4 WHIM mutation.Citation70 A later review by the same authors showed that all patients with wild-type MYD88 who had achieved a major response actually harbored a mutation in the MYD88 gene that was not captured by allele-specific polymerase chain reaction analysis for MYD88 L265P. The updated ORR and MRR for WM patients with wild-type MYD88 and CXCR4 were 60% and 0%, respectively.Citation71

The initial rapid and profound decrease in IgM levels (median time to response, 4 weeks) was not matched by a decrease in bone marrow involvement, which suggests that ibrutinib affects IgM secretion rather than cell killing. Improvement in lymphadenopathy (68%), splenomegaly (57%), and peripheral neuropathy (55%) was seen in most patients after initiation of ibrutinib. All patients in whom treatment was initiated for hyperviscosity symptoms required no additional plasma exchange after two cycles of therapy.Citation70

Overall, ibrutinib was well tolerated. Grade 3 or higher neutropenia and thrombocytopenia were seen in 14% and 13% of the patients, respectively. Most of these patients received three or more prior types of therapy. Grade 2 or higher bleeding complications were seen in 6% of the patients, all of which were associated with concomitant use of fish oil supplements. Atrial fibrillation was reported in 5% of patients, all of whom had a history of paroxysmal atrial fibrillation, which resolved after ibrutinib was withdrawn.Citation70 The effect of BTK on the platelet glycoprotein IV/collagen interaction, a critical component of platelet aggregation, is well known and was first identified in patients with X-linked agammaglobulinemia, which is caused by BTK deficiency.Citation72 The bleeding associated with ibrutinib is therefore thought to be secondary to its inhibitory effect on BTK and collagen-mediated platelet aggregation.Citation73 Platelet aggregation normalizes after 2–3 days of ibrutinib cessation.Citation74

Preliminary results of a multicenter Phase III trial demonstrated activity of ibrutinib in rituximab-refractory WM patients. With an ORR rate of 84%, single-agent ibrutinib shows activity in heavily pretreated rituximab-refractory WM patients. Grade 3 or higher adverse events rate was similar to prior studies of single-agent ibrutinib.Citation75

Despite the promising results, CR was not achieved in any patients with WM with the use of ibrutinib.Citation70 Primary and secondary cases of resistance to ibrutinib have also been described. Although ibrutinib resistance is uncommon, its true incidence is unknown given the relatively short follow-up of available clinical trials. Multiple mechanisms have been proposed to explain acquired resistance to BTK inhibitors, including mutations in BTK that interfere with drug binding and alternative pathways that bypass BTK entirely.Citation76 The BTK C481S mutation has been associated with secondary ibrutinib resistance in patients with CLL and MCL. At the structural level, the BTK C481S mutation disrupts covalent binding and allows for reversible, instead of irreversible, binding of ibrutinib by BTK.Citation77,Citation78 Primary resistance to ibrutinib has also been associated with sustained PI3K–Akt activity, rather than BTK activity, in some patients with MCL.Citation78

In WM, primary ibrutinib resistance is seen in patients harboring CXCR4 WHIM mutations. WHIM-like mutations reduce CXCR4 internalization and allow for sustained enzymatic activity of Akt (protein kinase B) and extracellular signal-regulated kinase, with subsequent increased cell survival, migration, and bone marrow homing.Citation79,Citation34 Plerixafor, an anti-CXCR4 agent used for many years as a stem-cell mobilizing agent, has been shown to effectively inhibit the CXCR4–CXCL12 pathway. Its use in patients with WM carrying a CXCR4 WHIM mutation has been proposed as a way to mobilize WM cells from the prosurvival stem cell niche in bone marrow, thus making them more amenable to cytotoxic chemotherapy.Citation80 Several other CXCR4 inhibitors are currently under investigation.Citation81

The use of BTK inhibitors together with CXCR4 antagonists as a way to decrease intrinsic drug resistance in patients with a CXCR4 WHIM mutation is yet to be studied. This combination has the theoretical potential to show even better response rates in WM patients. Furthermore, the patient’s MYD88 and CXCR4 mutation status could be used for a more individualized treatment selection in patients with WM.

Conclusion

Despite expansion of the therapeutic options in recent years, WM remains an incurable disease. Ibrutinib has become the first drug approved by the US Food and Drug Administration specifically for WM, owing to its impressive clinical efficacy, easy route of administration, and favorable toxicity profile reported in a recent Phase II trial.Citation70 Ibrutinib has also been approved in Europe, Canada, Japan, and is under review in Mexico. Ibrutinib as single agent has activity in WM comparable to that with combination therapies; however, no head-to-head trials have been conducted. Furthermore, the data that led to ibrutinib approval by the US Food and Drug Administration are based only on response rates. Long-term data are required to ascertain end points such as PFS, OS, resistance rates, long-term complications, and quality of life before a definitive conclusion on the role of ibrutinib in WM treatment can be made.

Nonetheless, ibrutinib has inaugurated a new era in the treatment of WM. Clinical trials of ibrutinib in combination with other established or novel agents are under way, and results could further solidify the role of ibrutinib in WM. Additionally, a myriad of strategies has been proposed to overcome ibrutinib resistance, such as ABT-199 (BLC2-ihibitor),Citation82 RP6530 (dual PI3Kδ/γ inhibitor),Citation83 and plerixafor (CXCR4 inhibitor).Citation84 The treatment of WM continues to evolve and expand rapidly, with the potential to affect the course of this disease.

Disclosure

The authors report no conflicts of interest in this work.

References

  • WaldenstromJIncipient myelomatosis or essential hyperglobulinemia with fibrinogenopenia – a new syndrome?Acta Med Scand194411734216247
  • CampoESwerdlowSHHarrisNLPileriSSteinHJaffeESThe 2008 WHO classification of lymphoid neoplasms and beyond: evolving concepts and practical applicationsBlood2011117195019503221300984
  • LinPMedeirosLJLymphoplasmacytic lymphoma/Waldenstrom macroglobulinemia: an evolving conceptAdv Anat Pathol200512524625516210920
  • KapoorPPaludoJAnsellSMWaldenstrom macroglobulinemia: familial predisposition and the role of genomics in prognosis and treatment selectionCurr Treat Options Oncol20161731626942591
  • San MiguelJFVidrialesMBOcioEImmunophenotypic analysis of Waldenstrom’s macroglobulinemiaSemin Oncol200330218719512720134
  • OwenRGTreonSPAl-KatibAClinicopathological definition of Waldenstrom’s macroglobulinemia: consensus panel recommendations from the Second International Workshop on Waldenstrom’s MacroglobulinemiaSemin Oncol200330211011512720118
  • KriangkumJTaylorBJTreonSPMantMJBelchARPilarskiLMClonotypic IgM V/D/J sequence analysis in Waldenstrom macroglobulinemia suggests an unusual B-cell origin and an expansion of polyclonal B cells in peripheral bloodBlood200410472134214214764523
  • AnsellSMKyleRAReederCBDiagnosis and management of Waldenstrom macroglobulinemia: Mayo stratification of macroglobu-linemia and risk-adapted therapy (mSMART) guidelinesMayo Clin Proc201085982483320702770
  • WangHChenYLiFTemporal and geographic variations of Waldenstrom macroglobulinemia incidenceCancer2012118153793380022139816
  • WangHChenYLiFTemporal and geographic variations of Waldenstrom macroglobulinemia incidence: a large population-based studyCancer2012118153793380022139816
  • GhobrialIMFonsecaRGertzMAPrognostic model for disease-specific and overall mortality in newly diagnosed symptomatic patients with Waldenstrom macroglobulinaemiaBr J Haematol2006133215816416611306
  • CastilloJJOlszewskiAJKananSMeidKHunterZRTreonSPOverall survival and competing risks of death in patients with Waldenstrom macroglobulinaemia: an analysis of the Surveillance, Epidemiology and End Results databaseBr J Haematol20151691818925521528
  • VallumsetlaNPaludoJAnsellSMOutcomes of young patients with Waldenstrom macroglobulinemia (WM)J Clin Oncol (Meeting Abstracts)20143215_suppl Abstract number 8609
  • KyleRABensonJTLarsonDRProgression in smoldering Waldenström macroglobulinemia: long-term resultsBlood2012119194462446622451426
  • KapoorPPaludoJVallumsetlaNGreippPRWaldenstrom macroglobulinemia: what a hematologist needs to knowBlood Rev201529530131925882617
  • KyleRATherneauTMRajkumarSVA long-term study of prognosis in monoclonal gammopathy of undetermined significanceN Engl J Med2002346856456911856795
  • VarettoniMArcainiLZibelliniSPrevalence and clinical significance of the MYD88 (L265P) somatic mutation in Waldenstrom’s macroglobulinemia and related lymphoid neoplasmsBlood2013121132522252823355535
  • TreonSPHunterZRAggarwalACharacterization of familial Waldenstrom’s macroglobulinemiaAnn Oncol200617348849416357024
  • HanzisCOjhaRPHunterZAssociated malignancies in patients with Waldenstrom’s macroglobulinemia and their kinClin Lymphoma Myeloma Leuk2011111889221454200
  • McMasterMLGoldinLRBaiYGenomewide linkage screen for Waldenström macroglobulinemia susceptibility loci in high-risk familiesAm J Hum Genet200679469570116960805
  • CastilloJJOlszewskiAJHunterZRKananSMeidKTreonSPIncidence of secondary malignancies among patients with Waldenstrom macroglobulinemia: an analysis of the SEER databaseCancer2015121132230223625757851
  • MorraEVarettoniMTedeschiAAssociated cancers in Waldenstrom macroglobulinemia: clues for common genetic predispositionClin Lymphoma Myeloma Leuk201313670070324070824
  • KristinssonSYLandgrenOWhat causes Waldenstrom’s macroglobulinemia: genetic or immune-related factors, or a combination?Clin Lymphoma Myeloma Leuk2011111858721454199
  • Nguyen-KhacFLambertJChapiroEChromosomal aberrations and their prognostic value in a series of 174 untreated patients with Waldenstrom’s macroglobulinemiaHaematologica201398464965423065509
  • BraggioEPhilipsbornCNovakAHodgeLAnsellSFonsecaRMolecular pathogenesis of Waldenstrom’s macroglobulinemiaHaematologica20129791281129022773606
  • BraggioEFonsecaRGenomic abnormalities of Waldenstrom macroglobulinemia and related low-grade B-cell lymphomasClin Lymphoma Myeloma Leuk201313219820123477936
  • Shapiro-ShelefMCalameKRegulation of plasma-cell developmentNat Rev Immunol20055323024215738953
  • GutierrezNCOcioEMde las RivasJGene expression profiling of B lymphocytes and plasma cells from Waldenstrom’s macroglobulinemia: comparison with expression patterns of the same cell counterparts from chronic lymphocytic leukemia, multiple myeloma and normal individualsLeukemia200721354154917252022
  • TreonSPXuLYangGMYD88 L265P somatic mutation in Waldenstrom’s macroglobulinemiaN Engl J Med2012367982683322931316
  • PertesiMGaliaPNazaretNRare circulating cells in familial Waldenstrom macroglobulinemia displaying the MYD88 L265P mutation are enriched by Epstein-Barr virus immortalizationPLoS ONE2015109e013650526352266
  • XuLHunterZRYangGMYD88 L265P in Waldenstrom macroglobulinemia, immunoglobulin M monoclonal gammopathy, and other B-cell lymphoproliferative disorders using conventional and quantitative allele-specific polymerase chain reactionBlood2013121112051205823321251
  • NgoVNYoungRMSchmitzROncogenically active MYD88 mutations in human lymphomaNature2011470733211511921179087
  • JimenezCSebastianEChillonMCMYD88 L265P is a marker highly characteristic of, but not restricted to, Waldenstrom’s macroglobulinemiaLeukemia20132781722172823446312
  • HunterZRXuLYangGThe genomic landscape of Waldenstrom macroglobulinemia is characterized by highly recurring MYD88 and WHIM-like CXCR4 mutations, and small somatic deletions associated with B-cell lymphomagenesisBlood2014123111637164624366360
  • NgoHTLeleuXLeeJSDF-1/CXCR4 and VLA-4 interaction regulates homing in Waldenstrom macroglobulinemiaBlood2008112115015818448868
  • CaoYHunterZRLiuXCXCR4 WHIM-like frameshift and nonsense mutations promote ibrutinib resistance but do not supplant MYD88(L265P)-directed survival signalling in Waldenstrom macroglobulinaemia cellsBr J Haematol2015168570170725371371
  • RoccaroAMSaccoAJimenezCC1013G/CXCR4 acts as a driver mutation of tumor progression and modulator of drug resistance in lymphoplasmacytic lymphomaBlood2014123264120413124711662
  • TreonSTripsasCMeidKIbrutinib in previously treated patients with Waldenstrom’s macroglobulinemia is highly active, produces durable responses, and is impacted by MYD88 and CXCR4 mutation statusHaematologica2015100311
  • VajdicCMLandgrenOMcMasterMLMedical history, lifestyle, family history, and occupational risk factors for lymphoplasmacytic lymphoma/Waldenstrom’s macroglobulinemia: the InterLymph Non-Hodgkin Lymphoma Subtypes ProjectJ Natl Cancer Inst Monogr2014201448879725174029
  • KyleRATherneauTMRajkumarSVLong-term follow-up of IgM monoclonal gammopathy of undetermined significanceBlood2003102103759376412881316
  • MorelPDuhamelAGobbiPInternational prognostic scoring system for Waldenstrom macroglobulinemiaBlood2009113184163417019196866
  • KastritisEKyrtsonisMCHadjiharissiEValidation of the International Prognostic Scoring System (IPSS) for Waldenstrom’s macroglobulinemia (WM) and the importance of serum lactate dehydrogenase (LDH)Leuk Res201034101340134320447689
  • TreonSPCaoYXuLYangGLiuXHunterZRSomatic mutations in MYD88 and CXCR4 are determinants of clinical presentation and overall survival in Waldenstrom macroglobulinemiaBlood2014123182791279624553177
  • OlszewskiAJFallahJEatonCBTreonSPCastilloJJThe evolution of management and survival outcomes of Waldenström macroglobulinemia (WM) in the United States (US)Blood201512623882882
  • GhobrialIMFonsecaRGreippPRInitial immunoglobulin M ‘flare’ after rituximab therapy in patients diagnosed with Waldenstrom macroglobulinemiaCancer2004101112593259815493038
  • GertzMAAbonourRHeffnerLTGreippPRUnoHRajkumarSVClinical value of minor responses after 4 doses of rituximab in Waldenstrom macroglobulinaemia: a follow-up of the Eastern Cooperative Oncology Group E3A98 trialBr J Haematol2009147567768019751237
  • TreonSPEmmanouilidesCKimbyEExtended rituximab therapy in Waldenstrom’s macroglobulinemiaAnn Oncol200516113213815598950
  • GertzMARueMBloodEKaminerLSVesoleDHGreippPRMulticenter phase 2 trial of rituximab for Waldenstrom macroglobulinemia (WM): an Eastern Cooperative Oncology Group Study (E3A98)Leuk Lymphoma200445102047205515370249
  • TreonSPHanzisCManningRJMaintenance rituximab is associated with improved clinical outcome in rituximab naïve patients with Waldenstrom macroglobulinaemia who respond to a rituximab-containing regimenBr J Haematol2011154335736221615385
  • RummelMJAl-BatranSEKimSZBendamustine plus rituximab is effective and has a favorable toxicity profile in the treatment of mantle cell and low-grade non-Hodgkin’s lymphomaJ Clin Oncol200523153383338915908650
  • RummelMJNiederleNMaschmeyerGBendamustine plus rituximab versus CHOP plus rituximab as first-line treatment for patients with indolent and mantle-cell lymphomas: an open-label, multicentre, randomised, phase 3 non-inferiority trialLancet201338198731203121023433739
  • DimopoulosMAAnagnostopoulosAKyrtsonisMCPrimary treatment of Waldenstrom macroglobulinemia with dexamethasone, rituximab, and cyclophosphamideJ Clin Oncol200725223344334917577016
  • KastritisEGavriatopoulouMKyrtsonisMCDexamethasone, rituximab, and cyclophosphamide as primary treatment of Waldenstrom macroglobulinemia: final analysis of a phase 2 studyBlood2015126111392139426359434
  • TreonSPHunterZRMatousJMulticenter clinical trial of bortezomib in relapsed/refractory Waldenstrom’s macroglobulinemia: results of WMCTG Trial 03-248Clin Cancer Res200713113320332517545538
  • GhobrialIMXieWPadmanabhanSPhase II trial of weekly bortezomib in combination with rituximab in untreated patients with Waldenstrom macroglobulinemiaAm J Hematol201085967067420652865
  • GhobrialIMHongFPadmanabhanSPhase II trial of weekly bortezomib in combination with rituximab in relapsed or relapsed and refractory Waldenstrom macroglobulinemiaJ Clin Oncol20102881422142820142586
  • DimopoulosMAGarcia-SanzRGavriatopoulouMPrimary therapy of Waldenstrom macroglobulinemia (WM) with weekly bortezomib, low-dose dexamethasone, and rituximab (BDR): long-term results of a phase 2 study of the European Myeloma Network (EMN)Blood2013122193276328224004667
  • TreonSPIoakimidisLSoumeraiJDPrimary therapy of Waldenstrom macroglobulinemia with bortezomib, dexamethasone, and rituximab: WMCTG clinical trial 05-180J Clin Oncol200927233830383519506160
  • GopalAKKahlBSde VosSPI3Kdelta inhibition by idelalisib in patients with relapsed indolent lymphomaN Engl J Med2014370111008101824450858
  • GerecitanoJFRobertsAWSeymourJFA Phase 1 study of venetoclax (ABT-199 / GDC-0199) monotherapy in patients with relapsed/refractory non-Hodgkin lymphomaBlood201512623254
  • CaoYYangGHunterZRThe BCL2 antagonist ABT-199 triggers apoptosis, and augments ibrutinib and idelalisib mediated cytotoxicity in CXCR4 wild-type and CXCR4 WHIM mutated Waldenstrom macroglobulinaemia cellsBr J Haematol2015170113413825582069
  • AnagnostopoulosAHariPNPerezWSAutologous or allogeneic stem cell transplantation in patients with Waldenstrom’s macroglobulinemiaBiol Blood Marrow Transplant200612884585416864055
  • GilleeceMHPearceRLinchDCThe outcome of haemopoietic stem cell transplantation in the treatment of lymphoplasmacytic lymphoma in the UK: a British Society Bone Marrow Transplantation studyHematology200813211912718616880
  • KyriakouCCanalsCSibonDHigh-dose therapy and autologous stem-cell transplantation in Waldenstrom macroglobulinemia: the Lymphoma Working Party of the European Group for Blood and Marrow TransplantationJ Clin Oncol201028132227223220368570
  • PaludoJGertzMAnsellSImpact of day-100 response post autologous stem cell transplantation (ASCT) in Waldenstrom macroglobulinemia (WM)Biol Blood Marrow Transplant2016223S130
  • KyriakouCCanalsCCornelissenJJAllogeneic stem-cell transplantation in patients with Waldenstrom macroglobulinemia: report from the Lymphoma Working Party of the European Group for Blood and Marrow TransplantationJ Clin Oncol201028334926493420956626
  • BuggyJJEliasLBruton tyrosine kinase (BTK) and its role in B-cell malignancyInt Rev Immunol201231211913222449073
  • CameronFSanfordMIbrutinib: first global approvalDrugs201474226327124464309
  • AdvaniRHBuggyJJSharmanJPBruton tyrosine kinase inhibitor ibrutinib (PCI-32765) has significant activity in patients with relapsed/refractory B-cell malignanciesJ Clin Oncol2013311889423045577
  • TreonSPTripsasCKMeidKIbrutinib in previously treated Waldenstrom’s macroglobulinemiaN Engl J Med2015372151430144025853747
  • TreonSPXuLHunterZMYD88 mutations and response to ibrutinib in Waldenstrom’s macroglobulinemiaN Engl J Med20153736584586
  • QuekLSBolenJWatsonSPA role for Bruton’s tyrosine kinase (Btk) in platelet activation by collagenCurr Biol1998820113711409778529
  • KamelSHortonLYsebaertLIbrutinib inhibits collagen-mediated but not ADP-mediated platelet aggregationLeukemia201529478378725138588
  • YsebaertLLevadeMCedricGElucidation of mild bleeding disorders reported under ibrutinib (Imbruvica(R)) therapy: implications for optimal clinical managementBlood20141242132963296
  • DimopoulosMATrotmanJTedeschiAIbrutinib therapy in rituximab-refractory patients with Waldenström’s macroglobulinemia: initial results from an international, multicenter, open-label phase 3 substudy (iNNOVATETM)Blood20151262327452745
  • ZhangSQSmithSMZhangSYLynn WangYMechanisms of ibrutinib resistance in chronic lymphocytic leukaemia and non-Hodgkin lymphomaBr J Haematol2015170444545625858358
  • FurmanRRChengSLuPIbrutinib resistance in chronic lymphocytic leukemiaN Engl J Med2014370242352235424869597
  • ChironDDi LibertoMMartinPCell-cycle reprogramming for PI3K inhibition overrides a relapse-specific C481S BTK mutation revealed by longitudinal functional genomics in mantle cell lymphomaCancer Discov2014491022103525082755
  • CaoYHunterZRLiuXThe WHIM-like CXCR4(S338X) somatic mutation activates AKT and ERK, and promotes resistance to ibrutinib and other agents used in the treatment of Waldenstrom’s macroglobulinemiaLeukemia201529116917624912431
  • BeiderKDarash-YahanaMBlaierOCombination of imatinib with CXCR4 antagonist BKT140 overcomes the protective effect of stroma and targets CML in vitro and in vivoMol Cancer Ther20141351155116924502926
  • ScalaSMolecular pathways: targeting the CXCR4-CXCL12 axis-untapped potential in the tumor microenvironmentClin Cancer Res201521194278428526199389
  • CaoYYangGHunterZRThe BCL2 antagonist ABT-199 triggers apoptosis, and augments ibrutinib and idelalisib mediated cytotoxicity in CXCR4 wild-type and CXCR4 WHIM mutated Waldenstrom macroglobulinaemia cellsBr J Haematol2015170113413825582069
  • VakkalankaSNyayapathySViswanadhaSAddition of RP6530, a dual PI3Kδ/γ inhibitor, overcomes ibrutinib resistance in DLBCL cells in vitroBlood20141242144974497
  • McDermottDHLiuQVelezDA phase 1 clinical trial of long-term, low-dose treatment of WHIM syndrome with the CXCR4 antagonist plerixaforBlood2014123152308231624523241