In the past 36 months, new developments have occurred both in the understanding of the biology of Waldenström macroglobulinemia (WM) and in therapeutic options for WM. Here, we review the classification, clinical features, and diagnostic criteria of the disease. WM is a B-cell neoplasm characterized by lymphoplasmacytic infiltration of the bone marrow and a monoclonal immunoglobulin M (IgM) protein. The symptoms of WM are attributable to the extent of tumor infiltration and to elevated IgM levels. The most common symptom is fatigue attributable to anemia. The prognostic factors predictive of survival include the patient's age, β2-microglobulin level, monoclonal protein level, hemoglobin concentration, and platelet count. Therapy is postponed for asymptomatic patients, and progressive anemia is the most common indication for initiation of treatment. The main therapeutic options include alkylating agents, nucleoside analogues, and rituximab. Studies involving combination chemotherapy are ongoing, and preliminary results are encouraging. No specific agent or regimen has been shown to be superior to another for treatment of WM. Novel agents such as bortezomib, perifosine, atacicept, oblimersen sodium, and tositumomab show promise as rational targeted therapy for WM.

Nearly 63 years have passed since Jan Gosta Waldenström first described 2 patients with oronasal bleeding, lymphadenopathy, anemia and thrombocytopenia, elevated erythrocyte sedimentation rate, high serum viscosity, normal bone radiographs, and bone marrow showing predominantly lymphoid cells.1  At a time when paper electrophoresis was unheard of, he attributed hyperviscosity symptoms to an abnormal high-molecular-weight serum protein. These preliminary observations proved to be the cornerstones of the widely recognized but relatively uncommon diagnosis of Waldenström macroglobulinemia (WM). The abnormal high-molecular-weight serum protein subsequently was shown to be monoclonal immunoglobulin M (IgM).

WM is a malignant lymphoplasmo-proliferative disorder with monoclonal pentameric IgM production. The most consistent feature of the bone marrow or lymph nodes of patients with WM is the presence of pleomorphic B-lineage cells at different stages of maturation, such as small lymphocytes, lymphoplasmacytoid cells (abundant basophilic cytoplasm but lymphocyte-like nuclei), and plasma cells.2  Bone marrow is infiltrated in a predominantly intertrabecular pattern. A significant increase in the number of mast cells has been noted in bone marrow biopsies of WM patients.3  Bone marrow mast cells of WM patients overexpress the CD40 ligand (CD154), which is a potent inducer of B-cell expansion.

WM was initially defined in broad terms with the Kiel classification4  as a lymphoma of Ig-secreting cells, often associated with a paraproteinemia. Subsequently, WM was combined with lymphoplasmacytic lymphoma (LPL) and designated LPL/WM in the Revised European-American Lymphoma5  and World Health Organization (WHO)6  classifications. The consensus group at the Second International Workshop on WM in 20027  redefined WM as a distinct clinicopathologic entity characterized by bone marrow infiltration by LPL and IgM monoclonal gammopathy.

WM has an overall incidence of approximately 3 per million persons per year, accounting for approximately 1% to 2% of hematologic cancers.8,9  The incidence of WM is higher among whites, with blacks representing only 5% of all patients.10  In large series of patients, the median age varies between 63 and 68 years, with 55% to 70% men.11  WM remains incurable, and most patients die of disease progression, with a median survival of 5 years.12  Because of the late age of presentation of WM, half of the patients succumb to causes unrelated to WM.

WM is believed to be predominantly a sporadic disease. Its cause is unknown, but various reports of multigenerational clustering and familial patterns indicate the possible role of a single genetic defect.13  Treon et al14  analyzed 257 consecutive and unrelated patients with WM: 48 (18.7%) had at least 1 first-degree relative with either WM or another B-cell disorder. Moreover, patients with a familial history of WM or a plasma cell disorder received the diagnosis at a younger age and with greater bone marrow involvement. Deletions in 6q21-22.1 were confirmed in most WM patients regardless of family history. In short, a high degree of clustering of B-cell disorders was seen among first-degree relatives of patients with WM.

The main risk factor for the development of WM is pre-existing IgM–monoclonal gammopathy of undetermined significance (MGUS) (46 times higher relative risk than for the general population).15  Morra et al16  showed a progressive increase in the risk of transformation from asymptomatic IgM-MGUS to symptomatic WM, with increasing IgM levels.

A possible association between hepatitis C virus and WM had been suggested,17  but this has been negated recently by Leleu et al.18  Reports of links between human herpesvirus-8 and WM are unconfirmed.

The origin of the B-cell clone has been long debated. Distinct B-cell subsets have been demonstrated in the bone marrow, marginal zone of the spleen and lymph nodes, and circulating in the peripheral blood.19  Analysis of 14q32 by fluorescence in situ hybridization (FISH) and Southern blot indicates the absence of Ig heavy chain (IgH) rearrangements in WM.20  Postswitch clonotypic Ig (IgG or IgA) is undetectable in WM B cells, confirming the absence of isotype switch events by deletional recombination.21–23  With the use of variable region (V) gene analysis, evidence shows that VH genes are somatically mutated in WM. Most WM VDJ sequences from the VH3/JH4 gene families are hypermutated and lack intraclonal heterogeneity.21,23  In contrast, one case has been reported in which WM cells have shown functional class switch recombination.24  This favors the hypothesis that WM cells have normal class switch recombination machinery but defective initiation of the switching process.

The only recurrent abnormality identified by Schop et al20,25  was deletion of the long arm of chromosome 6 in 55% of cases. In contrast, FISH analysis by Terre et al26  on 39 WM cases found 6q deletions in only 21%, and trisomy 4 was the most recurrent chromosomal abnormality (18%). In a study involving 37 patients with LPL/WM, Mansoor et al27  reported that the most common chromosomal numeric abnormalities were trisomy 5 and monosomy 8 in 3 cases each; the most common structural abnormality was deletion of 6q in 6 cases.

Liu et al28  reported a single case of WM with deletion of 20q as the sole initial cytogenetic abnormality. Ten cases with plasma cell dyscrasias and del(20q) were reviewed. None of the cases without genotoxic chemotherapy exposure had development of myelodysplastic syndrome/acute myelogenous leukemia during follow-up. This suggests that the significance of del(20q) differs depending on whether it appears at diagnosis or after chemotherapy.

Of the various genes that have been localized to 6q21, BLIMP-1 is postulated to be of importance in WM. BLIMP-1, a tumor-suppressor gene, is the master gene regulator for B-lymphocytic cell proliferation and differentiation.29–31  It facilitates the transition from the mature B-cell stage to the plasma cell stage. Partial or whole losses in this gene could result in the predisposition for B-cell malignancies such as WM.

Another area of interest in WM is the B-lymphocyte stimulator (BLyS), also known as B-cell–activating factor of the tumor necrosis factor family.32  BLyS is expressed on monocytes and is critical for maintenance of normal B-cell development and homeostasis.33,34  It is overexpressed in various B-cell malignancies and has been shown to inhibit apoptosis in malignant B cells. Moreover, in a study by Elsawa et al,35  lymphoplasmacytic cell infiltrates in the bone marrow of patients with WM stained positive for BLyS expression, and serum BLyS levels in patients with WM were significantly higher than in healthy controls. Because of the role of BLyS in WM, strategies to inhibit BLyS potentially may have therapeutic efficacy.

TACI (transmembrane activator and calcium-modulator and cyclophilin ligand interactor), a TNF-receptor family member expressed on B lymphocytes, has been shown to have a high affinity for APRIL (a proliferation-inducing ligand) and BLyS.36  Mutations in TACI signaling are commonly seen in common variable immune deficiency. In WM, IgG and IgA hypogammaglobulinemia are more prevalent among cases with mutations in the TACI signaling process.37 

Abnormal expression of hyaluronan synthases (HASs) has been reported as a possible pathogenetic factor in WM.38  Hyaluronan has a role in malignant cell migration and metastasis. Of the 3 HAS isoenzymes detected in humans, Adamia et al38  postulated that overexpressed HAS1 and HAS3 form a hyaluronan matrix around WM cells, thereby preventing their elimination by the immune system and promoting spread of the disease. A later report stated that a single nucleotide polymorphism in the HAS1 gene resulted in an enhanced risk of WM development.39  These findings are important considering that serum hyaluronan levels have been shown to have prognostic value in multiple myeloma.

Serum interleukin-6 (IL-6) levels have been reported to reflect disease severity and high tumor burden in patients with WM. It has also been shown that clonal blood B cells from patients with WM spontaneously differentiate in vitro to plasma cells through an IL-6 pathway. These findings suggest that IL-6 may be a marker reflecting tumor burden and response to treatment in WM.40 

The typical immunophenotype of WM consists of expression of pan–B-cell surface markers (CD19, CD20, CD22), cytoplasmic Igs, FMC7, BCL2, PAX5, CD38, and CD79a; CD10 and CD23 are mostly absent, and CD5 is expressed in 5% to 20% of cases.41  Variations in the immunophenotypic profile exist, but most patients meet the newly proposed criteria: monoclonal surface Ig positive (5:1 κ/λ ratio), CD19+, CD20+, CD5, CD10, and CD23.7,42  CD5 positivity does not rule out the diagnosis of WM.

According to the WHO classification, WM cells lack CD23,6  and the 2002 WM consensus group stated that CD23 expression is found in a minority of patients.7  However, in a retrospective study of the immunophenotypic profile of 75 patients by Konoplev et al,43  61% of LPL/WM cases were CD23+. As part of the 2004 consensus conference on WM, Hunter et al44  reported the detection of CD23 in 35% of cases.

The immunophenotypic profile in combination with the presence of somatic mutations of V genes (IgM) without intraclonal diversity strongly suggests that the malignant cells in WM originate from cells at a late stage of differentiation.22  The clonal population in WM appears to be derived from a B cell arrested after somatic hypermutation in the germinal center and before terminal differentiation to a plasma cell.45 

Most patients with the diagnosis of WM have symptoms attributable to tumor infiltration or monoclonal serum protein (or both).

Symptoms attributable to tumor infiltration

Extensive bone marrow infiltration leads to cytopenias, and progressive anemia is the most common indication for initiation of treatment. Lytic bone disease is very uncommon in WM. Although the neoplastic clone predominantly infiltrates the bone marrow, it can also infiltrate other organs, including lymph nodes, liver, and spleen, presenting as organomegalies.46  In rare cases, diffuse lymphoplasmacytic infiltration of the pulmonary parenchyma can occur, and patients present with diffuse pulmonary infiltrates, nodules, or pleural effusion.47  Malignant infiltration of the stomach and bowel as well as the skull base and orbit has also been reported.48  Bing-Neel syndrome (long-standing hyperviscosity altering the vascular permeability and leading to perivascular malignant infiltration) consists of headache, vertigo, impaired hearing, ataxia, nystagmus, diplopia, and eventually coma.49  Malignant vitreitis and conjunctival infiltration are rare ocular manifestations of the disease.50 

Symptoms attributable to circulating IgM

The larger size and increased concentration of the monoclonal protein leads to an increase in vascular resistance and viscosity. Serum hyperviscosity is the most distinguishing feature of WM, but it is only observed in less than 15% of patients at diagnosis.51  Symptoms of hyperviscosity usually appear when the normal serum viscosity of 1.4 to 1.8 cP reaches 4 to 5 cP (corresponding to a serum IgM level of at least 30 g/L [3 g/dL]) and include constitutional symptoms, bleeding, and ocular, neurologic, and cardiovascular manifestations.52  High-output cardiac failure may develop because of the expanded plasma volume arising from increased osmotic pressure. Abnormalities in bleeding and clotting times occur from the interaction of IgM with coagulation factors. IgM may also coat platelets, thereby impairing their function. The circulating IgM also may undergo precipitation at cooler temperatures and present as type I or type II cryoglobulinemia. Cryoglobulins may be detected in 20% of patients, but fewer than 5% present with symptoms such as Raynaud syndrome, arthralgia, purpura, and skin ulcers.53  Priapism has also been described as an unusual complication.

Symptoms attributable to tissue deposition of IgM

IgM deposition can occur in glomerular loops, intestine, and skin, presenting as proteinuria, diarrhea, and macroglobulinemia cutis (papules and nodules), respectively. Primary amyloidosis due to deposition of monoclonal light chains occurs mainly in the heart, peripheral nerves, kidneys, soft tissues, liver, and lungs (in descending order of frequency).54  Secondary amyloidosis is seen rarely in WM, with nephrotic syndrome and gastrointestinal symptoms as the initial presentation.55  Acute renal failure is rare in WM; in most cases, there is slowly progressive loss of function.56  Although most patients have detectable light chains in the urine, renal insufficiency and cast nephropathy are rare.11 

Symptoms attributable to autoantibody activity of IgM

The IgM protein has been proved to induce various autoimmune symptoms in WM. In fewer than 10% of patients, the IgM κ reacts with specific red blood cell antigens at temperatures below 37°C to produce a chronic immune hemolytic anemia, which is associated with elevated cold agglutinin titers.57  Schnitzler syndrome is the term for IgM monoclonal gammopathy associated with urticarial skin lesions, fever, and arthralgia.58,59  Peripheral neuropathies have been reported in 15% to 30% of patients with IgM-MGUS or WM.60  The most commonly encountered symptomatic neuropathy in WM is symmetric polyneuropathy; other forms include cranial nerve palsies and mononeuropathies or multineuropathies.61  In a study of 119 WM patients and 58 controls by Levine et al,62  polyneuropathy symptoms were observed more frequently in patients with WM (47%) than in controls (9%) (P <.001). Other less common neuropathies associated with WM include those related to amyloidosis and cryoglobulinemia. Anti–myelin-associated glycoprotein (MAG) antibody has been implicated in the demyelinating neuropathy found in WM.63  Analyses of nerve tissue from patients with anti-MAG antibodies in nerve or skin biopsy specimens have demonstrated IgM deposits at the site of MAG localization. The antibody properties of IgM toward glomerular basement membrane may present as glomerulonephritis. Angioedema and acquired von Willebrand disease have also been reported.

The presence of clonal B cells with lymphoplasmacytic differentiation in the bone marrow or a serum monoclonal IgM protein are not pathognomonic for WM and may be seen in other B-cell lymphoproliferative disorders including splenic marginal zone lymphoma (SMZL). With increasing frequency, patients who fulfill the diagnostic criteria of WM are being diagnosed without having any symptoms or signs and are classified as having asymptomatic or smoldering WM.64 

Asymptomatic patients with monoclonal IgM and without morphologic evidence of bone marrow infiltration (< 10% clonal marrow cells) are classified as having IgM-MGUS, which is the most common differential diagnosis for patients with an IgM monoclonal protein. Some patients may have detectable bone marrow clonal B cells by flow cytometry but no morphologic evidence of bone marrow infiltration at trephine biopsy. These patients should be classified as having IgM-MGUS and monitored without therapeutic intervention.65  Results from FISH studies indicate that deletion of the long arm of chromosome 6 (6q−) is not seen in IgM-MGUS, and 6q− has been suggested as a clinical marker to distinguish WM from IgM-MGUS.25 

SMZL can be distinguished from WM on the basis of immunophenotypic and molecular cytogenetic studies. Ocio et al66  demonstrated that CD22 and CD11c were overexpressed in patients with SMZL, whereas CD25 was more common in WM (88% vs 44%). The CD103 antigen (which was always negative in WM) was positive in 40% of SMZL cases. The chromosomal abnormality most commonly seen in WM was 6q deletion, whereas in SMZL, it was the loss of 7q along with +3q and +5q.

The clinical differentiation of multiple myeloma (MM) from WM is straightforward. When a patient presents with features typical of MM and an IgM component, a diagnosis of IgM-MM is made. The distinction between IgM-MM and WM is based on the pure plasma cell morphology in myeloma and presence of lytic bone lesions in myeloma. Renal insufficiency is more common in IgM-MM than in WM. Typical WM expresses all the B-cell antigens (CD19, CD20, and CD22), whereas IgM-MM typically expresses plasma cell antigens CD38 and CD138, which are absent in WM. IgH gene translocations are more common in IgM-MM,67  particularly t(11;14)(q13;q32).

B-cell chronic lymphocytic leukemia (CLL) may mimic WM clinically. The most common physical finding in CLL is lymphadenopathy. Morphology and immunophenotyping are adequate to diagnose CLL. Lymphocytes are typically small and mature, without visible nucleoli, and smudge cells are characteristic.68  The lymphocytes in CLL are positive for CD5 and CD23, whereas both are usually negative in WM. The presence of strong cytoplasmic Ig in WM also helps in making the distinction.6  Patients with CLL frequently have IgM-MGUS.69 

The evolution of WM to diffuse large B-cell lymphoma (DLBCL) as a result of histologic transformation has been described.70  Onset of DLBCL is usually characterized by an aggressive clinical course and usually manifests as worsening constitutional symptoms, profound cytopenias, extramedullary disease, and organomegaly. It is also associated with a poor outcome. The clinicopathologic features at diagnosis of WM do not predict the risk of DLBCL.70  Ig light chain expression is usually identical in WM and DLBCL.

In practice, a surface IgM-positive CD5CD10CD19+CD20+CD23 immunophenotype in association with a nonparatrabecular pattern of infiltration is diagnostic of WM.41,71  The diagnostic criteria as stated by Owen71  are summarized in Table 1. The various diagnostic studies required in daily practice for suspected cases of WM are summarized in Table 2.

The median survival of patients with WM ranges between 5 and 10 years in different series. Several studies have evaluated the effects of different clinical and laboratory variables on patient outcome.72  Age, hemoglobin concentration, serum albumin level, and β2-microglobulin level were identified as the predominant outcome predictors in these studies.73  Most of these studies concluded that IgM levels had no prognostic value.74 

An International Prognostic Scoring System was presented at the 2006 American Society of Hematology panel as a staging system for survival for symptomatic patients in need of therapy.75  The parameters used to stratify risk were age older than 65 years, β2-microglobulin level higher than 3 mg/L, monoclonal protein level higher than 70 g/L (7.0 g/dL), hemoglobin concentration less than 115 g/L (11.5 g/dL), and platelet count less than 100 × 109/L. Low risk was defined as the presence of fewer than 1 adverse characteristic except age; high risk, as the presence of more than 2 adverse characteristics; the remaining patients with 2 adverse characteristics or older than 65 years had intermediate risk.

The aim of treatment is to improve the quality and duration of life with minimal adverse effects in the most cost-effective manner. Whether achievement of complete remission confers clinical benefit is still debatable.

Whom to treat

The Third International Workshop on WM76  affirmed the previous consensus panel's recommendations77  that treatment must be reserved for symptomatic patients and should not be initiated on the basis of serum monoclonal protein levels alone. Some of the considerations for initiation of treatment include hemoglobin concentration less than 100 g/L, platelet count less than 100 × 109/L, significant adenopathy or organomegaly, symptomatic hyperviscosity, severe neuropathy, amyloidosis, cryoglobulinemia, cold-agglutinin disease, or evidence of disease transformation. Nevertheless, a serum monoclonal protein level higher than 50 g/L places patients at considerable risk of hyperviscosity. This finding requires a thorough history and physical and funduscopic examinations to diagnose early symptoms and signs of hyperviscosity, and treatment should not be postponed. Therapeutic outcomes should be evaluated using updated consensus panel criteria.76 

Treatment options

The main choices for primary treatment of WM are alkylating agents (chlorambucil, cyclophosphamide, melphalan), purine analogues (cladribine, fludarabine), and monoclonal antibody (rituximab [anti-CD20]). Plasma exchange (1-1.5 volume) is indicated for the acute management of patients with symptoms of hyperviscosity because 80% of the IgM protein is intravascular. Patients may be candidates for initial combination therapy with purine nucleoside analogues or antibody therapy.76 Table 3 summarizes the largest therapeutic trials, including their overall response rates and median response duration.

Chlorambucil (0.1 mg/kg) was the first agent used, with response rates varying between 31% and 92%.78,94  The most common complication of therapy with alkylating agents is development of myelodysplasia and acute nonlymphocytic leukemia from therapy-induced chromosomal breakage.95  Cladribine (0.1 mg/kg) has shown response rates in the range of 44% to 90%.79,96–98  Response rates to fludarabine (30 mg/m2) as initial therapy range from 38% to 100%.80,99–102  Fludarabine and cladribine are cross-resistant.96  The principal dose-limiting toxicity of both these agents is bone marrow suppression and immunosuppression, predisposing patients to infections. Response rates to rituximab (375 mg/m2) vary between 20% and 50%.81,103–107  Rituximab may be regarded as a reasonable choice for treating patients with IgM autoantibody-related neuropathies.108  Patients with polyneuropathy associated with anti-MAG antibodies treated with high-dose rituximab have shown clinical improvement as well as improvement of nerve conduction velocities and decreased anti-MAG antibody titers.109  Polymorphisms in the FcγRIIIA (CD16) receptor gene may affect response to rituximab.110  Transient increases in IgM titers have been reported in 54% of patients after initiation of rituximab therapy. These levels may persist for up to 4 months and do not indicate treatment failure, but they may necessitate plasmapheresis to reduce hyperviscosity, which may result in the loss of the therapeutic antibody.111,112  Patients who had initial IgM flares had poorer response rates than those who did not (28% vs 80%).111 

Because several cell cycle regulators and modulators of apoptosis are degraded through the ubiquitin-proteasome pathway, proteasome inhibitors such as bortezomib have become the focus of clinical research in WM. Chen et al113  described 27 patients in whom bortezomib was found to be active in WM; 21 patients had more than a 25% decrease in IgM.

Experience with the use of thalidomide in WM, as a single agent or in combination with dexamethasone or clarithromycin, is limited. Thalidomide is nonmyelosuppressive, immunomodulatory, and antiangiogenic and may be a reasonable choice for (1) patients for whom first-line therapies have failed, (2) those who have had disease relapse and are not candidates for alkylating or nucleoside analog therapy, or (3) patients with pancytopenia.114  Lenalidomide has been studied in MM and myelodysplastic syndrome and found to be more potent and also to lack the neurotoxic and prothrombotic adverse effects of thalidomide.115  High-dose chemoradiotherapy with autologous stem cell transplantation is occasionally used, but the number of patients treated to date is inadequate to assess its overall role in management. Pentostatin is useful for patients who may benefit from high-dose chemotherapy with autologous stem cell transplantation.93,116  Nonmyeloablative allogeneic peripheral blood stem cell transplantation may be a promising alternative in patients with refractory disease.117  Splenectomy is rarely indicated, but limited case reports suggest that it may be helpful for managing symptomatic painful splenomegaly and hypersplenism.108 

The Third International Workshop on WM76  consensus panel concluded that it was not possible to delineate a particular first-line therapeutic agent and that the choice must be made based on individual patient considerations. Alkylating agents deplete stem cells and hence should not be used among patients who may be eligible for autologous transplantation. The results of ongoing studies will help determine the role of allogeneic or nonmyeloablative allogeneic transplantation in the treatment of WM.

Other therapeutic options currently being evaluated for WM include oblimersen sodium (BCL-2 antisense oligonucleotide),118 131I-tositumomab,119  imatinib mesylate (targets signaling through the stem cell factor and platelet-derived growth factor receptors on WM and mast cells), and dolastatin (microtubule inhibitor).

Dimopoulos et al120  showed that a combination of dexamethasone, rituximab, and cyclophosphamide is an active and well-tolerated treatment for symptomatic patients requiring therapy. Disease control was achieved in the majority of patients without the risks of myelosuppression and immunosuppression, which may occur when nucleoside analogues are used. The interim results of a study by Treon et al121  suggest that a combination regimen of bortezomib, dexamethasone, and rituximab is highly active and well tolerated in the primary treatment of WM.

One pathway that conveys survival signals in mammalian cells is based on phosphoinositide 3-kinase/Akt. Akt is a principal signaling protein in the cellular pathways that leads to muscle hypertrophy and tissue growth. This pathway also has an important role in the migration, adhesion, and homing of WM in vitro and in vivo.122  Akt inhibitors such as perifosine lower the resistance of tumor cells to various therapeutic modalities and induce apoptosis in WM. Perifosine also activates mitogen-activated protein kinase pathways and protein kinase C proteins, thereby promoting cell proliferation. Specific Akt inhibitors such as triciribine induce cytotoxicity without enhancing MEK/ERK (mitogen-activated protein kinase kinase/extracellular signal-regulated kinase) activity. Combining perifosine with MEK inhibitors or protein kinase C inhibitors such as AZD6244 and enzastaurin may provide therapeutic advantages in WM. Perifosine in combination with bortezomib, rituximab, and other agents has been shown to have enhanced cytotoxicity on WM cell lines.123  Additional clinical trials are required to establish the efficacy and safety of such combinations.

Rossi et al124  reported a phase 1/2 study of atacicept (TACI-Ig) in refractory WM. Atacicept acts as a decoy receptor by binding to and neutralizing soluble BLyS and APRIL, and preventing these TNF family members from binding to their cognate receptors (TACI, BCMA, and BAFF-R) on B-cell tumors, thereby enhancing cytotoxicity. Mast cells may have a role in tumor cell expansion through constitutive CD154-CD40 signaling; therefore, CD154 blocking agents may prove to be a therapeutic option in WM.3  Ho et al125  described the role of sCD27 in the pathogenesis of WM and demonstrated the feasibility of targeting CD70 and sCD27-CD70 interactions with the SGN-70 monoclonal antibody.

WM cells in the bone marrow and mast cells express CD52. In a phase 2 study by Hunter et al,126  alemtuzumab (humanized monoclonal antibody against CD52) has been reported to be highly active in WM. Sildenafil citrate has been shown to induce apoptosis in WM cell lines.127  In a prospective study,128  30 patients with slowly progressing WM who did not meet consensus eligibility for active therapy were treated with sildenafil citrate; disease progression was suppressed in more than half the patients. After 3 months of therapy, 63% showed a significant decrease in IgM levels and 17% showed a minor response. These results encourage additional clinical trials.

WM is a B-cell neoplasm characterized by lymphoplasmacytic infiltration of the bone marrow and elevated serum monoclonal IgM levels. The symptoms of WM are attributable to the extent of tumor infiltration and/or elevated IgM levels. The prognostic factors of significance include the patient's age, β2-microglobulin level, monoclonal protein level, hemoglobin concentration, and platelet count. Therapy is reserved exclusively for symptomatic patients; the main therapeutic options include alkylating agents, nucleoside analogues, and rituximab. Preliminary results of ongoing studies involving combination chemotherapy are encouraging. At present, no specific agent or regimen is superior to another in the treatment of WM. Other novel agents are being investigated; their unique mechanisms of action and toxicity profiles hold promise in the development of rational targeted therapy in WM.

Editing, proofreading, and reference verification were provided by the Section of Scientific Publications, Mayo Clinic.

Contribution: A.V. and M.A.G. contributed to concept and design, acquisition of data, and analysis and interpretation of data.

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

Correspondence: Morie A. Gertz, Division of Hematology, Mayo Clinic, 200 First St SW, Rochester, MN 55905; e-mail: gertz.morie@mayo.edu.

1
Waldenström
 
J
Incipient myelomatosis or ‘essential’ hyperglobulinemia with fibrinogenopenia: a new syndrome?
Acta Med Scand
1944
, vol. 
117
 (pg. 
216
-
247
)
2
Treon
 
SP
Morel
 
P
Leblond
 
V
Fermand
 
JP
Report of the Third International Workshop on Waldenström's macroglobulinemia.
Clin Lymphoma
2005
, vol. 
5
 (pg. 
215
-
216
)
3
Tournilhac
 
O
Santos
 
DD
Xu
 
L
, et al. 
Mast cells in Waldenström's macroglobulinemia support lymphoplasmacytic cell growth through CD154/CD40 signaling.
Ann Oncol
2006
, vol. 
17
 (pg. 
1275
-
1282
)
4
Stansfeld
 
AG
Diebold
 
J
Noel
 
H
, et al. 
Updated Kiel classification for lymphomas [erratum in: Lancet. 1988;1:372]
Lancet
1988
, vol. 
1
 (pg. 
292
-
293
)
5
Harris
 
NL
Jaffe
 
ES
Stein
 
H
, et al. 
A revised European-American classification of lymphoid neoplasms: a proposal from the International Lymphoma Study Group.
Blood
1994
, vol. 
84
 (pg. 
1361
-
1392
)
6
Jaffe
 
ES
Harris
 
NL
Stein
 
H
Vardiman
 
JW
World Health Organization Classification of Tumours: Pathology and Genetics of Tumours of Haematopoietic and Lymphoid Tissues.
2001
Lyon, France
IARC Press
7
Owen
 
RG
Treon
 
SP
Al-Katib
 
A
, et al. 
Clinicopathological definition of Waldenström's macroglobulinemia: consensus panel recommendations from the Second International Workshop on Waldenström's Macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
110
-
115
)
8
Herrinton
 
LJ
Weiss
 
NS
Incidence of Waldenström's macroglobulinemia.
Blood
1993
, vol. 
82
 (pg. 
3148
-
3150
)
9
Groves
 
FD
Travis
 
LB
Devesa
 
SS
Ries
 
LA
Fraumeni
 
JF
Waldenström's macroglobulinemia: incidence patterns in the United States, 1988-1994.
Cancer
1998
, vol. 
82
 (pg. 
1078
-
1081
)
10
Benjamin
 
M
Reddy
 
S
Brawley
 
OW
Myeloma and race: a review of the literature.
Cancer Metastasis Rev
2003
, vol. 
22
 (pg. 
87
-
93
)
11
Dimopoulos
 
MA
Panayiotidis
 
P
Moulopoulos
 
LA
Sfikakis
 
P
Dalakas
 
M
Waldenström's macroglobulinemia: clinical features, complications, and management.
J Clin Oncol
2000
, vol. 
18
 (pg. 
214
-
226
)
12
Dimopoulos
 
MA
Galani
 
E
Matsouka
 
C
Waldenström's macroglobulinemia.
Hematol Oncol Clin North Am
1999
, vol. 
13
 (pg. 
1351
-
1366
)
13
McMaster
 
ML
Familial Waldenström's macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
146
-
152
)
14
Treon
 
SP
Hunter
 
ZR
Aggarwal
 
A
, et al. 
Characterization of familial Waldenström's macroglobulinemia.
Ann Oncol
2006
, vol. 
17
 (pg. 
488
-
494
)
15
Kyle
 
RA
Therneau
 
TM
Rajkumar
 
SV
, et al. 
A long-term study of prognosis in monoclonal gammopathy of undetermined significance.
N Engl J Med
2002
, vol. 
346
 (pg. 
564
-
569
)
16
Morra
 
E
Cesana
 
C
Klersy
 
C
, et al. 
Predictive variables for malignant transformation in 452 patients with asymptomatic IgM monoclonal gammopathy.
Semin Oncol
2003
, vol. 
30
 (pg. 
172
-
177
)
17
Silvestri
 
F
Barillari
 
G
Fanin
 
R
, et al. 
Risk of hepatitis C virus infection, Waldenström's macroglobulinemia, and monoclonal gammopathies.
Blood
1996
, vol. 
88
 (pg. 
1125
-
1126
)
18
Leleu
 
X
O'Connor
 
K
Ho
 
AW
, et al. 
Hepatitis C viral infection is not associated with Waldenström's macroglobulinemia.
Am J Hematol
2007
, vol. 
82
 (pg. 
83
-
84
)
19
Weller
 
S
Braun
 
MC
Tan
 
BK
, et al. 
Human blood IgM “memory” B cells are circulating splenic marginal zone B cells harboring a prediversified immunoglobulin repertoire.
Blood
2004
, vol. 
104
 (pg. 
3647
-
3654
)
20
Schop
 
RF
Kuehl
 
WM
Van Wier
 
SA
, et al. 
Waldenström macroglobulinemia neoplastic cells lack immunoglobulin heavy chain locus translocations but have frequent 6q deletions.
Blood
2002
, vol. 
100
 (pg. 
2996
-
3001
)
21
Kriangkum
 
J
Taylor
 
BJ
Treon
 
SP
Mant
 
MJ
Belch
 
AR
Pilarski
 
LM
Clonotypic IgM V/D/J sequence analysis in Waldenström macroglobulinemia suggests an unusual B-cell origin and an expansion of polyclonal B cells in peripheral blood.
Blood
2004
, vol. 
104
 (pg. 
2134
-
2142
)
22
Sahota
 
SS
Forconi
 
F
Ottensmeier
 
CH
Stevenson
 
FK
Origins of the malignant clone in typical Waldenström's macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
136
-
141
)
23
Sahota
 
SS
Forconi
 
F
Ottensmeier
 
CH
, et al. 
Typical Waldenström macroglobulinemia is derived from a B-cell arrested after cessation of somatic mutation but prior to isotype switch events.
Blood
2002
, vol. 
100
 (pg. 
1505
-
1507
)
24
Martin-Jimenez
 
P
Garcia-Sanz
 
R
Sarasquete
 
ME
, et al. 
Functional class switch recombination may occur ‘in vivo’ in Waldenström macroglobulinaemia.
Br J Haematol
2007
, vol. 
136
 (pg. 
114
-
116
)
25
Schop
 
RF
Van Wier
 
SA
Xu
 
R
, et al. 
6q deletion discriminates Waldenstrom macroglobulinemia from IgM monoclonal gammopathy of undetermined significance.
Cancer Genet Cytogenet
2006
, vol. 
169
 (pg. 
150
-
153
)
26
Terre
 
C
Nguyen-Khac
 
F
Barin
 
C
, et al. 
Trisomy 4, a new chromosomal abnormality in Waldenström's macroglobulinemia: a study of 39 cases.
Leukemia
2006
, vol. 
20
 (pg. 
1634
-
1636
)
27
Mansoor
 
A
Medeiros
 
LJ
Weber
 
DM
, et al. 
Cytogenetic findings in lymphoplasmacytic lymphoma/Waldenström macroglobulinemia: chromosomal abnormalities are associated with the polymorphous subtype and an aggressive clinical course.
Am J Clin Pathol
2001
, vol. 
116
 (pg. 
543
-
549
)
28
Liu
 
YC
Miyazawa
 
K
Sashida
 
G
Kodama
 
A
Ohyashiki
 
K
Deletion (20q) as the sole abnormality in Waldenström macroglobulinemia suggests distinct pathogenesis of 20q11 anomaly.
Cancer Genet Cytogenet
2006
, vol. 
169
 (pg. 
69
-
72
)
29
Turner
 
CA
Mack
 
DH
Davis
 
MM
Blimp-1, a novel zinc finger-containing protein that can drive the maturation of B lymphocytes into immunoglobulin-secreting cells.
Cell
1994
, vol. 
77
 (pg. 
297
-
306
)
30
Schebesta
 
M
Heavey
 
B
Busslinger
 
M
Transcriptional control of B-cell development.
Curr Opin Immunol
2002
, vol. 
14
 (pg. 
216
-
223
)
31
Calame
 
KL
Plasma cells: finding new light at the end of B cell development.
Nat Immunol
2001
, vol. 
2
 (pg. 
1103
-
1108
)
32
Schneider
 
P
MacKay
 
F
Steiner
 
V
, et al. 
BAFF, a novel ligand of the tumor necrosis factor family, stimulates B cell growth.
J Exp Med
1999
, vol. 
189
 (pg. 
1747
-
1756
)
33
Locksley
 
RM
Killeen
 
N
Lenardo
 
MJ
The TNF and TNF receptor superfamilies: integrating mammalian biology.
Cell
2001
, vol. 
104
 (pg. 
487
-
501
)
34
MacKay
 
F
Schneider
 
P
Rennert
 
P
Browning
 
J
BAFF AND APRIL: a tutorial on B cell survival.
Annu Rev Immunol
2003
, vol. 
21
 (pg. 
231
-
264
)
35
Elsawa
 
SF
Novak
 
AJ
Grote
 
DM
, et al. 
B-lymphocyte stimulator (BlyS) stimulates immunoglobulin production and malignant B-cell growth in Waldenström macroglobulinemia.
Blood
2006
, vol. 
107
 (pg. 
2882
-
2888
)
36
Wu
 
Y
Bressette
 
D
Carrell
 
JA
, et al. 
Tumor necrosis factor (TNF) receptor superfamily member TACI is a high affinity receptor for TNF family members APRIL and BLyS.
J Biol Chem
2000
, vol. 
275
 (pg. 
35478
-
35485
)
37
Hunter
 
Z
Leleu
 
X
Hatjiharissi
 
E
, et al. 
IgA and IgG hypogammaglobulinemia are associated with mutations in the APRIL/BLYS receptor TACI in Waldenström's macroglobulinemia (WM) [abstract].
Blood
2006
, vol. 
108
  
Abstract 228
38
Adamia
 
S
Crainie
 
M
Kriangkum
 
J
Mant
 
MJ
Belch
 
AR
Pilarski
 
LM
Abnormal expression of hyaluronan synthases in patients with Waldenström's macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
165
-
168
)
39
Adamia
 
S
Treon
 
SP
Reiman
 
T
, et al. 
Potential impact of a single nucleotide polymorphism in the hyaluronan synthase 1 gene in Waldenström's macroglobulinemia.
Clin Lymphoma
2005
, vol. 
5
 (pg. 
253
-
256
)
40
Hatzimichael
 
EC
Christou
 
L
Bai
 
M
Kolios
 
G
Kefala
 
L
Bourantas
 
KL
Serum levels of IL-6 and its soluble receptor (sIL-6R) in Waldenström's macroglobulinemia.
Eur J Haematol
2001
, vol. 
66
 (pg. 
1
-
6
)
41
Owen
 
RG
Barrans
 
SL
Richards
 
SJ
, et al. 
Waldenström macroglobulinemia: development of diagnostic criteria and identification of prognostic factors.
Am J Clin Pathol
2001
, vol. 
116
 (pg. 
420
-
428
)
42
San Miguel
 
JF
Vidriales
 
MB
Ocio
 
E
, et al. 
Immunophenotypic analysis of Waldenström's macroglobulinemia [erratum in: Semin Oncol. 2002;15:427].
Semin Oncol
2003
, vol. 
30
 (pg. 
187
-
195
)
43
Konoplev
 
S
Medeiros
 
LJ
Bueso-Ramos
 
CE
Jorgensen
 
JL
Lin
 
P
Immunophenotypic profile of lymphoplasmacytic lymphoma/Waldenström macroglobulinemia.
Am J Clin Pathol
2005
, vol. 
124
 (pg. 
414
-
420
)
44
Hunter
 
ZR
Branagan
 
AR
Manning
 
R
, et al. 
CD5, CD10, and CD23 expression in Waldenström's macroglobulinemia.
Clin Lymphoma
2005
, vol. 
5
 (pg. 
246
-
249
)
45
Walsh
 
SH
Laurell
 
A
Sundstrom
 
G
Roos
 
G
Sundstrom
 
C
Rosenquist
 
R
Lymphoplasmacytic lymphoma/Waldenström's macroglobulinemia derives from an extensively hypermutated B cell that lacks ongoing somatic hypermutation.
Leuk Res
2005
, vol. 
29
 (pg. 
729
-
734
)
46
Gertz
 
MA
Fonseca
 
R
Rajkumar
 
SV
Waldenström's macroglobulinemia.
Oncologist
2000
, vol. 
5
 (pg. 
63
-
67
)
47
Fadil
 
A
Taylor
 
DE
The lung and Waldenström's macroglobulinemia.
South Med J
1998
, vol. 
91
 (pg. 
681
-
685
)
48
Rosenthal
 
JA
Curran
 
WJ
Schuster
 
SJ
Waldenström's macroglobulinemia resulting from localized gastric lymphoplasmacytoid lymphoma.
Am J Hematol
1998
, vol. 
58
 (pg. 
244
-
245
)
49
Civit
 
T
Coulbois
 
S
Baylac
 
F
Taillandier
 
L
Auque
 
J
Waldenström's macroglobulinemia and cerebral lymphoplasmocytic proliferation: Bing and Neel syndrome: apropos of a new case [French].
Neurochirurgie
1997
, vol. 
43
 (pg. 
245
-
249
)
50
Orellana
 
J
Friedman
 
AH
Ocular manifestations of multiple myeloma, Waldenström's macroglobulinemia and benign monoclonal gammopathy.
Surv Ophthalmol
1981
, vol. 
26
 (pg. 
157
-
169
)
51
Gertz
 
MA
Kyle
 
RA
Hyperviscosity syndrome.
J Intensive Care Med
1995
, vol. 
10
 (pg. 
128
-
141
)
52
Kwaan
 
HC
Bongu
 
A
The hyperviscosity syndromes.
Semin Thomb Hemost
1999
, vol. 
25
 (pg. 
199
-
208
)
53
Brouet
 
JC
Clauvel
 
JP
Danon
 
F
Klein
 
M
Seligmann
 
M
Biologic and clinical significance of cryoglobulins: a report of 86 cases.
Am J Med
1974
, vol. 
57
 (pg. 
775
-
788
)
54
Gertz
 
MA
Kyle
 
RA
Noel
 
P
Primary systemic amyloidosis: a rare complication of immunoglobulin M monoclonal gammopathies and Waldenström's macroglobulinemia.
J Clin Oncol
1993
, vol. 
11
 (pg. 
914
-
920
)
55
Gardyn
 
J
Schwartz
 
A
Gal
 
R
Lewinski
 
U
Kristt
 
D
Cohen
 
AM
Waldenström's macroglobulinemia associated with AA amyloidosis.
Int J Hematol
2001
, vol. 
74
 (pg. 
76
-
78
)
56
Veltman
 
GA
van Veen
 
S
Kluin-Nelemans
 
JC
Bruijn
 
JA
van Es
 
LA
Renal disease in Waldenström's macroglobulinaemia.
Nephrol Dial Transplant
1997
, vol. 
12
 (pg. 
1256
-
1259
)
57
Pruzanski
 
W
Shumak
 
KH
Biologic activity of cold-reacting autoantibodies (first of two parts).
N Engl J Med
1977
, vol. 
297
 (pg. 
538
-
542
)
58
Schnitzler
 
L
Schubert
 
B
Boasson
 
M
Gardais
 
J
Tourmen
 
A
Urticaire chronique, lésions osseuses, macroglobulinémie IgM: maladie de Waldenström?
Bull Soc Fr Derm Syph
1974
, vol. 
81
 (pg. 
363
-
367
)
59
Asli
 
B
Bienvenu
 
B
Cordoliani
 
F
, et al. 
Chronic urticaria and monoclonal IgM gammopathy (Schnitzler's syndrome): report of 11 cases treated by pefloxacine.
Arch Derm
 
In press
60
Nobile-Orazio
 
E
IgM paraproteinaemic neuropathies.
Curr Opin Neurol
2004
, vol. 
17
 (pg. 
599
-
605
)
61
Baldini
 
L
Nobile-Orazio
 
E
Guffanti
 
A
, et al. 
Peripheral neuropathy in IgM monoclonal gammopathy and Waldenström's macroglobulinemia: a frequent complication in elderly males with low MAG-reactive serum monoclonal component.
Am J Hematol
1994
, vol. 
45
 (pg. 
25
-
31
)
62
Levine
 
T
Pestronk
 
A
Florence
 
J
, et al. 
Peripheral neuropathies in Waldenström's macroglobulinaemia.
J Neurol Neurosurg Psychiatry
2006
, vol. 
77
 (pg. 
224
-
228
)
63
Ropper
 
AH
Gorson
 
KC
Neuropathies associated with paraproteinemia.
N Engl J Med
1998
, vol. 
338
 (pg. 
1601
-
1607
)
64
Alexanian
 
R
Weber
 
D
Delasalle
 
K
Cabanillas
 
F
Dimopoulos
 
M
Asymptomatic Waldenström's macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
206
-
210
)
65
Kyle
 
RA
Therneau
 
TM
Rajkumar
 
SV
, et al. 
Long-term follow-up of IgM monoclonal gammopathy of undetermined significance.
Blood
2003
, vol. 
102
 (pg. 
3759
-
3764
)
66
Ocio
 
EM
Hernandez
 
JM
Mateo
 
G
, et al. 
Immunophenotypic and cytogenetic comparison of Waldenström's macroglobulinemia with splenic marginal zone lymphoma.
Clin Lymphoma
2005
, vol. 
5
 (pg. 
241
-
245
)
67
Avet-Loiseau
 
H
Garand
 
R
Lode
 
L
Robillard
 
N
Bataille
 
R
14q32 translocations discriminate IgM multiple myeloma from Waldenström's macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
153
-
155
)
68
Binet
 
JL
Caligaris-Cappio
 
F
Catovsky
 
D
, et al. 
Perspectives on the use of new diagnostic tools in the treatment of chronic lymphocytic leukemia.
Blood
2006
, vol. 
107
 (pg. 
859
-
861
)
69
Noel
 
P
Kyle
 
RA
Monoclonal proteins in chronic lymphocytic leukemia.
Am J Clin Pathol
1987
, vol. 
87
 (pg. 
385
-
388
)
70
Lin
 
P
Mansoor
 
A
Bueso-Ramos
 
C
Hao
 
S
Lai
 
R
Medeiros
 
LJ
Diffuse large B-cell lymphoma occurring in patients with lymphoplasmacytic lymphoma/Waldenström macroglobulinemia: clinicopathologic features of 12 cases.
Am J Clin Pathol
2003
, vol. 
120
 (pg. 
246
-
253
)
71
Owen
 
RG
Developing diagnostic criteria in Waldenström's macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
196
-
200
)
72
Merlini
 
G
Baldini
 
L
Broglia
 
C
, et al. 
Prognostic factors in symptomatic Waldenström's macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
211
-
215
)
73
Ghobrial
 
IM
Fonseca
 
R
Gertz
 
MA
, et al. 
Prognostic model for disease-specific and overall mortality in newly diagnosed symptomatic patients with Waldenström macroglobulinaemia.
Br J Haematol
2006
, vol. 
133
 (pg. 
158
-
164
)
74
Dimopoulos
 
M
Gika
 
D
Zervas
 
K
, et al. 
The international staging system for multiple myeloma is applicable in symptomatic Waldenström's macroglobulinemia.
Leuk Lymphoma
2004
, vol. 
45
 (pg. 
1809
-
1813
)
75
Morel
 
P
Duhamel
 
A
Gobbi
 
P
, et al. 
International prognostic scoring system (IPSS) for Waldenström's macroglobulinemia (WM) [abstract].
Blood
2006
, vol. 
108
  
Abstract 127
76
Treon
 
SP
Gertz
 
MA
Dimopoulos
 
M
, et al. 
Update on treatment recommendations from the Third International Workshop on Waldenström's macroglobulinemia.
Blood
2006
, vol. 
107
 (pg. 
3442
-
3446
)
77
Kyle
 
RA
Treon
 
SP
Alexanian
 
R
, et al. 
Prognostic markers and criteria to initiate therapy in Waldenström's macroglobulinemia: consensus panel recommendations from the Second International Workshop on Waldenström's Macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
116
-
120
)
78
Facon
 
T
Brouillard
 
M
Duhamel
 
A
, et al. 
Prognostic factors in Waldenström's macroglobulinemia: a report of 167 cases.
J Clin Oncol
1993
, vol. 
11
 (pg. 
1553
-
1558
)
79
Dimopoulos
 
MA
Weber
 
D
Delasalle
 
KB
Keating
 
M
Alexanian
 
R
Treatment of Waldenström's macroglobulinemia resistant to standard therapy with 2-chlorodeoxyadenosine: identification of prognostic factors.
Ann Oncol
1995
, vol. 
6
 (pg. 
49
-
52
)
80
Dhodapkar
 
MV
Jacobson
 
JL
Gertz
 
MA
, et al. 
Prognostic factors and response to fludarabine therapy in patients with Waldenström macroglobulinemia: results of United States intergroup trial (Southwest Oncology Group S9003).
Blood
2001
, vol. 
98
 (pg. 
41
-
48
)
81
Gertz
 
MA
Rue
 
M
Blood
 
E
Kaminer
 
LS
Vesole
 
DH
Greipp
 
PR
Multicenter phase 2 trial of rituximab for Waldenström macroglobulinemia (WM): an Eastern Cooperative Oncology Group Study (E3A98).
Leuk Lymphoma
2004
, vol. 
45
 (pg. 
2047
-
2055
)
82
Petrucci
 
MT
Avvisati
 
G
Tribalto
 
M
Giovangrossi
 
P
Mandelli
 
F
Waldenström's macroglobulinaemia: results of a combined oral treatment in 34 newly diagnosed patients.
J Intern Med
1989
, vol. 
226
 (pg. 
443
-
447
)
83
Case
 
DC
Ervin
 
TJ
Boyd
 
MA
Redfield
 
DL
Waldenström's macroglobulinemia: long-term results with the M-2 protocol.
Cancer Invest
1991
, vol. 
9
 (pg. 
1
-
7
)
84
Dimopoulos
 
MA
Alexanian
 
R
Waldenström's macroglobulinemia.
Blood
1994
, vol. 
83
 (pg. 
1452
-
1459
)
85
Leblond
 
V
Levy
 
V
Maloisel
 
F
, et al. 
Multicenter, randomized comparative trial of fludarabine and the combination of cyclophosphamide-doxorubicin-prednisone in 92 patients with Waldenström macroglobulinemia in first relapse or with primary refractory disease.
Blood
2001
, vol. 
98
 (pg. 
2640
-
2644
)
86
Hunter
 
Z
Branagan
 
A
Treon
 
SP
CHOP plus rituximab (CHOP-R) in Waldenström's macroglobulinemia (WM) [abstract].
Blood
2004
, vol. 
104
 pg. 
314b
  
Abstract 4930
87
Dimopoulos
 
MA
Kyrtsonis
 
MC
Tsatalas
 
C
, et al. 
Primary treatment of Waldenström's macroglobulinemia (WM) with dexamethasone, rituximab and cyclophosphamide [abstract].
Blood
2004
, vol. 
104
 pg. 
215a
  
Abstract 752
88
Annibali
 
O
Petrucci
 
MT
Martini
 
V
, et al. 
Treatment of 72 newly diagnosed Waldenström macroglobulinemia cases with oral melphalan, cyclophosphamide, and prednisone: results and cost analysis.
Cancer
2005
, vol. 
103
 (pg. 
582
-
587
)
89
Weber
 
DM
Dimopoulos
 
MA
Delasalle
 
K
Rankin
 
K
Gavino
 
M
Alexanian
 
R
2-Chlorodeoxyadenosine alone and in combination for previously untreated Waldenström's macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
243
-
247
)
90
Tam
 
CS
Wolf
 
MM
Westerman
 
D
Januszewicz
 
EH
Prince
 
HM
Seymour
 
JF
Fludarabine combination therapy is highly effective in first-line and salvage treatment of patients with Waldenström's macroglobulinemia.
Clin Lymphoma Myeloma
2005
, vol. 
6
 (pg. 
136
-
139
)
91
Tamburini
 
J
Lévy
 
V
Chaletiex
 
C
, et al. 
Fludarabine plus cyclophosphamide in Waldenström's macroglobulinemia: results in 49 patients.
Leukemia
2005
, vol. 
19
 (pg. 
1831
-
1834
)
92
Branagan
 
A
Hunter
 
Z
Santos
 
DD
Treon
 
SP
Thalidomide and rituximab in Waldenström's macroglobulinemia [abstract].
Blood
2004
, vol. 
104
  
Abstract 1484
93
Hensel
 
M
Villalobos
 
M
Kornacker
 
M
Krasniqi
 
F
Ho
 
AD
Pentostatin/cyclophosphamide with or without rituximab: an effective regimen for patients with Waldenström's macroglobulinemia/lymphoplasmacytic lymphoma.
Clin Lymphoma Myeloma
2005
, vol. 
6
 (pg. 
131
-
135
)
94
Kyle
 
RA
Greipp
 
PR
Gertz
 
MA
, et al. 
Waldenström's macroglobulinaemia: a prospective study comparing daily with intermittent oral chlorambucil.
Br J Haematol
2000
, vol. 
108
 (pg. 
737
-
742
)
95
Rodriguez
 
JN
Fernandez-Jurado
 
A
Martino
 
ML
Prados
 
D
Waldenström's macroglobulinemia complicated with acute myeloid leukemia: report of a case and review of the literature.
Haematologica
1998
, vol. 
83
 (pg. 
91
-
92
)
96
Dimopoulos
 
MA
Weber
 
DM
Kantarjian
 
H
Keating
 
M
Alexanian
 
R
2Chlorodeoxyadenosine therapy of patients with Waldenström macroglobulinemia previously treated with fludarabine.
Ann Oncol
1994
, vol. 
5
 (pg. 
288
-
289
)
97
Delannoy
 
A
Van Den Neste
 
E
Michaux
 
JL
Bosly
 
A
Ferrant
 
A
Cladribine for Waldenström's macroglobulinaemia.
Br J Haematol
1999
, vol. 
104
 (pg. 
933
-
934
)
98
Fridrik
 
MA
Jager
 
G
Baldinger
 
C
Krieger
 
O
Chott
 
A
Bettelheim
 
P
First-line treatment of Waldenström's disease with cladribine. Arbeitsgemeinschaft Medikamentose Tumortherapie.
Ann Hematol
1997
, vol. 
74
 (pg. 
7
-
10
)
99
Dimopoulos
 
MA
O'Brien
 
S
Kantarjian
 
H
, et al. 
Fludarabine therapy in Waldenström's macroglobulinemia.
Am J Med
1993
, vol. 
95
 (pg. 
49
-
52
)
100
Dhodapkar
 
MV
Jacobson
 
JL
Gertz
 
MA
Crowley
 
JJ
Barlogie
 
B
Prognostic factors and response to fludarabine therapy in Waldenström's macroglobulinemia: an update of a US intergroup trial (SWOG S9003).
Semin Oncol
2003
, vol. 
30
 (pg. 
220
-
225
)
101
Foran
 
JM
Rohatiner
 
AZ
Coiffier
 
B
, et al. 
Multicenter phase II study of fludarabine phosphate for patients with newly diagnosed lymphoplasmacytoid lymphoma, Waldenström's macroglobulinemia, and mantle-cell lymphoma.
J Clin Oncol
1999
, vol. 
17
 (pg. 
546
-
553
)
102
Thalhammer-Scherrer
 
R
Geissler
 
K
Schwarzinger
 
I
, et al. 
Fludarabine therapy in Waldenström's macroglobulinemia.
Ann Hematol
2000
, vol. 
79
 (pg. 
556
-
559
)
103
Foran
 
JM
Rohatiner
 
AZ
Cunningham
 
D
, et al. 
European phase II study of rituximab (chimeric anti-CD20 monoclonal antibody) for patients with newly diagnosed mantle-cell lymphoma and previously treated mantle-cell lymphoma, immunocytoma and small B-cell lymphocytic lymphoma [erratum in: J Clin Oncol. 2000;18:2006].
J Clin Oncol
2000
, vol. 
18
 (pg. 
317
-
324
)
104
Treon
 
SP
Agus
 
DB
Link
 
B
, et al. 
CD20-directed antibody-mediated immunotherapy induces responses and facilitates hematologic recovery in patients with Waldenström's macroglobulinemia.
J Immunother
2001
, vol. 
24
 (pg. 
272
-
279
)
105
Treon
 
SP
Emmanouilides
 
C
Kimby
 
E
, et al. 
Extended rituximab therapy in Waldenström's macroglobulinemia.
Ann Oncol
2005
, vol. 
16
 (pg. 
132
-
138
)
106
Byrd
 
JC
White
 
CA
Link
 
B
, et al. 
Rituximab therapy in Waldenström's macroglobulinemia: preliminary evidence of clinical activity.
Ann Oncol
1999
, vol. 
10
 (pg. 
1525
-
1527
)
107
Dimopoulos
 
MA
Zervas
 
C
Zomas
 
A
, et al. 
Extended rituximab therapy for previously untreated patients with Waldenström's macroglobulinemia.
Clin Lymphoma
2002
, vol. 
3
 (pg. 
163
-
166
)
108
Gertz
 
MA
Anagnostopoulos
 
A
Anderson
 
K
, et al. 
Treatment recommendations in Waldenström's macroglobulinemia: consensus panel recommendations from the Second International Workshop on Waldenström's Macroglobulinemia.
Semin Oncol
2003
, vol. 
30
 (pg. 
121
-
126
)
109
Renaud
 
S
Fuhr
 
P
Gregor
 
M
, et al. 
High-dose rituximab and anti-MAG-associated polyneuropathy.
Neurology
2006
, vol. 
66
 (pg. 
742
-
744
)
110
Treon
 
SP
Hansen
 
M
Branagan
 
AR
, et al. 
Polymorphisms in FcγRIIIA (CD16) receptor expression are associated with clinical response to rituximab in Waldenström's macroglobulinemia.
J Clin Oncol
2005
, vol. 
23
 (pg. 
474
-
481
)
111
Ghobrial
 
IM
Fonseca
 
R
Greipp
 
PR
, et al. 
Initial immunoglobulin M ‘flare’ after rituximab therapy in patients diagnosed with Waldenström macroglobulinemia: an Eastern Cooperative Oncology Group Study.
Cancer
2004
, vol. 
101
 (pg. 
2593
-
2598
)
112
Treon
 
SP
Branagan
 
AR
Hunter
 
Z
Santos
 
D
Tournhilac
 
O
Anderson
 
KC
Paradoxical increases in serum IgM and viscosity levels following rituximab in Waldenström's macroglobulinemia.
Ann Oncol
2004
, vol. 
15
 (pg. 
1481
-
1483
)
113
Chen
 
CI
Kouroukis
 
T
White
 
D
, et al. 
Bortezomib is active in Waldenström's macroglobulinemia (WM): results of a National Cancer Institute of Canada (NCIC) phase II study in previously untreated or treated WM [abstract].
J Clin Oncol
2006
, vol. 
24
 
suppl 432S
114
Dimopoulos
 
MA
Tsatalas
 
C
Zomas
 
A
, et al. 
Treatment of Waldenström's macroglobulinemia with single-agent thalidomide or with the combination of clarithromycin, thalidomide and dexamethasone.
Semin Oncol
2003
, vol. 
30
 (pg. 
265
-
269
)
115
Crane
 
E
List
 
A
Lenalidomide: an immunomodulatory drug.
Future Oncol
2005
, vol. 
1
 (pg. 
575
-
583
)
116
Ho
 
AD
Hensel
 
M
Pentostatin for the treatment of indolent lymphoproliferative disorders.
Semin Hematol
2006
, vol. 
43
 
suppl 2
(pg. 
S2
-
S10
)
117
Ueda
 
T
Hatanaka
 
K
Kosugi
 
S
Kishino
 
BI
Tamaki
 
T
Successful non-myeloablative allogeneic peripheral blood stem cell transplantation (PBSCT) for Waldenström's macroglobulinemia with severe pancytopenia.
Bone Marrow Transplant
2001
, vol. 
28
 (pg. 
609
-
611
)
118
Frankel
 
SR
Oblimersen sodium (G3139 Bcl-2 antisense oligonucleotide) therapy in Waldenström's macroglobulinemia: a targeted approach to enhance apoptosis.
Semin Oncol
2003
, vol. 
30
 (pg. 
300
-
304
)
119
Tsai
 
DE
Maillard
 
I
Downs
 
LH
, et al. 
Use of iodine 131I-tositumomab radioimmunotherapy in a patient with Waldenström's macroglobulinemia.
Leuk Lymphoma
2004
, vol. 
45
 (pg. 
591
-
595
)
120
Dimopoulos
 
MA
Anagnostopoulos
 
A
Kyrtsonis
 
MC
, et al. 
Primary treatment of Waldenstöm's macroglobulinemia (WM) with dexamethasone, rituximab and cycloposphamide [abstract].
Blood
2006
, vol. 
108
  
Abstract 128
121
Treon
 
SP
Soumerai
 
JD
Patterson
 
CJ
, et al. 
Bortezomib dexamethasone and rituximab (BDR) is a highly active regimen in the primary therapy of Waldenström's macroglobulinemia: planned interim results of WMCTG clinical trial 05-180 [abstract].
Blood
2006
, vol. 
108
  
Abstract 2765
122
Leleu
 
X
Ngo
 
H
Runnels
 
J
, et al. 
The PI3K/Akt pathway is an important regulator of homing and adhesion in Waldenström's macroglobulinemia [abstract].
Blood
2006
, vol. 
108
  
Abstract 2417
123
Leleu
 
X
O'Sullivan
 
G
Jia
 
X
, et al. 
Novel agent perifosine enhances antitumor activity of bortezomib, rituximab and other conventional therapies in Waldenström's macroglobulinemia [abstract].
Blood
2006
, vol. 
108
  
Abstract 2517
124
Rossi
 
JF
Moreaux
 
J
Rose
 
M
, et al. 
A Phase I/II study of atacicept (TACI-Ig) to neutralize APRIL and BLyS in patients with refractory or relapsed multiple myeloma (MM) or active previously treated Waldenström's macroglobulinemia (WM) [abstract].
Blood
2006
, vol. 
108
  
Abstract 3578
125
Ho
 
AW
Hatjiharissi
 
E
Branagan
 
A
, et al. 
Therapeutic targeting of CD70 and CD27–70 interactions with the monoclonal antibody SGN-70 in Waldenström's macroglobulinemia (WM) [abstract].
J Clin Oncol
2006
, vol. 
24
 
suppl 102S
126
Hunter
 
ZR
Boxer
 
M
Kahl
 
B
Patterson
 
CJ
Soumerai
 
JD
Treon
 
SP
Phase II study of alemtuzumab in lymphoplasmacytic lymphoma: results of WMCTG trial 02-079 [abstract].
J Clin Oncol
2006
, vol. 
24
 pg. 
427s
 
127
Treon
 
SP
Tournilhac
 
O
Branagan
 
AR
, et al. 
Clinical responses to sildenafil in Waldenström's macroglobulinemia.
Clin Lymphoma
2004
, vol. 
5
 (pg. 
205
-
207
)
128
Patterson
 
CJ
Soumerai
 
J
Hunter
 
Z
Leleu
 
X
Ghobrial
 
I
Treon
 
SP
Sildenafil citrate suppresses disease progression in patients with Waldenström's macroglobulinemia [abstract].
J Clin Oncol
2006
, vol. 
24
 
suppl 435S
Sign in via your Institution