ash 2009 meeting report—top 10 clinically oriented abstracts in chronic lymphocytic leukemia

26

Click here to load reader

Upload: susan-obrien

Post on 06-Jun-2016

215 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

ASH 2009 meeting report—Top 10 clinically orientedabstracts in myeloproliferative neoplasmsAyalew Tefferi*

The 2009 American Society of Hematology annual meeting took place

December 5–8, in New Orleans, Louisiana, USA. Official sources indi-

cate over 20,000 people attended the meeting and more than 4,000 sci-

entific abstracts were presented. Among the latter, close to 100 were

in the category of BCR-ABL1-negative myeloproliferative neoplasms

(MPN), including polycythemia vera, essential thrombocythemia and

primary myelofibrosis. Here, I summarize and discuss the top 10 MPN

abstracts, which I believe contain information that is immediately rele-

vant to clinical practice. All 10 abstracts are published in full in the

November 20, 2009 issue of Blood (volume 114, No. 22) and are identi-

fied here by their abstract numbers.

No. 1: A Phase I Evaluation of TG101348, a Selective JAK2

Inhibitor, in Myelofibrosis: Clinical Response Is Accompanied

by Significant Reduction in JAK2V617F Allele Burden

(Abstract # 755)Abstract summary

TG101348 is an oral JAK2-selective inhibitor with in vitro and in vivo activity

against JAK2V617F-driven myeloproliferation [1–4]. Fifty nine patients with pri-

mary myelofibrosis (PMF) or post-polycythemia vera/essential thrombocythemia

myelofibrosis (PV/ET MF) were treated with TG101348 in a Phase I/II study.

Among the 40 patients treated at or above the maximum tolerated dose (MTD;

680 mg/day), Grade 4 neutropenia or thrombocytopenia were rare. Dose-

dependent and expected on target effect on erythropoiesis was more prevalent

and attenuated with dose reduction. The majority of patients who started at

�680 mg developed mostly grade 1 or 2 nausea, vomiting, or diarrhea that

was self-limited or easily controlled. Other side effects included grade 1 or 2

transaminitis and asymptomatic elevation of serum creatinine and lipase. Thirty

three patients who started at �680 mg have completed at least three cycles of

treatment and during that period 67% experienced >50% reduction in palpable

spleen size and leukocytosis was controlled in all affected patients. Further-

more, among evaluable patients, 44% had a >50% reduction in JAK2V617F

allele burden. The drug also resulted in resolution of constitutional symptoms,

including early satiety, fatigue, cough, pruritus, and night sweats.

Discussion

For the first time, this study provides proof-of-principle for the utility of JAK2

inhibitors in the treatment of myeloproliferative neoplasms (MPN). To date,

TG101348 is the only JAK2 inhibitor shown to favorably affect JAK2V617F

allele burden, in addition to its remarkable activity in reducing spleen size,

controlling leukocytosis, and alleviating constitutional symptoms and pruritus.

It appears that the mechanism of TG101348 response involves direct

anticlonal myeloproliferation rather than a nonspecific suppressive effect on

proinflammatory cytokines that are often abnormally increased in MF [5].

No. 2: Long-Term Follow-up and Optimized Dosing Regimen

of INCB018424 in Patients with Myelofibrosis: Durable Clinical,

Functional and Symptomatic Responses with Improved

Hematological Safety (Abstract # 756)Abstract summary

INCB018424 is an oral JAK1 and JAK2 inhibitor [6]. One hundred and

fifty-five patients were treated in a Phase I/II study. Safe starting doses have

been determined to be 10 or 15 mg twice daily. At such doses, 48% of

evaluable patients achieved >50% reduction in palpable spleen size. The

drug also resulted in marked improvement of constitutional symptoms includ-

ing fatigue, abdominal discomfort, night sweats, and pruritus. Improvement

in symptoms correlated with marked reduction in serum proinflammatory

cytokines. Toxicity included thrombocytopenia and anemia whose incidences

was lower with the use of reduced drug doses and cytokine rebound phe-

nomenon with marked and rapid exacerbation of symptoms and signs of dis-

ease, in some patients, if the drug is discontinued for any reason.

Discussion

INCB018424 has shown clear benefit in alleviating severe constitutional

symptoms and cachexia in patients with MF and dramatically improves

patients’ sense of well being. This has been attributed to the drug’s

potent anticytokine activity, which should come in handy for the treatment

of other inflammatory conditions. In MF, the drug has the added benefit of

reducing spleen size. The drug has little effect on JAK2V617F allele

burden. Physicians using INCB018424 should be aware of the cytokine

rebound phenomenon during drug discontinuation and must use a taper-

ing schedule, similar to what is applied with systemic corticosteroid

therapy.

No. 3: RAD001, An Inhibitor of mTOR, Shows Clinical Activity

in a Phase I/II Study in Patients with Primary Myelofibrosis

and Post PolycythemiaVera/Essential Thrombocythemia

Myelofibrosis (Abstract # 307)Abstract summary

Mammalian target of rapamycin (mTOR), a serine/threonine protein

kinase, is a downstream effector of oncogenic signals. RAD001 (everolimus)

is an oral inhibitor of mTOR [7] and has been shown to inhibit JAK2V617F-

driven proliferation in both cell lines and primary myeloid progenitor cells

(abstract # 2914). The current abstract reports on a Phase I trial of oral

RAD001 in nine patients with JAK2V617F positive and negative MF. Cohorts

of three patients were treated with the drug at 5.0, 7.5, or 10 mg daily.

Toxicity included grade 2 mouth ulcers and hyperlipidemia. Reduction in

spleen size was seen in five patients. Pruritus disappeared in all five

affected patients. JAK2V617F allele burden was not affected.

Discussion

This abstract suggests therapeutically exploitable pathogenetic contribu-

tion of mTOR in MPN. Considering the fact that the drug is already FDA-

approved for the treatment of advanced renal cancer, [7] it is accessible for

an off-label use in some MPN patients whose quality of life has been marred

by intractable pruritus and constitutional symptoms.

No. 4: Pomalidomide Therapy in Myelofibrosis: 2-Year

Follow-up of a Randomized Phase 2 Study (Abstract # 1904)Abstract summary

This abstract provides long-term follow-up information in 16 patients with

PMF or post-PV/ET MF who had responded to oral pomalidomide therapy

(0.5–2 mg/day) during a previously published Phase 2 study of 84 patients [8].

It is to be recalled that response to pomalidomide was limited to improvement

in anemia (i.e., the drug had little effect on splenomegaly) and more likely to

occur in patients without leukocytosis or marked splenomegaly [8]. At a median

treatment duration of 19 months, 10 (63%) of the 16 treatment responders

were still in remission with a median response duration of 17 months (range,

13–24). The drug was neither neurotoxic nor myelosuppressive at the doses

used in this study. Among nine patients with available cytogenetic information,

all three with unfavorable karyotype relapsed, whereas none of the six with

favorable karyotype did.

Discussion

Pomalidomide is a new IMiD with therapeutic activity in both multiple mye-

loma and MF [8,9]. This abstract confirms the drug’s favorable side effect

profile and suggests durability of anemia response in the majority of treat-

ment responders, especially in the absence of unfavorable karyotype. How-

ever, pomalidomide therapy does not appear to favorably affect either sple-

nomegaly or bone marrow changes associated with MF. In another Ameri-

can Society of Hematology 2009 meeting abstract (# 2911), higher than 2

mg/day doses of the drug were shown to be associated with significant mye-

losuppression.

Letters

VVC 2009 Wiley-Liss, Inc.

American Journal of Hematology 190 http://www3.interscience.wiley.com/cgi-bin/jhome/35105

Page 2: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

No. 5: A Dynamic Prognostic Model to Predict Survival in

Primary Myelofibrosis: A Study of the International Working

Group for Myeloproliferative Neoplasm Research and

Treatment (Abstract # 3891)Abstract summary

The International Prognostic Scoring System (IPSS) for PMF identifies

five adverse risk factors (age >65 years, hemoglobin <10 g/dL, leukocyte

count >25 3 109/L, circulating blasts �1%, and constitutional symptoms) to

risk stratify patients into low, intermediate-1, intermediate-2, and high risk

disease [10]. This abstract examines the prognostic relevance of acquiring

each one of the five IPSS risk factors during the disease course. Five hun-

dred and twenty-five PMF patients were studied. The acquisition during the

disease course of each one of the five IPSS risk factors was found to be sig-

nificantly detrimental to survival. Subsequently, each variable was assigned

a score close to the corresponding hazard ratio (each one of the IPSS risk

factors was assigned a score of 1 except anemia that was assigned a score

of 2). The sum of these values was used to develop a dynamic IPSS

(DIPSS) that can be applied at any time during the disease course. The

authors also established age-adjusted DIPSS for patients less than age

65 years.

Discussion

This study validates IPSS, which is applicable at time of diagnosis, and

provides a modified version (i.e., DIPSS) for use at any time in the disease

course of patients with PMF. The age-adjusted DIPSS should come in handy

in facilitating treatment decision making, especially when allogeneic hemato-

poietic cell transplantation (AHCT) is considered.

No. 6: Transfusion Need at Diagnosis or Its Development

During the First Year of Diagnosis in Primary Myelofibrosis:

Effect On Survival and Correlation with JAK2 and TET2

Mutational Status (Abstract # 1909)Abstract summary

The main objective of this study was to examine the IPSS-independent

prognostic effect of transfusion need in PMF. Two hundred and fifty-four

patients were studied. Transfusion need at diagnosis was documented in

62 patients, whereas an additional 22 patients became transfusion

dependent during the 1st year of their diagnosis. Median survivals of

patients requiring transfusions at diagnosis or in their first year of diagno-

sis were equally poor and significantly worse than the median survival in

patients remaining transfusion-free at least in the 1st year postdiagnosis:

2.9, 2.2, and 9.7 years, respectively (P < 0.0001). Multivariable analysis

confirmed the IPSS-independent prognostic value of transfusion need

in PMF.

Discussion

This study is an extension of a previously published study [11] and

underscores the fact that transfusion need in PMF, as it is in MDS, [12] is

a marker of advanced disease with poor survival. Regardless of one’s

IPSS risk category, the need for red cell transfusion is an ominous sign

and warrants aggressive treatment such as AHCT or experimental drug

therapy.

No. 7: IPSS-Independent Cytogenetic Risk Categorization in

Primary Myelofibrosis (Abstract # 2909)Abstract summary

Two hundred patients with PMF were studied for the impact of karyotype

at diagnosis on survival. Eighty three (42%) patients displayed an abnormal

karyotype. Median survival in patients with sole 19 was not reached and in

those with sole del(13q), sole del(20q), normal karyotype, complex abnor-

malities and sole 18 were 112, 108, 80, 37, and 27 months, respectively,

while it was 46 months for patients with other cytogenetic abnormalities (P

5 0.01). Accordingly, sole abnormalities of 19, del(20q), and del(13q) were

categorized as being favorable (n 5 35) and complex abnormalities and sole

18 as unfavorable (n 5 20). Multivariable analysis confirmed the IPSS-inde-

pendent prognostic value of cytogenetic risk categorization.

Discussion

This study builds on observations from previous studies [13,14] and clari-

fies the prognostic role of cytogenetic information in PMF. Such information

is critical during treatment decision making. In case of a dry tap, one can

get cytogenetic information using peripheral blood karyotypic studies or mul-

tiprobe FISH analysis [15–17].

No. 8: A Phase 2 Study of INCB018424, An Oral, Selective

JAK1/JAK2 Inhibitor, in Patients with Advanced Polycythemia

Vera and Essential Thrombocythemia Refractory to

Hydroxyurea (Abstract # 311)Abstract summary

INCB018424 is a JAK1 and JAK2 inhibitor with anticytokine mediated ben-

eficial effect on constitutional symptoms and pruritus associated with MF [6].

The drug also reduces spleen size in MF. In this Phase II study, the drug

was used for the treatment with hydroxyurea intolerant or resistant PV (n 5

34) or ET (n 5 39). The optimal dose levels were determined to be twice

daily 10 mg for PV and 25 mg for ET. According to consensus criteria, 94%

of patients with PV and 61% with ET have achieved PR or CR. As expected

from the experience in MF, pruritus, constitutional symptoms, and splenome-

galy were favorably affected.

Discussion

Most of the results from this study were predictable based on the

drug’s previous activity in patients with MF including alleviation of consti-

tutional symptoms, pruritus, and reduction in spleen size. The drug’s

modest activity in ET as well as the less than optimal CR rates in both

ET and PV is consistent with its primary mode of action which is sup-

pression of proinflammatory cytokines. Broadly speaking, median sur-

vival in both ET and PV easily exceeds 15 years and most patients do

extremely well with current therapy that includes relatively inexpensive

and safe drugs such as aspirin and hydroxyurea [18–23]. Similarly,

patients who can not take hydroxyurea for one reason or another can be

effectively treated with interferon alpha [24,25] or busulfan [26–28].

Therefore, in the absence of selective activity against clonal myelopoie-

sis, it will be difficult to defend the use of new drugs in PV or ET.

No. 9: Chronic Eosinophilic Leukemia with

FIP1L1-PDGFRalpha Rearrangement (F/P): The Response

to Imatinib Is Durable. A Report of 33 Patients with a

Follow-up of 30 to 92 Months (Abstract # 3894)Abstract summary

In a prospective Phase 2 study, 33 patients with FIP1L1-PDGFRA-positive

MPN were treated with imatinib (IM) 100 to 400 mg daily and followed for a

median of 51 months (range 30–92). All but one were males. All patients

achieved a complete hematologic response in less than 1 month and molec-

ular remission in a median time of 3 months. All patients who continued IM

therapy remained in remission. The results were as good in patients receiv-

ing 100 mg daily. Discontinuation of IM treatment was almost always associ-

ated with reversible disease relapse.

Discussion

This study is comforting in terms of the long-term durability of response to

low doses of IM in patients with FIP1L1-PDGFRA-positive MPN. The study

also suggests the need for indefinite therapy. Additional studies are needed

to establish the minimum dose of IM that is required to maintain major

molecular remission and whether or not the latter is essential for preventing

disease progression. In this regard, a recent study suggested that a single

100 or 200 mg dose weekly was capable of maintaining hematological and

molecular remission [29].

No. 10: Platelet Antiaggregant Therapy Prevents Venous

Thrombosis in Patients with JAK2V617F Positive Essential

Thrombocythemia without Indication of Cytoreductive

Treatment (Abstract # 3906)Abstract summary

This study focuses on low-risk ET defined by age <60 years and

absence of thrombosis history. The main objective was to identify risk

factors for thrombosis. Three hundred patients were retrospectively

studied. During a median follow up of 8.6 years, a total of 32 thrombotic

events (arterial 21 and venous 11) were documented and the 5 year

thrombosis free survival in the absence of cytoreductive therapy was 94%

for arterial thrombosis and 96% for venous thrombosis. Smoking and ele-

vated serum LDH level at diagnosis, but not leukocytosis, JAK2 mutational

status or absence of treatment with aspirin, were significantly associated

with increased risk of arterial thrombosis. The risk of venous thrombosis

was higher in JAK2V617F-positive patients and appeared to be favorably

affected by the presence of aspirin therapy.

letters

American Journal of Hematology 191

Page 3: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

Discussion

The findings in this study provide support for current management of low-

risk ET where common sense measures such as cessation of smoking and

treatment with low dose aspirin are often considered adequate. Consistent

with a recently published Mayo Clinic study [30] and in contrast to another

Italian study, [31] leukocytosis was not relevant to thrombosis free survival.

The Mayo Clinic study in low-risk ET identified higher hemoglobin as a risk

factor for venous thrombosis [30].

Concluding Remarks

Between ET, PV, and PMF, the need for new therapy is most evident for

PMF and post-PV/ET MF. High risk patients with MF are currently offered

AHCT, which is unfortunately associated with high treatment-related mortal-

ity and morbidity [32]. Fortunately, recent data suggest improvement in sur-

vival for PMF patients diagnosed in the last 20 years compared to those

diagnosed earlier [33]. The reason behind this unexpected survival gains in

recent years is unclear but the specific observation underscores the inap-

propriateness of using historical controls to determine survival benefit for

new treatment modalities. At present, there is much activity regarding new

mutations [34–38] and new drugs [39] in MPN, and although it is easy to

jump on the band wagon of hype and misinformation, it takes courage (and

deeper understanding of disease natural history) to step back and carefully

interpret preliminary observations from new treatment modalities.

Division of Hematology, Mayo Clinic, Rochester, Minnesota*Correspondence to: Ayalew Tefferi, Division of Hematology,

Department of Medicine, Mayo Clinic, 200 First street SW, Rochester, MN 55905.E-mail: [email protected]

Received for publication 14 December 2009; Accepted 14 December 2009Conflict of interest: Nothing to report.

Published online 22 December 2009 in Wiley InterScience(www.interscience.wiley.com).

DOI: 10.1002/ajh.21620

References

1. Pardanani A, Hood J, Lasho T, et al. TG101209, a small molecule JAK2-selective kinase inhibitor potently inhibits myeloproliferative disorder-associ-ated JAK2V617F and MPLW515L/K mutations. Leukemia 2007;21:1658–1668.

2. Lasho TL, Tefferi A, Hood JD, et al. TG101348, a JAK2-selective antagonist,inhibits primary hematopoietic cells derived from myeloproliferative disorderpatients with JAK2V617F, MPLW515K or JAK2 exon 12 mutations as well asmutation negative patients. Leukemia 2008;22:1790–1792.

3. Wernig G, Kharas MG, Okabe R, et al. Efficacy of TG101348, a selective JAK2inhibitor, in treatment of a murine model of JAK2V617F-induced polycythemiavera. Cancer Cell 2008;13:311–320.

4. Geron I, Abrahamsson AE, Barroga CF, et al. Selective inhibition of JAK2-drivenerythroid differentiation of polycythemia vera progenitors. Cancer Cell 2008;13:321–330.

5. Tefferi A. Pathogenesis of myelofibrosis with myeloid metaplasia. J Clin Oncol2005;23:8520–8530.

6. Pardanani A. JAK2 inhibitor therapy in myeloproliferative disorders: Rationale,preclinical studies and ongoing clinical trials. Leukemia 2008; 22:23–30.

7. Atkins MB, Yasothan U, Kirkpatrick P. Everolimus. Nat Rev Drug Discov2009;8:535–536.

8. Tefferi A, Verstovsek S, Barosi G, et al. Pomalidomide is active in the treat-ment of anemia associated with myelofibrosis. J Clin Oncol 2009;27:4563–4569.

9. Lacy MQ, Hayman SR, Gertz MA, et al. Pomalidomide (CC4047) plus low-dosedexamethasone as therapy for relapsed multiple myeloma. J Clin Oncol2009;27:5008–5014.

10. Cervantes F, Dupriez B, Pereira A, et al. New prognostic scoring systemfor primary myelofibrosis based on a study of the International WorkingGroup for Myelofibrosis Research and Treatment. Blood 2009;113:2895–2901.

11. Tefferi A, Mesa RA, Pardanani A, et al. Red blood cell transfusion need at diag-nosis adversely affects survival in primary myelofibrosis-increased serum ferritinor transfusion load does not. Am J Hematol 2009;84:265–267.

12. Chee CE, Steensma DP, Wu W, et al. Neither serum ferritin nor the number ofred blood cell transfusions affect overall survival in refractory anemia with ringedsideroblasts. Am J Hematol 2008;83:611–613.

13. Hussein K, Huang J, Lasho T, et al. Karyotype complements the InternationalPrognostic Scoring System for primary myelofibrosis. Eur J Haematol 2009;82:255–259.

14. Hussein K, Van Dyke DL, Tefferi A. Conventional cytogenetics in myelofibrosis:Literature review and discussion. Eur J Haematol 2009;82:329–338.

15. Hussein K, Ketterling RP, Hulshizer RL, et al. Peripheral blood cytogenetic stud-ies in hematological neoplasms: Predictors of obtaining metaphases for analy-sis. Eur J Haematol 2008;80:318–321.

16. Hussein K, Ketterling RP, Dewald GW, et al. Peripheral blood cytogenetic stud-ies in myelofibrosis: Overall yield and comparison with bone marrow cytogeneticstudies. Leuk Res 2008;32:1597–1600.

17. Tefferi A, Meyer RG, Wyatt WA, Dewald GW. Comparison of peripheral bloodinterphase cytogenetics with bone marrow karyotype analysis in myelofibrosiswith myeloid metaplasia. Br J Haematol 2001;115:316–319.

18. Gangat N, Wolanskyj AP, McClure RF, et al. Risk stratification for survival andleukemic transformation in essential thrombocythemia: A single institutionalstudy of 605 patients. Leukemia 2007;21:270–276.

19. Gangat N, Strand J, Li CY, et al. Leucocytosis in polycythaemia vera predictsboth inferior survival and leukaemic transformation. Br J Haematol 2007;138:354–358.

20. Passamonti F, Rumi E, Pungolino E, et al. Life expectancy and prognostic fac-tors for survival in patients with polycythemia vera and essential thrombocythe-mia. Am J Med 2004;117:755–761.

21. Passamonti F, Rumi E, Arcaini L, et al. Prognostic factors for thrombosis, myelo-fibrosis, and leukemia in essential thrombocythemia: A study of 605 patients.Haematologica 2008;93:1645–1651.

22. Finazzi G, Barbui T. Evidence and expertise in the management of polycythemiavera and essential thrombocythemia. Leukemia 2008;22:1494–1502.

23. Palandri F, Catani L, Testoni N, et al. Long-term follow-up of 386 consecutivepatients with essential thrombocythemia: Safety of cytoreductive therapy. Am JHematol 2009;84:215–220.

24. Silver RT. Long-term effects of the treatment of polycythemia vera with recombi-nant interferon-alpha. Cancer 2006;107:451–458.

25. Kiladjian JJ, Cassinat B, Turlure P, et al. High molecular response rate of polycy-themia vera patients treated with pegylated interferon alpha-2a. Blood 2006;108:2037–2040.

26. Messinezy M, Pearson TC, Prochazka A, Wetherley-Mein G. Treatment of pri-mary proliferative polycythaemia by venesection and low dose busulphan: Ret-rospective study from one centre. Br J Haematol 1985;61:657–666.

27. Shvidel L, Sigler E, Haran M, et al. Busulphan is safe and efficient treatmentin elderly patients with essential thrombocythemia. Leukemia 2007;21:2071–2072.

28. Zittoun R. Busulfan versus 32P in polycythaemia vera. Drugs Exp Clin Res1986;12:283–285.

29. Helbig G, Moskwa A, Swiderska A, et al. Weekly imatinib dosage for chroniceosinophilic leukaemia expressing FIP1L1-PDGFRA fusion transcript: Extendedfollow-up. Br J Haematol 2009;145:132–134.

30. Gangat N, Wolanskyj AP, Schwager SM, et al. Leukocytosis at diagnosisand the risk of subsequent thrombosis in patients with low-risk essentiaLthrombocythemia and polycythemia vera. Cancer 2009;115:5740–5745.

31. Carobbio A, Finazzi G, Guerini V, et al. Leukocytosis is a risk factor for thrombo-sis in essential thrombocythemia: Interaction with treatment, standard risk fac-tors, and Jak2 mutation status. Blood 2007;109:2310–2313.

32. Siragusa S, Passamonti F, Cervantes F, Tefferi A. Survival in young patients withintermediate- / high-risk myelofibrosis: estimates derived from databases fornon transplant patients. Am J Hematol 2009;84:140–143.

33. Vaidya R, Siragusa S, Huang J, et al. Mature survival data for 176 patients youngerthan 60 years with primary myelofibrosis diagnosed between 1976 and 2005:Evidence for survival gains in recent years. Mayo Clin Proc 2009;84:1114–1119.

34. Tefferi A, Levine RL, Lim KH, et al. Frequent TET2 mutations in systemic masto-cytosis: clinical, KITD816V and FIP1L1-PDGFRA correlates. Leukemia 2009;23:900–904.

35. Tefferi A, Pardanani A, Lim KH, et al. TET2 mutations and their clinical corre-lates in polycythemia vera, essential thrombocythemia and myelofibrosis.Leukemia 2009;23:905–911.

36. Delhommeau F, Dupont S, Della Valle V, et al. Mutation in TET2 in myeloid can-cers. N Engl J Med 2009;360:2289–2301.

37. Tefferi A, Lim KH, Abdel-Wahab O, et al. Detection of mutant TET2 in myeloidmalignancies other than myeloproliferative neoplasms: CMML, MDS, MDS/MPN and AML. Leukemia 2009;23:1343–1345.

38. Carbuccia N, Murati A, Trouplin V, et al. Mutations of ASXL1 gene in myeloproli-ferative neoplasms. Leukemia 2009;23:2183–2186.

39. Tefferi A. Essential thrombocythemia, polycythemia vera, and myelofibrosis:Current management and the prospect of targeted therapy. Am J Hematol2008;83:491–497.

letters

192 American Journal of Hematology

Page 4: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

ASH 2009 meeting report—Top 10 clinically-orientedabstracts in hematopoietic stem cell transplantationDavid L. Porter*

Paradigms in allogeneic stem cell transplantation (SCT) have been shift-

ing for several years. At the 2009 ASH meeting, over 1,200 abstracts pre-

sented data dealing with transplantation. A large number of presenta-

tions continue to build on several important themes that generate great

interest and controversy in the field. It is misleading, despite the title of

this summary, to suggest that there is a ‘‘Top 10’’ list of transplant-related

abstracts. There was an incredible breadth and depth of work presented

this year that goes a long way in moving the transplant field forward.

Many more abstracts that can be presented here are worthy of note, rec-

ognition, and discussion, and the exclusion of many important studies

from this review is a reflection of space limitations rather than quality of

work. Nevertheless, there were a number of repetitive themes at ASH,

and this highlight attempts to review abstracts that at least touch on sev-

eral of the important and increasingly controversial issues. All 10

abstracts are published in full in the November 20, 2009 issue of Blood

(volume 114, No. 22) and are identified here by their abstract number.

Adoptive Immunotherapy With Tregs Prevents GvHD and

Favors Immune Reconstitution After HLA Haploidentical

Transplants for Hematological Malignancies (Abstract #4)

The only available curative option for many patients with hematologic malig-

nancies is allogeneic SCT, but often well matched donors can not be identified. If

effective and safe, haploidentical donors would dramatically increase the donor

pool for almost every patient. To limit the risk of severe GVHD after haploidenti-

cal SCT, intensive T cell depletion is often performed [1]. Unfortunately GVHD

remains significant and the extensive T cell depletion causes delayed immune

reconstitution resulting in a high incidence of infection-related deaths. Preclinical

models demonstrate that CD41/CD251/FoxP31 regulatory T cells (Treg) can

control GVHD and may promote immune reconstitution [2]. In the plenary ses-

sion, Dr Martelli presented results from a phase I/II trial of haploidentical trans-

plantation designed to evaluate the role of an engineered graft consisting of

immunoselected CD341 cells, mature donor T cells, and isolated donor Tregs.

Twenty-two adult patients with a median age of 40 years were treated for high

risk or relapsed AML (n 5 17), ALL (n 5 4) or NHL (n 5 1). All but one patient

received 23 106/kg Tregs and one patient received 13 106/kg Tregs isolated by

immunoselection. No other GVHD prophylaxis was used. Engraftment was

prompt in 20/22 patients with achievement of full donor chimerism. Of 20 evalu-

able patients, two had grade I GVHD and only one had grade III acute GVHD

(this patient received the lowest dose of Treg). There were four deaths (20%)

from severe infectious complications. However, the authors report reconstitution

with a high frequency of both CD41 and CD81 T cells specific for opportunistic

pathogens such as Aspergillus, Candida, CMV, ADV, HSV, VZV, and Toxo-

plasma. They also noted a low incidence of CMV reactivation.

This presentation showed that it is possible to isolate and administer a

purified population of Tregs at relatively high doses, GVHD was minimal,

and immune reconstitution assays showed rapid recovery of potentially

important effector T cells. There was still a relatively high risk of death from

opportunistic infections that was worrisome. Nevertheless, this trial should

lead to a better understanding of the effects of Tregs on GVHD, immune

recovery and GVL activity. Further refinement of the selection and expansion

process for Tregs and other cells may lead to improved outcomes and

enhance immune recovery after haploidentical SCT.

Loss of Mismatched HLA as a Mechanism of Leukemia Immune

Escape in Family Haploidentical and Unrelated HSCT: Analysis

of 103 Transplants FromAlternative Donors (Abstract 203)

Relapse remains a major limitation to allogeneic SCT, and mechanisms

causing relapse are poorly defined. Furthermore, therapy for relapse after

transplant is often ineffective [3,4]. Designing new immunologic strategies

for relapse is dependent on understanding the biology of graft-vs-leukemia

(GVL) induction. These topics were recently emphasized in the NCI-spon-

sored conference on ‘‘Prevention, and Treatment of Relapse After Allogeneic

Hematopoietic Stem Cell Transplantation’’. In a rather striking finding, Vago

et al. reported earlier this year that in the setting of haploidentical SCT, loss

of the patient-specific mismatched HLA haplotype occurred in patients with

relapse but not in patients without relapse [5]. This abstract expands these

findings surveying 103 recipients of partially HLA-matched transplantation,

including 67 patients with haploidentical donors and 35 patients with unre-

lated donors (URD). Relapse occurred in 28/67 (42%) patients after haploi-

dentical SCT and in 8/35 (23%) patients after URD SCT. In 10 of 28 cases

of relapse after haploidentical transplantation, mutations were found in leu-

kemia cells resulting in loss of the patient-specific HLA haplotype. No loss of

HLA haplotypes were noted in the eight patients who relapsed after URD

transplantation.

This report significantly expands on this group’s earlier publication and

shows that loss of the patient-specific haplotype is common after haploidenti-

cal transplant but not after URD SCT, and is likely an important mechanism

leading to relapse. Importantly, this information allows alternative therapy for

relapse; it would be anticipated that DLI would have no effect in this group of

patients. Rather than losing time and risking toxicity from DLI, 5 of the 10

relapsed patients received second transplants from a different donor mis-

matched for the remaining haplotype in hopes of generating GVL activity. This

report is particularly significant since it defines an important mechanism of

relapse after haploidentical SCT, and uses this information to select treatment

strategies for relapse.

Reduced-Intensity Conditioning (RIC) Allogeneic Stem Cell

Transplantation (allo-SCT) for Patients Aged �60 Years:

A Retrospective Study of 629 Patients From the Societe

Francaise De Greffe De Moelle Et De Therapie Cellulaire

(SFGM-TC) (Abstract 194)

The role of reduced intensity conditioning (RIC) and allogeneic SCT in

older patients is a particularly important topic. Available data suggests that

outcomes may be similar after RIC SCT than after more conventional trans-

plant [6–8], though the contribution of patient age, diagnosis, disease status,

and many other factors are poorly defined. As yet, no randomized trials are

available to direct decision making. The French Societe Francaise De Greffe

De Moelle Et De Therapie Cellulaire have provided an important addition to

the reported experience with a large retrospective study on role of RIC allo

SCT in 629 patients transplanted between 1998 and 2008 All patients were

�60 years old, and outcomes were compared between patients 60–65 years

and patients >65 years old. Although the indications for transplant were het-

erogeneous, importantly, 76% had high risk disease features, and 24% had

a standard risk disease. Sixty-one percent received a graft from an HLA-

matched related donor, while 199 (32%) received the graft from a matched-

unrelated donor. Disease characteristics were well matched between the

516 patients aged 60–65 and the 113 patients age >65. The follow-up was

relatively short (median 9 months (range, 1–90). Grade II-IV acute GVHD

developed in 29% and chronic GVHD in 23%. For patients 60–65, TRM was

29% compared to 27% in patients >65. For the entire group, disease related

mortality was 16% and overall survival (OS) at 1 and 2 years was 57%

(95%CI, 53–62%) and 47% (95%CI, 42–52%), respectively. There were no

significant differences is GVHD incidence, TRM, or OS in the different age

groups and Cox multivariate analysis showed that age >65 years was not a

significant factor associated with survival.

Many unanswered questions surround the application of RIC allogeneic

SCT in older patients including the impact of patient age of survival. This

retrospective analysis, in an older group of patients with predominantly high

risk hematologic malignancies, demonstrates impressive 1 and 2 year sur-

vival rates and found no difference in outcomes for patients 60–65 or >65.

This lends important support to include older patients in future clinical trials

letters

American Journal of Hematology 193

Page 5: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

and for offering potentially curative allogeneic SCT to patients beyond 65

years old. Consideration of other factors such as co-morbidities and func-

tional status is likely to be more important than age.

Allogeneic Hematopoietic Stem Cell Transplantation (HCT)

Compared to Chemotherapy Only in Acute Myeloid Leukemia

(AML) Patients 60 Years and Older: A Center for International

Blood and Marrow Transplantation Research (CIBMTR)/Cancer

and Leukemia Group B (CALGB) Study (Abstract 657)

Much of the data regarding the role of RIC SCT, and in particular compa-

rative studies to non-transplant therapies, is limited by small numbers of

patients studied, [9], patient heterogeneity, and selection bias [10]. In an

effort to account for this, the CIBMTR and CALGB analyzed outcomes in

100 patients with AML 60–70 years old who received RIC allogeneic SCT in

CR1 and compared their outcomes to 96 patients treated with standard

induction on two sequential CALGB protocols (9,720 and 10,201) between

1998 and 2006. Somewhat different than other comparative studies, chemo-

therapy-only patients had to remain in CR1 for at least 4 months to be

included in order to reduce selection bias. The groups were relatively well

matched except for statistically significant differences (though not necessa-

rily clinically meaningful) in age and duration to CR; transplant recipients

were slightly younger (median age 63 v 65 years; P < 0.001;) and the time

from diagnosis to CR1 was slightly longer for transplant patients (median 46

v 38 days; P 5 0.007;). Grafts were obtained from both HLA-identical sib-

lings (48%) and closely matched unrelated donors (52%). In addition follow-

up was longer in the chemotherapy-treated patients than that of SCT

patients (51 v 30 months). There was a trend towards improved 3 year LFS

for transplanted patients (34% (95% CI, 24–44%)) compared to chemother-

apy-treated patients (17% (95% CI, 10–25%), P 5 0.06). This was due to a

higher relapse rate of 86% in chemotherapy patients compared to 29% in

recipients of RIC SCT (P < 0.001). Non-relapse mortality was higher in

transplant recipients compared to chemotherapy-treated patients (39% v

17%, P < 0.001). However these factors balanced each other and overall

survival from CR1 did not differ between the groups (P 5 0.47), with 3 year

survival estimates of 35% (95% CI, 25–46%) for transplant patients com-

pared to 25% (95% CI, 17–34%) for chemotherapy patients.

This is an important large retrospective analysis comparing RIC SCT to

non-transplant therapy for patients in CR1. The inclusion of a uniform diag-

nosis (AML) and the efforts made to eliminate bias adds strength to this

report. The survival data in this study is reasonable, and similar to other tri-

als of RIC transplant in older AML patients [10–12]. The low relapse rate

associated with RIC SCT supports the presence of a potent GVL effect after

RIC SCT in older patients with AML. Continued efforts need to be directed

at minimizing treatment related mortality before obvious benefits for trans-

plant are achieved.

Prospective Evaluation of An Agvhd-Specific Proteomic

Pattern in More Than 340 Patients (Abstract 347)

Graft-vs-host disease (GVHD) remains a major limitation to successful

allogeneic SCT. Initial treatment for acute GVHD is effective for approxi-

mately 50% of patients but is often inadequate [13]. Ultimately, preventing or

pre-empting GVHD is likely to be more effective than trying to treat estab-

lished disease. However, to date there is no test predictive of impending

GVHD though many investigators have tested numerous potential bio-

markers [14]. Dr Mischak-Weissinger and her colleagues presented data

showing that proteomic analysis can identify patterns predictive of subse-

quent GVHD. In a blinded evaluation, y 961 samples from 345 patients

undergoing allogeneic SCT at 8 different centers were studied. Most patients

were transplanted for hematological malignancies and a variety of condition-

ing and GVHD prophylactic regimens were used. A panel of 31 different

peptides as selected based on prior work showing that these peptides were

either increased/decreased or present/absent with acute GVHD [15]. This

large study correctly classified patients with acute GVHD �7 days prior to

the development of clinical symptoms for with a sensitivity 76% and specific-

ity of 85%.

Identifying non-invasive biomarkers for GVHD is critically important for

patient care, to allow for earlier (and potentially more effective and safer)

therapy, and to monitor response. Ultimately this would serve to limit toxicity

from both GVHD and immunosuppression. This study not only verified this

group’s previous results, but showed that altered protein patterns were iden-

tified almost a week before clinical symptoms developed. Finding a suffi-

ciently sensitive and specific proteomics pattern that can be analyzed with

rapid turn around has led this group appropriately to initiate a multicenter

trial testing the efficacy of pre-emptive therapy on the incidence and severity

of acute GVHD and determining the impact of this strategy on overall survival.

Allogeneic Hematopoietic Cell Transplantation Can Cure

Some Patients with Acute Leukemia in Relapse or Primary

Induction Failure: A CIBMTR Study (Abstract 528)

Refractory leukemia is uniformly fatal without allogeneic SCT. However,

the role of transplant in these patients is controversial [16,17]. Duval and

colleagues performed a retrospective study of 2255 adult recipients of a

myeloablative allogeneic SCT reported to the CIBMTR between 1995 and

2004 with either refractory AML (n 5 1,673) or ALL (n 5 582). Over 40% of

patients had more than 25% blasts and almost half of all patients had a Kar-

nofsky performance status (KPS) of <90. Mortality by day 100 was high

(39% in AML and 41% in ALL) and overall survival low (19% (CI 17–21) for

AML and 16% (CI 13–20) for ALL patients. Leukemia was the cause of

death in 42% of AML patients and 37% of ALL patients. Five pre-transplant

factors were significantly associated with survival: first remission <6 months,

blasts in the blood, donor other than HLA-identical sibling or partially matched

unrelated donor, KPS <90%, and poor-risk cytogenetics. For the 106 AML

patients with no high risk criteria, the probability of 3 year OS was 44%

(35–54). For ALL, factors associated with worse survival included primary

refractory disease or �2nd relapse, �25% marrow blasts, CMV seroposi-

tive recipient, and age �10 years. Patients with 3 or more disease-specific

risk factors had a probability of survival of 6% for AML and 9% for ALL, or

as Dr Duval reported, survival was ‘‘approaching the definition of futility’’.

The approach to patients with refractory leukemia is always difficult and

controversial. This very large, though retrospective analysis from the

CIBMTR provides important insight into factors that will influence the ulti-

mate outcome of transplant. For patients with few risk factors, anticipated

outcomes are quite reasonable. For patients with multiple risk factors, alter-

native therapies and supportive care may be more appropriate.

Long-Term Survival and Late Deaths in 2-Year Survivors of

Myeloablative Allogeneic Hematopoietic-Cell Transplantation

for Hematologic Disorders (Abstract 520)

A large number of trials in allogeneic SCT report outcomes at 2 years and

occasionally even out to 5 years [18]. Data on long-term outcomes are lim-

ited, and there remains a concern that relapse, chronic GVHD, and long-

term organ toxicity shorten patient’s survival after a presumed ‘‘cure’’. To

study long term outcomes, Dr Wingard and colleagues from the CIBMTR

identified 10,632 patients who received a myeloablative allogeneic HCT

through 2003 and survived in CR at least 2 years after transplant. The

patient population is summarized in Table I.

At 10 years, the estimated probability of overall survival was 84% (95% CI,

82–85%) for AML, 84% (82–85%) for ALL, 80% (77–83%) for MDS, 84% (81–

87%) for lymphoma and 92% (91–93%) for SAA. Late relapses were significant

and the most common cause of death for AML, ALL, MDS, and lymphoma.

The cumulative incidence of relapse at 10-years post-HCT was 10% (9–11%)

for AML, 9% (8–10%) for ALL, 10% (8–12%) for MDS and 6% (4–8%) for lym-

phoma. GVHD was the most common cause of death for SAA. Older age at

transplant and chronic GVHD were both associated with greater late mortality

for all diseases. At 15-years after transplants, the relative mortality of patients

who had received SCT for AML, ALL, MDS and SAA was significantly higher

than in age-, race- and gender-matched normal populations. Mortality rates for

lymphoma patients were not significantly different than those of the matched

general population after 8-years post-HCT.

This study represents one of the longest follow-up reports after allogeneic

SCT. It gives a reasonable estimate of expectations and demonstrates that

patients alive in CR 2 years after transplant continue to have a excellent

prognosis though survival is shorter compared to a matched normal popula-

tion (for all but patients with lymphoma). Further studies should continue to

analyze long term complications and take into account quality of life for

transplant survivors.

letters

194 American Journal of Hematology

Page 6: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

Early Human Herpes Virus Type 6 Reactivation in Patients

Undergoing Allogeneic Stem Cell Transplantation (allo-SCT)

(Abstract 794)

Despite more than 30 years of experience, a great deal remains unknown

regarding the mechanisms leading to acute GVHD and other toxicity after allo-

geneic SCT. Human herpes virus 6 (HHV6) frequently reactivates after alloge-

neic SCT but the contribution of this virus to clinical outcomes remains a subject

of debate [19]. Dr Dulery and colleagues report on a study involving 225 con-

secutive patients receiving either matched sibling (n 5 110) or unrelated donor

(n 5 115) bone marrow (n-124), PBSC (n 5 69) or cord blood (n 5 32) allog-

rafts and measured HHV6 plasma loads by weekly quantitative PCR. The con-

ditioning regimen was myeloablative in 151 patients and reduced intensity in 74

patients. Patients were compared in 2 groups; 105 patients (46 patients had

either no reactivation or a positive PCR more than 100 days from SCT. Of the

105 patients with early reactivation, 10 were asymptomatic and 95 had clinical

symptoms that included fever, rash, diarrhea, pulmonary or neurological compli-

cations, liver abnormalities or other medical issues. Compared to patients with

no or late HHV6 reactivation, patients with early HHV6 reactivation had delayed

platelet engraftment (28 vs 22 days (P 5 0.003) and more grade III-IV acute

GVHD (32% vs 12%, P 5 0.008). Interestingly, HHV6 PCR was positive in 53%

of patients with grade III-IV acute GVHD before onset of symptoms. In multivari-

able analysis, the most important factors influencing acute grade III-IV GvHD

were early HHV6 reactivation, [HR: 2.10; 95%CI, 1.20-3.66] (P 5 0.011) and

use of an unrelated donor graft [HR: 1.17; 95%CI 1.04–1.32] (P5 0.007)

Mechanisms leading to acute GVHD and other organ toxicity are complex

and not completely understood. Though HHV6 reactivation is frequent after

allogeneic SCT, the relationship to clinical symptoms has been difficult to

define. This study confirms frequent reactivation of HHV6 after allogeneic

SCT and shows that it is associated with delayed platelet engraftment and

severe acute GVHD. This suggests that monitoring for HHV6 by PCR with

potential early intervention could positively impact on transplant outcomes.

Whether preemptive therapy for HHV6 will have an impact similar to the dra-

matic advances made with CMV therapy over the past decade will need to

be determined by prospective and randomized trials.

Association Between Genetic Variants in the Base Excision

Repair Pathway and Outcomes After Hematopoietic Cell

Transplant (Abstract 870)

It is widely known that allogeneic SCT is associated with extensive morbidity

and mortality, but mechanisms leading to organ damage in some patients but

not others are not well understood. In an interesting presentation, Arora and col-

leagues hypothesized that since conditioning therapies induce DNA damage,

genetic variation in base excision repair (BER) pathway genes could influence

transplant outcomes. They evaluated the association of 179 single nucleotide

polymorphisms (SNPs) in BER pathway genes with transplant related mortality

(TRM) at 1 year and relapse in 470 recipients of allogeneic SCT for hematologic

malignancies between 1998 and 2007. Grafts were from HLA matched siblings,

matched or mismatched unrelated donors or single umbilical cord blood units.

They found that after adjustment for age at transplant, donor type, race, and con-

ditioning regimen, four specific SNPs were associated with increased risk of

TRM and two were associated with lower risk of TRM at 1 year. Furthermore,

patients with more deleterious alleles in the BER pathway showed an increased

cumulative incidence of TRM at one year (14% for �1 deleterious allele vs. 51%

for �4 deleterious alleles; P < 0.001). One SNP was associated with decreased

risk of disease relapse (P < 0.001) after SCT.

This study showed that specific SNPs in genes associated with the BER

pathway could be associated with protection from, or worsening of, TRM

and relapse after allogeneic SCT. This supports the hypothesis that hetero-

geneity in the BER pathway could lead to variable DNA and tissue damage

from conditioning therapy and impact survival. If confirmed, SNP analysis

would be a rapid and efficient biomarker for risk assessment and treatment

planning prior to transplant. Furthermore, more detailed studies may provide

clearer insight into the mechanisms associated with tissue damage and

relapse after transplant.

Reducing the Risk for Transplant Related Mortality After

Allogeneic Hematopoietic Cell Transplantation: How Much

Progress Has Been Made? (Abstract 649)

In one of the more optimistic presentations from this meeting, Dr. Horan pre-

sented data on behalf of the CIBMTR showing that significant progress has

been made in reducing the risk of treatment related mortality after allogeneic

SCT. The last 25 years have been highlighted by significant advances in sup-

portive care and infection treatment, newer therapies for GVHD, and the wide

application of the therapy designed to limit regimen-related toxicity. To deter-

mine the impact of practice changes over time on TRM and survival, this

abstract reported on outcomes in 5,972 patients who received bone marrow or

peripheral blood stem cell grafts between 1985 and 2004. All patients received

a myeloablative conditioning regimen for AML in either the first or second

remission and were younger than 50 years old. Outcomes were determined for

4 consecutive 5 year intervals and appropriate adjustments were made for any

differences in characteristics over time. For recipients of matched related

donor grafts, there was a steady drop in TRM with a significant reduction in the

risk of death. The reduction was most pronounced in the matched related

donor group in CR2. Treatment related mortality also improved in the later time

periods for recipients of unrelated donor grafts though no overall survival bene-

fit was noted for this group in first complete remission. Patients in CR2 had

a lower risk of mortality in the later years compared to the earlier time

periods (RR 0.74 (range 0.6–0.9, P 5 0.03). Subgroup analyses was done

to account for major practice changes over time and showed similar benefits

when the analysis was limited to recipients of bone marrow grafts only,

cyclosporine/methotrexate as GVHD prophylaxis, or partially matched URD grafts.

These results suggest that improvements in outcomes over time were not

solely related to the use of PBSC grafts, new GVHD prophylactic regimens,

or better HLA typing technology, but rather an overall improvement in

care of these patients. Although TRM improved, it remains relatively high,

particularly after unrelated donor transplantation, and all would agree that

there is significant room for continued improvement. Nevertheless, this

report from the CIBMTR serves as a barometer of progress within the trans-

plant community. The improved outcomes over this relatively short time

period is a testament to the perseverance of thousands of clinicians, scien-

tists, nurses, research staff, and countless other professionals, and a tribute

to the courage of thousands of patients who have trusted their lives to the

potential of hematopoietic stem cell transplantation.

Blood and Marrow Transplantation, Division of Hematology-Oncology,University of Pennsylvania Medical Center, Philadelphia, Pennsylvania 19104

Correspondence to: David L Porter, Blood and Marrow Transplantation,Division of Hematology-Oncology, University of Pennsylvania Medical Center,3400 Civic Center Boulevard, PCAM 2 West Pavilion, Philadelphia, PA 19104

E-mail: [email protected] grant sponsor: NIH; Contract grant number: 5K24CA117879-04.

Published online 23 December 2009 in Wiley InterScience(www.interscience.wiley.com).

DOI: 10.1002/ajh.21628

References

1. Aversa F, Terenzi A, Tabilio A, et al. Full haplotype-mismatched hematopoieticstem-cell transplantation: A Phase II study in patients with acute leukemia athigh risk of relapse. J Clin Oncol 2005;23:3447–3454.

2. Riley JL, June CH, Blazar BR, et al. Human Tregulatory cell therapy: Take a bil-lion or so and call me in the morning. Immunity 2009;30:656–665.

3. Saiko K, Takahiro F, Kinuko T, et al. Outcome of 93 patients with relapse or pro-gression following allogeneic hematopoietic cell transplantation. Am J Hematol2009;84:815–820.

TABLE I. The Patient Population

Median follow up 9 years (range, 2–31)

Survivors for �10-years 37%

Survivors for �15-years 12%

AML N 5 4,017

ALL N 5 2,895

MDS N 5 930

SAA N 5 2,171

Lymphoma N 5 619

Donors HLA identical siblings 72%

Donors unrelated 22%

Age <20 45%

Age >40 17%

TBI containing regimen 60%

Acute GVHD grade II–IV 39%

Chronic GVHD by 2 years 43%

letters

American Journal of Hematology 195

Page 7: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

4. Loren AW, Porter DL. Donor leukocyte infusions for the treatment of relapsedacute leukemia after allogeneic stem cell transplantation. Bone Marrow Trans-plant 2008;41:483–493.

5. Vago L, Perna SK, Zanussi M, et al. Loss of mismatched HLA in leukemia afterstem-cell transplantation. [see comment]. New Engl J Med 2009;361:478–488.

6. Martino R, Iacobelli S, Brand R, et al. Retrospective comparison of reduced-intensity conditioning and conventional high-dose conditioning for allogeneichematopoietic stem cell transplantation using HLA-identical sibling donors inmyelodysplastic syndromes. Blood 2006;108:836–846.

7. Flynn CM, Hirsch B, Defor T, et al. Reduced intensity compared with high doseconditioning for allotransplantation in acute myeloid leukemia and myelodysplas-tic syndrome: A comparative clinical analysis. Am J Hematol 2007;82:867–872.

8. Tomblyn M, Brunstein C, Burns LJ, et al. Similar and promising outcomes in lym-phoma patients treated with myeloablative or nonmyeloablative conditioning andallogeneic hematopoietic cell transplantation. Biol Blood Marrow Transplant2008;14:538–545.

9. de Lavallade H, El-Cheikh J, Faucher C, et al. Reduced-intensity conditioningallogeneic SCTas salvage treatment for relapsed multiple myeloma. Bone Mar-row Transplant 2008;41:953–960.

10. Estey E, de Lima M, Tibes R, et al. Prospective feasibility analysis of reduced-intensity conditioning (RIC) regimens for hematopoietic stem cell transplantation(HSCT) in elderly patients with acute myeloid leukemia (AML) and high-riskmyelodysplastic syndrome (MDS). Blood 2007;109:1395–1400.

11. Valcarcel D, Martino R, Caballero D, et al. Sustained remissions of high-risk

acute myeloid leukemia and myelodysplastic syndrome after reduced-intensity

conditioning allogeneic hematopoietic transplantation: Chronic graft-versus-host

disease is the strongest factor improving survival. J Clin Oncol 2008;26:577–584.

12. Tauro S, Craddock C, Peggs K, et al. Allogeneic stem-cell transplantation

using a reduced-intensity conditioning regimen has the capacity to produce

durable remissions and long-term disease-free survival in patients with

high-risk acute myeloid leukemia and myelodysplasia. J Clin Oncol 2005;

23:9387–9393.13. MacMillan M, Weisdorf D, Wagner J, et al. Response of 443 patients to steroids

as primary therapy for acute graft-versus-host disease: Comparison of grading

systems. Biol Blood Marrow Transplant 2002;8:387–394.14. Mathias C, Mick R, Grupp S, et al. Soluble interleukin-2 receptor concentra-

tion as a biochemical indicator for acute graft-versus-host disease after

allogeneic bone marrow transplantation. J Hematother Stem Cell Res

2000;9:393–400.15. Weissinger EM, Schiffer E, Hertenstein B, et al. Proteomic patterns predict

acute graft-versus-host disease after allogeneic hematopoietic stem cell trans-plantation. Blood 2007;109:5511–5519.

16. Song KW, Lipton J. Is it appropriate to offer allogeneic hematopoietic stem celltransplantation to patients with primary refractory acute myeloid leukemia?Bone Marrow Transplant 2005;36:183–191.

17. Oyekunle AA, Kroger N, Zabelina T, et al. Allogeneic stem-cell transplantation inpatients with refractory acute leukemia: A long-term follow-up. Bone MarrowTransplant 2006;37:45–50.

18. Weisser M, Ledderose G, Jochem Kolb H, et al. Long-term follow-up of alloge-neic HSCT for CML reveals significant improvement in the outcome over the lastdecade. Ann Hematol 2007;86:127–132.

19. Zerr DM, Corey L, Kim HW, et al. Clinical outcomes of human herpesvirus 6reactivation after hematopoietic stem cell transplantation. Clin Infect Dis2005;40:932–940.

ASH 2009 meeting report—Top 10 clinically oriented abstractsin myelodysplastic syndromes

David P. Steensma*

Among the 4,097 abstracts presented orally or in poster format at the

December 2009 American Society of Hematology (ASH) annual meeting

in New Orleans, LA, 139 abstracts (36 presentations in six oral ses-

sions and 103 posters in three poster sessions) were assigned to the

general category of myelodysplastic syndromes (MDS). Here, I summa-

rize and discuss 10 clinically relevant abstracts related to MDS, all of

which are published in their entirety in the November 20, 2009 issue of

Blood (volume 114, issue 22) and are identified here by their abstract

numbers. These 10 abstracts—six oral presentations and four post-

ers—concern prognostic modeling, treatment with hypomethylating

agents (azacitidine or decitabine), and therapy with immunomodulatory

or immunosuppressive drugs (lenalidomide or alemtuzumab).

Section 1: Prognostic FeaturesAbstract #2772: Cytogenetic risk features in myelodysplastic

syndromes—Update and present state

More than 600 distinct karyotypes have been observed in patients with

myelodysplastic syndromes (MDS) [1–3]. However, because many of these

MDS-associated clonal cytogenetic abnormalities are uncommon or rare,

their prognostic importance has been unclear. The 1997 International Prog-

nostic Scoring System (IPSS), based on a review of 816 patients with pri-

mary MDS (i.e., the International Myelodysplasia Risk Analysis Workshop

(IMRAW) cohort), identified just six prognostically useful karyotypes: a nor-

mal karyotype, isolated deletion of the long arm of chromosome 20 [del(20q)],

isolated del(5q), loss of the Y chromosome, abnormalities of chromosome 7,

and complex karyotypes (defined by the IPSS as �3 abnormalities) [4].

In 2007, a German–Austrian Consortium, led by Haase et al. from the

University of Gottingen, published a 4-tier risk model incorporating a more

comprehensive cytogenetic database (39 karyotypic subgroups were

defined), based on a review of 2,124 cases from eight centers, including

1,084 patients (52%) with an abnormal karyotype [2]. Patients treated with

supportive care or chemotherapy were included in the German–Austrian

analysis, but those treated with lenalidomide were excluded; 93.3% had pri-

mary MDS. The validity of the German–Austrian risk model was confirmed

in 2007 using an independent 1,738 patient dataset from the MD Anderson

Cancer Center (MDACC) in Houston [5].

At the 2009 ASH meeting, Julie Schanz from Gottingen reported an

update of the German–Austrian model, derived by combining existing con-

sortium data with the IMRAW cohort and data from the Spanish Cytoge-

netics Working Group (GCECGH)—cumulatively comprising 3,803 patients.

Additionally, 53 cases with rare cytogenetic abnormalities were contributed

by the International Cytogenetics Working Group (ICWG) of the MDS Foun-

dation, increasing the total number of patients to 3,856. Based on prognos-

tic modeling of this large cohort, more than 30 different MDS-associated

karyotype patterns were stratified into four different risk groups (Table I),

with median survival ranging from 5.7 months for poor-risk karyotypes to

50.6 months for good-risk karyotypes. Notably, the placement of a number

of karyotypes within this new 4-tier risk model differs from the 2007 publica-

tion of the German–Austrian Consortium and from data presented by Pro-

fessor Haase at the MDS International Symposium in Patras, Greece in

May 2009.

The 2008 World Health Organization (WHO) classification of hematologi-

cal neoplasms [6] revised the MDS diagnostic category of ‘‘MDS-unclassifi-

able’’ to include cases in which diagnostic morphologic features for MDS are

not present, yet 1 of 13 specific clonal MDS-associated karyotypic abnor-

malities is detected (Table II) [8,9]. Isolated trisomy 8, loss of the Y chromo-

some, and del(20q) were specifically excluded from this list, as these find-

ings are less specific for MDS (and, in the case of loss of the Y chromo-

some, can be seen in healthy people [10]). It seems likely that future

revisions of the WHO classification will expand the list of diseases-defining

MDS-associated karyotypes.

Another 2009 ASH annual meeting abstract (#3817), presented by the

DACO-020 (ADOPT) decitabine trial [11] investigators, correlated karyotypic

response to therapy with improved survival, compared with patients who

experienced a clinical response to decitabine but continued to have a cyto-

genetic abnormality, or those who failed to have any response to the drug.

This is not a surprising finding, but to my knowledge, it is the first time that

the importance of a cytogenetic response to therapy has been formally dem-

onstrated in MDS outside of the stem cell transplant setting.

Abstract #945: Survival, prognostic factors, and rates of

leukemic transformation in a multicenter study of

303 untreated patients with MDS and Del(5q)

There has been considerable confusion over the diagnostic boundaries of

del(5q) MDS and the 5q-syndrome [a subset of del(5q) MDS with isolated

del(5q), <5% blasts, normal or increased platelet count, and a characteristic

marrow morphology including monolobated megakaryocytes] [12]. In particu-

letters

196 American Journal of Hematology

Page 8: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

lar, most series of del(5q) patients that have assessed risk factors for dis-

ease progression to acute myeloid leukemia (AML) or to higher risk forms of

MDS have been small, so there is uncertainty about which disease-associ-

ated features confer a high-risk of progression.

Ulrich Germing of Heinrich-Heine University in Dusseldorf led a multina-

tional consortium that assessed outcomes and prognostic variables in 303

patients with MDS-associated with del(5q), with a focus on disease progres-

sion. Patients included in this series received supportive care alone; the 62

patients who received lenalidomide were censored at the time lenalidomide

treatment started.

In this series, there were several factors associated with better outcomes.

Patients with del(5q) as a sole chromosomal aberration did better than those

with more than one aberration in addition to del(5q): a median survival of 73

months versus 19.3 months for ‘‘del(5q) plus.’’ The most important prognos-

tic factor was the need for red cell transfusions at diagnosis: patients requir-

ing transfusions at diagnosis survived only 39 months, compared to 97

months in transfusion-independent patients. Both differences were highly

statistically significant. The IPSS and WHO-based Prognostic Scoring Sys-

tem (WPSS) stratified patients well.

The cumulative AML progression rate in this cohort, calculated with the

Kaplan-Meier method, was 7% at 2 years and 18.2% at 5 years. Factors

associated with a high risk of AML transformation included intermediate-1

IPSS risk group (instead of low risk) and high-risk WPSS score (instead of

very low, low, or intermediate risk), marrow blast proportion >5%, and trans-

fusion need at diagnosis.

Several other investigative groups have also recently assessed prognostic

factors in del(5q) MDS. A Japanese study of 183 patients published in 2008

also reported that the WPSS has prognostic utility in this group, and the

Japanese investigators found that addition of either monosomy 7 or a com-

plex karyotype to del(5q) correlates with shorter overall survival [13]. A

Mayo Clinic study of 130 del(5q) patients, also from the prelenalidomide era,

reported that the least favorable outcomes are associated with complex

cytogenetics, lack of any normal metaphases, normocytic rather than macro-

cytic erythrocyte indices, and a low baseline absolute lymphocyte count (i.e.,

<1.2 3 109/L) [14]. Finally, a European multicenter analysis of del(5q) MDS

that included 60 patients with carefully defined 5q- syndrome (a small subset

of del(5q) MDS [14]) found that this entity has a relatively benign prognosis,

with a median overall survival of 107 months and a low probability (<5%) of

AML progression [12]. In this European series, an increase in the marrow

blast proportion to �5% or the addition of another karyotypic abnormality in

addition to del(5q) were each associated with dramatic reduction in overall

survival. It is likely that lenalidomide modifies the natural history of del(5q)

MDS, and not all of these variables may retain independent prognostic sig-

nificance in the lenalidomide era [15].

Abstract #3814: Independent validation of the MD Anderson

Cancer Center Risk Model for MDS, and comparison to the

International Prognostic Scoring System (IPSS) and the World

Health Organization-Based Prognostic Scoring System (WPSS)Although the 1997 IPSS remains the gold standard for MDS prognostica-

tion and continues to be used widely, a number of criticisms have been

leveled at the IPSS (summarized in my 2009 annual meeting MDS Educa-

tion Session manuscript [7]). As a consequence, there have been several

recent proposals for new prognostic tools incorporating disease features that

were not included in the IPSS. Prominent among these new systems are

the WPSS—initially published in 2007 by Luca Malcovati from Pavia, Italy

and his colleagues and expanded in 2009 by Matteo Della Porta and his

colleagues to include marrow fibrosis as a disease feature with adverse

prognosis—and a generalized MDS/chronic myelomonocytic leukemia

(CMML) risk model published in 2008 by Kantarjian and coworkers at

MDACC [16–18].

To compare the IPSS directly with the WPSS and the MDACC risk model

in an independent group of patients, my colleagues at Mayo Clinic and I

evaluated the medical records of 1,503 adult patients with MDS (n 5 1,249)

or CMML (n 5 254) evaluated at Mayo Clinic between January 1996 and

December 2007. Follow-up was complete until death in 1,122 patients

(74%). We found that all three systems—the IPSS, WPSS, and MDACC risk

model—stratify patients accurately, but the MDACC risk model best identifies

the lowest-risk patients and also classifies the broadest group of patients

(i.e., primary and secondary MDS and primary and secondary CMML with

or without leukocytosis). In a multivariable proportional hazards model, each

of the components of the MDACC risk model (Table III) except a white cell

count of >20 3 109/L retained independent prognostic significance in this

series. Therefore, this abstract represents an independent validation of the

MDACC risk model.

The IPSS is currently being revised by a working group of the MDS Foun-

dation (Available at www.mds-foundation.org), led by Peter Greenberg from

Stanford University, using a multinational dataset of >5,000 patients. The

Mayo Clinic results presented at the 2009 ASH annual meeting suggest that

‘‘version 2.0’’ of the IPSS should include the patient factors accounted for by

the MDACC risk model, with the possible exception of leukocytosis. The

TABLE I. International MDS Cytogenetic Risk Model Derived

from 3,856 Patients

Risk group Karyotypes

Median survival

(months)

Good Normal karyotype, 21/1p2,

der(1;7)/t(1;7), del(5q),

t(5q), del(11q),

del(12p), 213/13q2,

del(16q), 119, del(20q),

2Y, any 2 abnormalities

including del(5q)

50.6

Intermediate-1 18, 29/9q, del(17p), iso(17q),

121, 221, 2X, 1mar,

any other single abnormality,

any 2 abnormalities

not including

abnormalities of

chromosomes 5q or

27/7q2

25.7

Intermediate-2 11/11q/dup(1q),

der(3)(q21/q26),

27/7q2, t(11q23), 111,

any 2 abnormalities including

27 or 7q2, complex with 3

abnormalities

16.0

Poor Complex with >3 abnormalities 5.7

Source: ASH 2009 Abstract #2772.

TABLE II. WHO 2008 Classification Revision—Karyotypes Suitable to

Diagnose ‘‘MDS-Unclassifiable,’’ in the Absence of Diagnostic Morphology

Abnormality

WHO-estimated

frequency

Gene(s) involved

(HGNC)

Unbalanced rearrangements

27 or del(7q) 10%; 50% in t-MDS Unknown

25 or del(5q) 10%; 40% in t-MDS Unknown; multiple

candidates

i(17q) or t(17p) 3–5% TP53

213 or del(13q) 3% Unknown

del(11q) 3% Unknown

del(12p) or t(12p) 3% ETV6 in some cases

del(9q) 1–2% Unknown

idic(X)(q13) 1–2% Unknown

Balanced rearrangements

t(11;16)(q23;p13.3) 3% in t-MDS MLL, CREBBP

t(3;21)(q26.2;q22.1) 2% in t-MDS RUNX1, MDS1-EVI1

t(1;3)(p36.3;q21.2) <1% PRDM16, MDS1-EVI1

t(2;11)(p21;q23) <1% miR-125b-1

Inv(3)(q21q26.2) <1% MDS1-EVI1,

RPN1 regulatory

elements

t(6;9)(p23;q34) <1% DEK-NUP214

Source: Refs. 6, 7.Excluded as disease-defining sole abnormalities: loss of the Ychromosome; trisomy 8; del(20q).WHO, World Health Organization; HGNC,Human Genome Nomenclature Committee.

letters

American Journal of Hematology 197

Page 9: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

new IPSS will likely include the expanded karyotype risk model described

above (abstract #2772).

Section 2: Hypomethylating Agents (Azacitidine and Decitabine)Abstract #3797: A study comparing dosing regimens and

efficacy of subcutaneous to intravenous azacitidine (AZA)

for the treatment of MDS

Because azacitidine is the only hypomethylating agent that has been

shown in a randomized trial (AZA-001) to be superior to supportive care

alone in patients with higher risk forms of MDS, this drug is now widely used

in clinical practice to treat patients with higher risk MDS [19]. Initial United

States regulatory approval of azacitidine by the Food and Drug Administra-

tion (FDA) in May 2004 was for a dose and schedule of 75 mg/m2/day,

administered subcutaneously for 7 consecutive days, based on the treatment

regimen used in the CALGB 9221 trial, which was published in 2002 and

formed the basis for FDA approval [20]. However, because for most patients

a dose of 75 mg/m2 requires multiple subcutaneous injections of azacitidine

and skin reactions (pain, erythema, etc.) are common (in the AZA-001 trial,

43% of patients reported injection site redness, and 29% of patients com-

plained of other injection site reactions), clinicians have expressed interest

in intravenous azacitidine [21].

A small bioequivalence study published in 2005 showed that area-under-

the-curve bioavailability for subcutaneous azacitidine is on an average 89%

of intravenous administration [22]. In a small Phase II study of intravenous

azacitidine that enrolled 22 evaluable patients, 27% of patients responded

[five complete responses (CR) and one partial response] [23]. Although the

FDA approved a supplement to the package insert for azacitidine describing

intravenous administration in January 2007, the widely cited AZA-001 sur-

vival study published in early 2009 used subcutaneous azacitidine, not intra-

venous.

Mikkael Sekeres from Cleveland Clinic and his colleagues assessed 331

azacitidine-treated patients with MDS or oligoblastic AML (i.e., <30% mar-

row blasts), who were included in the industry-sponsored AVIDA registry.

Among these patients, 190 (57%) received intravenous azacitidine and 141

patients (43%) received subcutaneous azacitidine, in each case for a

median of four cycles. Notably, only 17% of patients in the registry received

the inconvenient FDA-approved continuous 7-day dosing schedule: 51%

received <7 days of therapy per cycle, 30% received 7 total days of therapy

but with breaks, and 2% received >7 days of therapy. There were no differ-

ences between subcutaneous and intravenous dosing in terms of hemato-

logical response rate (24% for each route of administration). Patients under-

going intravenous administration received a slightly lower dose of azacitidine

overall (mean, 12 mg less), and there was also a trend toward less-frequent

dosing cycles (i.e., >28 days), with intravenous rather than with subcutane-

ous dosing.

While hematological response rates may be similar for subcutaneous and

intravenous azacitidine, and intravenous administration is one way of avoid-

ing skin reactions, most MDS experts believe that the subcutaneous route

should still be considered the standard of care, because of the AZA-001

study results.

Abstract #2773: Different clinical results with the use

of different dosing schedules of azacitidine in

patients with MDS managed in community-based

practice: Effectiveness and safety data from

the Spanish Azacitidine Compassionate

Use Registry

The study presented in this abstract also examined practical azacitidine

administration issues, as discussed earlier.

A Spanish investigative group reviewed data from 144 patients with WHO-

defined MDS, who received azacitidine and were enrolled in clinical trials or

compassionate using protocols in Spain, prior to azacitidine’s formal market-

ing approval from the Spanish Medicines Agency in May 2009. Patients

were separated into three cohorts: cohort A had received only 5 days of

azacitidine per cycle; cohort B had received 7 days of azacitidine per cycle,

but skipped weekend days (days 6 and 7, with makeup doses administered

on days 8 and 9); and cohort C had received the standard dose and sched-

ule of 75 mg/m2/day, administered subcutaneously for 7 consecutive days,

as in the CALGB 9221 and AZA-001 studies. Even though the baseline per-

formance status of patients in cohort C was poorer (other baseline charac-

teristics were balanced), the overall response rate to azacitidine was highest

in cohort C: 74% (using International Working Group (IWG) 2006 criteria),

compared to 65% in cohort B and only 58% in cohort A.

A study by Lyons et al. in the US Oncology community practice network

was published in 2009; this study compared overall hematological response

rates in MDS using three weekend-sparing azacitidine schedules [24]. Lyons

and colleagues observed similar response rates (44–56%) between the

three different regimens. However, the study did not include a control arm

receiving the standard 7 consecutive day regimen, so, it is possible that all

three weekend-sparing regimens would have been inferior to the standard

regimen.

The Spanish results presented at the ASH annual meeting suggest that

whenever possible, clinicians using azacitidine should try to stick with the 7

consecutive day schedule described in the package insert. However, in

some communities this is simply not feasible—there is no infusion center

open on Saturday or Sunday. In these cases, patients and clinicians need to

accept the possibility of a lower response rate.

Abstract #117: A Phase 1, Open-label, dose-escalation study

to evaluate the safety, pharmacokinetics, and pharmacody-

namics of oral azacitidine in patients with MDS or AML

One solution to the challenges associated with weekend administration of

parenteral azacitidine is to switch to an oral preparation that patients can

self-administer at home. Oral agents offer patients other advantages such

as convenience, and an oral formulation of azacitidine might allow extended-

dosing schedules to be explored more easily. However, in the United States,

reimbursement for oral chemotherapeutic agents is not equivalent to that for

parenteral agents, so oral agents are often costlier for patients.

Guillermo Garcia-Manero from MDACC presented pilot data from a classic

dose-escalating ‘‘3 1 3’’ cohort design Phase I trial of an oral formulation of

azacitidine, in patients with MDS or AML. Patients received a single cycle of

7-day subcutaneous azacitidine as a benchmark for pharmacokinetics, and

then subsequent cycles were administered orally.

The initial oral azacitidine dose used was 120 mg/day for 7 consecutive

days, repeated every 28 days. The dose-limiting toxicity was reached at 600

mg/day 3 7 days (Grade 3 or 4 diarrhea developed in two of three subjects

enrolled at that dose level), so the maximally tolerated dose was 480 mg/

day 3 7 days, administered once in every 28 days. Unfortunately, there was

wide interpatient variability in the pharmacokinetics of oral azacitidine, with

area-under-the-curve bioavailability ranging from 15 to 167% compared with

TABLE III. M.D. Anderson Cancer Center Risk Model for MDS/CMML

Prognostic factor

Points

(0–17)

Performance status �2 2

Age: 60–64 1

�65 2

Platelets: <30 3 109/L 3

30–49 3 109/L 2

50–199 3 109/L 1

Hemoglobin <12 g/dL 2

Bone marrow blasts: 5–10% 1

11–29% 2

White blood count �20 3 109/L 2

Karyotype: chromosome

7 abnormality or

complex (�3 abnormalities)

3

Prior transfusion 1

Risk

group

Total score

(see above)

% Patients Median

survival

% Alive

(3 years)

% Alive

(6 years)

Low 0–4 16 54 63 38

Int-1 5–6 24 25 34 13

Int-2 7–8 24 14 16 6

High 9–17 36 6 4 0.4

Source: Ref. 18. Sum all points from the above eight disease-associated factors.Total possible points: 17.

letters

198 American Journal of Hematology

Page 10: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

subcutaneous exposure. Among the first 14 evaluable patients, four (29%)

experienced a CR, and six (43%) had stable disease, using 2006 IWG MDS

response criteria or 2003 IWG AML response criteria.

It is difficult to know what role oral azacitidine might eventually play in

MDS therapy, given the wide interpatient differences observed in drug

absorption and metabolism. If, for example, clinicians will need to follow aza-

citidine levels for patients on oral therapy and titrate dose accordingly, this

would markedly decrease the convenience associated with an oral agent.

Abstract #119: A Phase II randomized Bayesian study of very

low dose subcutaneous decitabine administered daily

or weekly times three in patients with lower risk MDS

Decitabine is the other hypomethylating agent available in the United

States; it was approved by the FDA for treatment of MDS in May 2006.

Although neither the D-0007 decitabine registration study [25] nor the

EORTC decitabine survival study presented at the 2008 ASH annual meet-

ing showed a statistically significant overall survival benefit for decitabine

compared with supportive care, the EORTC study in particular has been

criticized, because it used an older 3-day inpatient decitabine regimen, and

patients also received short duration of therapy (median <4 cycles with 40%

of patients receiving 2 cycles or less, compared to a median of 9 cycles of

azacitidine in the AZA-001 study). Nevertheless, the fact that there is a posi-

tive survival study for azacitidine and no such study for decitabine led the

National Comprehensive Cancer Network (NCCN) in 2009 to recommend

azacitidine, as the preferred agent for treatment of patients with higher-risk

forms of MDS (Available at www.nccn.org).

There is a disconnect between how leukemia cell lines best respond to

hypomethylating agents in vitro (i.e., with prolonged low-dose exposure) and

how these agents are dosed in the clinic (typically monthly boluses of

between 3 and 7 consecutive days). Several clinical groups are currently

exploring alternative schedules of decitabine or azacitidine. Lower and less

frequent doses are especially attractive in patients with lower risk MDS, for

whom the risk-benefit ratio of hypomethylating agents used at standard

doses may be less favorable.

On behalf of the MDACC, Guillermo Garcia-Manero presented a Phase II

study with an adaptive randomization exploring decitabine in lower-risk MDS

patients (IPSS low or intermediate-1), 84% of whom had de novo disease, a

median of 2.5 months from diagnosis. Decitabine was administered subcuta-

neously either once daily for 3 consecutive days (20 mg/m2/day) or once

weekly for 3 weeks using the same dose, with treatment cycles repeated

every 4 weeks. Among 43 enrolled patients, the overall CR rate (IWG 2006

criteria) was 9% (i.e., three patients with CR in the 3-day arm, one in the

weekly arm); there were also two marrow responses (5%), one partial

response (2%), and four hematological improvements (9%). The overall

response rate was 25%: 32% in the 3-day arm and 19% in the weekly arm,

a difference that was not significant. No patients died of treatment-related

complications, but two patients died of AML progression; most of the rest of

the patients (81%) remain on study, which the investigators felt was indica-

tive of a well-tolerated regimen.

The response rates seen with these three-dose-per-month regimens

approximate those observed in the D-0007 registration trial, but are lower

than those reported with a 5-day decitabine regimen either in a single-center

MDACC study or in the multicenter DACO-020 (ADOPT) trial [11,25,26].

However, the rate of febrile neutropenia was higher on the DACO-020 trial—

17%—than in this study, and several early septic deaths occurred in patients

enrolled on DACO-020 compared to none in this study [11]. Although addi-

tional data are needed, the three-dose-per-month study regimens reported by

Garcia-Manero and colleagues are active and seem unlikely to be harmful to

patients, and might be appropriate for patients with lower-risk MDS who have

an indication for treatment, such as transfusion dependence.

Section 3: Immunomodulatory and Immunosuppressive

Agents (Lenalidomide and Alemtuzumab)Abstract #944: RBC transfusion independence and safety

profile of lenalidomide 5 or 10 mg in PTS with low- or

Int-1-Risk MDS with del5q: Results from a randomized

Phase III trial (MDS-004)

Most therapeutic studies presented at the 2009 ASH annual meeting were

single-arm pilot studies. An exception was the MDS-004 study: a random-

ized, multicenter, prospective, placebo-controlled study of two different

schedules of lenalidomide in lower-risk MDS patients with del(5q), presented

by Pierre Fenaux from Paris.

Three previous MDS lenalidomide monotherapy studies have been pub-

lished. MDS-001, the 43 patient pilot study that first identified patients with

del(5q) MDS, as the subset most likely to respond to lenalidomide, was

published in the New England Journal of Medicine by List et al. in 2005

[27]. In MDS-001, patients received 10 mg of lenalidomide for 21 days of

every 28 day cycle. A follow-up study, MDS-003, limited enrollment to

transfusion-requiring IPSS low or intermediate-1 risk patients with del(5q);

results were published in the New England Journal of Medicine in 2006

[15]. In MDS-003, 67% of 148 enrolled patients became transfusion inde-

pendent, the median hemoglobin increment was 5.4 g/dL, 45% of evaluable

patients experienced a complete cytogenetic remission, and the median

time to response was just 4.6 weeks. MDS-003 included patients who

received 10 mg daily dosing, as well as a small cohort who received 10

mg for 21 days of 28 days; overall response rates were comparable

between these groups. Finally, MDS-002 was a study of 214 red-cell-trans-

fusion-requiring patients with lower-risk MDS without del(5q), published in

Blood in 2008 by Raza et al. [28]. The response rate to lenalidomide was

lower in MDS-002 (26% transfusion independence) than in MDS-003, and

responses were also less durable (median 41 weeks, compared to 2.2

years in the del(5q) subset).

Although the recommended initial dose of lenalidomide for MDS is 10 mg

each day (the FDA-approved schedule), more than half of patients who start

taking this dose eventually require a dose reduction, because of severe neu-

tropenia or thrombocytopenia [29]. Therefore, the MDS-004 trial was under-

taken to see if a starting dose of 5 mg/day is just as efficacious as 10 mg/

day. In addition, a placebo control was included in MDS-004, in part to sat-

isfy European regulators who denied an application for approval of lenalido-

mide for MDS in 2008, because of a lack of randomized trial data and con-

cern about possible acceleration of disease progression [30,31].

MDS-004 enrolled 205 patients with IPSS low/intermediate-1 MDS, red

cell transfusion requirement, and del(5q). del(5q) was present as an isolated

karyotypic abnormality in 75% of patients, and patients were enrolled a

median of 2.5 years from the time of their initial MDS diagnosis. Patients

were randomized to receive lenalidomide 5 mg each day (i.e., days 1–28 of

each 28 day cycle), 10 mg per day for 21 of 28 days, or an oral placebo.

The duration of the trial was 52 weeks, and the primary endpoint was red

cell transfusion independence lasting at least 26 weeks.

The evaluable intention-to-treat population reported by Fenaux at the

2009 annual meeting included 138 patients: 46 who received 5 mg of lenali-

domide, 41 who received 10 mg lenalidomide, and 51 who got a placebo.

(The majority of the 67 inevaluable patients were excluded, because of inad-

equate bone marrow examinations.) The best results were seen with the 10

mg lenalidomide dosing: 56% of patients achieved the primary endpoint of

transfusion independence, compared to 41% with 5 mg of lenalidomide and

6% with placebo. Cytogenetic responses were also more common with the

10 mg dose (41% response rate, including 24% cytogenetic CRs, versus

17% response rate with 11% cytogenetic CRs for 5 mg, and no responses

for placebo).

From a safety standpoint, with the whole 205 patient cohort considered,

the highest rate of AML progression was in the 5 mg lenalidomide arm (6%,

compared to 1% with 10 mg and 2% with placebo). A similar proportion of

patients required dose reduction in each active treatment arm—58% for 10

mg versus 52% for 5 mg and 0% for placebo—and Grade 3 or 4 neutrope-

nia was reported in 75% of patients on the 10 mg arm, 74% on the 5 mg

arm, and 15% with placebo.

The starting dose of lenalidomide in MDS should remain 10 mg, as this

appears to be the most efficacious dose and is no more likely to require

dose reduction than a starting dose of 5 mg/day. However, the use of 21 of

28-day dosing in the MDS-004 study may create some confusion, as the

current package insert for lenalidomide suggests 10 mg daily dosing without

an interruption, based on the MDS-003 trial results. It seems likely that prob-

lems with patient adherence to the regimen would increase with a 21/28 day

regimen. If 21/28 day dosing becomes commonplace, perhaps blister packs

including 7 days of placeholder placebos could be developed, such as those

currently used to dispense oral contraceptives, so that patients do not ‘‘lose

track’’ of when the next dose of drug is due.

letters

American Journal of Hematology 199

Page 11: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

Abstract #115: Lenalidomide in high-risk MDS and AML

with chromosome 5 abnormalities

Patients enrolled in the MDS-001 and MDS-003 trials described above had

lower-risk disease and were required to have both a platelet count >50 3 109/L

and an absolute neutrophil count >0.53 109/L. There are anecdotes of meaning-

ful clinical responses to lenalidomide in patients who would not have met enroll-

ment criteria for MDS-001/003, including higher-risk patients such as those with

del(5q) MDS with excess blasts, or even AML [32–34]. del(5q) may not be neces-

sary for a response to lenalidomide in higher risk disease; for instance, one report

described two patients with trisomy 13 AML without del(5q), who experienced CR

with lenalidomide monotherapy [35].

A 47-patient French study of lenalidomide in IPSS intermediate-2 or high-

risk MDS/oligoblastic AML associated with del(5q) was reported in early

2009 [36]. The dose employed was 10 mg/day for 21 of every 28 days, with

dose escalation to 15 mg permitted for nonresponse. Among nine patients

with isolated del(5q) in the French study, six (67%) achieved CR using IWG

2006 criteria, but only 1 of 11 (9%) and 0 of 27 patients with one or more

than one additional cytogenetic abnormality, respectively, had a CR. The

pretreatment platelet count also predicted response: 7/20 (35%) patients

with platelets >100 3 109/L had a CR, compared with 0 of 27 with a lower

platelet count.

On behalf of the Nordic MDS Group, Eva Hellstrom-Lindberg reported early

results from an investigator-initiated Phase II trial of lenalidomide in patients

with higher-risk MDS and AML who had either del(5q) or monosomy 5 and

who were not considered candidates for induction chemotherapy. The initial

dose of lenalidomide was 10 mg/day, with dose escalations up to 30 mg permit-

ted for patients who were tolerating the drug acceptably but not yet responding.

Among 26 enrolled patients (12 MDS, 14 AML), only nine (35%) were able to

complete 16 weeks of lenalidomide therapy. Of 13 patients who stopped treat-

ment before 12 weeks, 10 did so because of adverse effects (mostly infection

or febrile neutropenia—six patients died), and three had progressive disease.

Major cytogenetic responses were seen in 4/26 (15%) of enrolled patients,

three of whom also experienced a hematological response and two of whom

had a bone marrow blast reduction, and 2/26 (7.5%) of the other patients expe-

rienced a reduction in marrow blasts without a karyotypic response.

Two abstracts (#841 and #842) presented at the 2009 annual meeting

reported initial results from Phase I/II studies of lenalidomide in patients with

AML >60 years old without del(5q) (30% CR rate), or relapse of AML after

allogeneic stem cell transplantation (16% CR rate). Reimbursement may be

an issue with lenalidomide in this setting; also, the rate of adverse events is

high, and overall response rates are low. However, given the paucity of other

options available to patients with del(5q) higher risk MDS/AML who do not

respond to existing therapies such as hypomethylating agents and who are

too frail to consider chemotherapy, a brief trial of lenalidomide can be con-

sidered, especially for those who are not candidates for clinical trials.

Abstract #116: Alemtuzumab treatment of intermediate-1

(INT-1) myelodysplasia patients is associated with

sustained improvement in blood counts and

cytogenetic remissions

The Hematology group at the National Institutes of Health (NIH) in

Bethesda, MD, have been exploring immunosuppressive therapies in MDS

for more than a decade, based on a variety of studies suggesting a role for

dysregulated T lymphocytes in the marrow failure associated with MDS, as

well as evidence of pathobiological and clinical overlap between MDS and

aplastic anemia [37]. Early studies from the NIH and other groups reported

clinical benefit for some patients with MDS treated with antithymocyte globu-

lin, cyclosporine A, or both [38,39].

However, these studies have been difficult to interpret or to apply broadly,

as enrolled patients have often differed from the more typical MDS patients

seen in clinical practice. The patients enrolled on studies of immunosuppres-

sive drugs tend to be younger, are more likely to have hypocellular or nor-

mocellular marrow, and are more likely to be HLA DRB15 positive (an HLA

type over-represented in patients with aplastic anemia) than the general

MDS population. When ATG has been used in a more typical MDS patient

group, including patients with excess blasts or a complex karyotype, it has

proven less effective [40].

In 2003, Saunthararajah et al. published a model that helps predict which

patients are most likely to respond to immunosuppressive therapy (Table IV);

patients who are young, HLA DRB15 positive, and have a short duration of red

cell transfusions are most likely to respond [41]. A European group subse-

quently observed that patients with a hypocellular marrow or a lower risk IPSS

score aremore likely to respond to immunosuppressive therapy [42].

At the ASH 2009 meeting, Matthew Olnes from the NIH presented data

from a pilot study of alemtuzumab (anti-CD52 monoclonal antibody) in MDS.

The dose and schedule of alemtuzumab used in the NIH study, 10 mg/day

intravenously for 10 consecutive days, are different from the regimen

approved for chronic lymphocytic leukemia. Enrolled patients were those

with a high positive predictive score for response to immunosuppressive

therapy, according to the Saunthararajah model discussed earlier.

As in previous NIH studies, patients enrolled in this trial were younger than

the median age for MDS by >15 years (median age <55 years, compared to a

median age of 76 years at the time of diagnosis for patients with MDS captured

by the National Cancer Institute Survey, Epidemiology, and End Results (SEER)

database [43]) and were more likely to have a hypocellular marrow (>40% of

patients, compared to 10–15% of patients in general MDS series [44]). Alemtu-

zumab was chosen because, compared to ATG, it is more profoundly immuno-

suppressive, and the immunosuppression with alemtuzumab may also be more

durable. In addition, unlike ATG, alemtuzumab is not associated with serum

sickness, and unlike cyclosporine, alemtuzumab is not nephrotoxic.

Among this carefully selected group of patients, 15 of 16 (93%) patients

with IPSS intermediate-1 risk disease and two of five (40%) patients with

IPSS intermediate-2 risk disease experienced hematological improvement

with alemtuzumab, whereas five of seven (71%) of patients with a normal

karyotype at baseline experience cytogenetic remission.

Patients on this study received trimethoprim–sulfamethoxazole to prevent

Pneumocystis lung infections and antiviral prophylaxis to prevent herpetic

and cytomegalovirus reactivation. Patients were monitored weekly with poly-

merase chain reaction (PCR) blood tests for Epstein-Barr virus (EBV) or

CMV reactivation; four of 22 patients became transiently positive for EBV,

but none developed disease, and there were no clinically significant infec-

tions during the study.

Clinicians might consider immunosuppressive therapy for the rare patient

with a high positive predictive score for response to immunosuppressive

therapy, according to the Saunthararajah model. However, it is unclear, if

alemtuzumab is superior to ATG or cyclosporine in this group, or whether

insurance companies will pay for alemtuzumab in MDS. If alemtuzumab is

chosen, then anti-infective prophylaxis and monitoring for reactivation of

viruses is essential.

Conclusion

With the exception of the oral azacitidine abstract (a new formulation of

drug already FDA-approved for MDS), all of the above abstracts describe

findings that can inform clinicians’ practice immediately. Notably, however,

MDS diagnostic and prognostic tools such as those described above will

continue to be of limited without a better understanding of the molecular

pathology of MDS, and new treatments are needed for the many patients

with MDS who do not respond to any of the therapies listed earlier.

Other MDS abstracts not mentioned here that might have future clinical

relevance included early-phase clinical trials of ON 01910.Na, a Polo like-

kinase 1 inhibitor (abstracts #120 and #3815); sapacitabine, a novel nucleo-

side analog with a unique ability to cause irreparable single-strand deoxyribo

nucleic acid (DNA) breaks and induce G2 cell cycle arrest (abstract #1758);

an oral formulation of clofarabine, a halogenated adenosine analog for which

an intravenous formulation is currently FDA-approved for treatment of refrac-

TABLE IV. Model for Predicting Response to Immunotherapy in

MDS

If HLA DRB15 2: If HLA DRB15 1: PPR

X > 57 X > 71 Low (0–40%)

X � 57 X � 71 High (41–100%)

PPR Total Response No response

Low (0–40%) 14 (61%) 1 13

High (41–100%) 9 (39%) 6 3

Source: Ref. 41.X 5 the age of the patient (in years), plus the duration of red celltransfusion dependence of the patient (in months).Validation cohort: n 5 23patients.PPR, predicted probability of response.

letters

200 American Journal of Hematology

Page 12: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

tory acute lymphoblastic leukemia (abstract #118); and a variety of histone

deacetylase inhibitors or combination regimens.

Division of Hematological Malignancies, Dana-Farber Cancer InstituteHarvard Medical School, Boston, Massachusetts

Correspondence to: David P. Steensma, Division of Hematological MalignanciesDana-Farber Cancer Institute, Harvard Medical School, 44 Binney Street

Suite D1B30 (Mayer 1B21), Boston, MA 02115*E-mail: [email protected]

Conflict of interest: Nothing to reportPublished online 31December 2009 inWiley InterScience (www.interscience.wiley.com).

DOI: 10.1002/ajh.21629

References

1. Olney HJ, Le Beau MM. The cytogenetics of myelodysplastic syndromes. BestPract Res Clin Haematol 2001;14:479–495.

2. Haase D, Germing U, Schanz J, et al. New insights into the prognostic impact ofthe karyotype in MDS and correlation with subtypes: Evidence from a core data-set of 2,124 patients. Blood 2007;110:4385–4395.

3. Haase D. Cytogenetic features in myelodysplastic syndromes. Ann Hematol2008;87:515–526.

4. Greenberg P, Cox C, LeBeau MM, et al. International scoring system for evaluat-ing prognosis in myelodysplastic syndromes. Blood 1997;89:2079–2088.

5. Haase D, Estey EH, Steidl C, et al. Multivariate evaluation of the prognostic andtherapeutic relevance of cytogenetics in a merged European-American cohortof 3,860 patients with MDS. Blood 2007;110; Abstract #247.

6. Swerdlow SH, Campo E, Harris NL, et al, editors. WHO Classification of Tumoursof Haematopoietic and Lymphoid Tissues, 4th ed. Lyon: IARC Press; 2008.

7. Steensma DP. The changing classification of myelodysplastic syndromes:What’s in a name? Hematology 2009;2009:645–655.

8. Vardiman JW, Thiele J, Arber DA, et al. The 2008 revision of the World HealthOrganization (WHO) classification of myeloid neoplasms and acute leukemia:Rationale and important changes. Blood 2009;114:937–951.

9. Steensma DP, Dewald GW, Hodnefield JM, et al. Clonal cytogenetic abnormal-ities in bone marrow specimens without clear morphologic evidence of dyspla-sia: A form fruste of myelodysplasia? Leuk Res 2003;27:235–242.

10. Loss of the Y chromosome from normal and neoplastic bone marrows. UnitedKingdom cancer cytogenetics group (UKCCG). Genes Chromosomes Cancer1992;5:83–88.

11. Steensma DP, Baer MR, Slack JL, et al. Multicenter study of decitabine adminis-tered daily for 5 days every 4 weeks to adults with myelodysplastic syndromes:The alternative dosing for outpatient treatment (ADOPT) trial. J Clin Oncol2009;27:3842–3848.

12. Giagounidis AA, Germing U, Wainscoat JS, et al. The 5q- syndrome. Hematol-ogy 2004;9:271–277.

13. Tasaka T, Tohyama K, Kishimoto M, et al. Myelodysplastic syndrome with chro-mosome 5 abnormalities: A nationwide survey in Japan. Leukemia 2008;22:1874–1881.

14. Holtan SG, Santana-Davila R, Dewald GW, et al. Myelodysplastic syndromesassociated with interstitial deletion of chromosome 5q: Clinicopathologic correla-tions and new insights from the prelenalidomide era. Am J Hematol 2008;83:708–713.

15. List A, Dewald G, Bennett J, et al. Lenalidomide in the myelodysplastic syn-drome with chromosome 5q deletion. N Engl J Med 2006;355:1456–1465.

16. Malcovati L, Germing U, Kuendgen A, et al. Time-dependent prognostic scoringsystem for predicting survival and leukemic evolution in myelodysplastic syn-dromes. J Clin Oncol 2007;25:3503–3510.

17. Della Porta MG, Malcovati L, Boveri E, et al. Clinical relevance of bone marrowfibrosis and CD34-positive cell clusters in primary myelodysplastic syndromes.J Clin Oncol 2009;27:754–762.

18. Kantarjian H, O’Brien S, Ravandi F, et al. Proposal for a new risk model in myelo-dysplastic syndrome that accounts for events not considered in the original inter-national prognostic scoring system. Cancer 2008;113:1351–1361.

19. Fenaux P, Mufti GJ, Hellstrom-Lindberg E, et al. Efficacy of azacitidine com-pared with that of conventional care regimens in the treatment of higher-riskmyelodysplastic syndromes: A randomised, open-label, phase III study. LancetOncol 2009;10:223–232.

20. Silverman LR, Demakos EP, Peterson BL, et al. Randomized controlled trial ofazacitidine in patients with the myelodysplastic syndrome: A study of the cancerand leukemia group B. J Clin Oncol 2002;20:2429–2440.

21. Gore SD. Intravenous azacitidine for MDS. Clin Adv Hematol Oncol2007;5:234.

22. Marcucci G, Silverman L, Eller M, et al. Bioavailability of azacitidine subcutane-ous versus intravenous in patients with the myelodysplastic syndromes. J ClinPharmacol 2005;45:597–602.

23. Martin MG, Walgren RA, Procknow E, et al. A phase II study of 5-day intrave-nous azacitidine in patients with myelodysplastic syndromes. Am J Hematol2009;84:560–564.

24. Lyons RM, Cosgriff TM, Modi SS, et al. Hematologic response to three alterna-tive dosing schedules of azacitidine in patients with myelodysplastic syndromes.J Clin Oncol 2009;27:1850–1856.

25. Kantarjian H, Issa JP, Rosenfeld CS, et al. Decitabine improves patient out-comes in myelodysplastic syndromes: Results of a phase III randomized study.Cancer 2006;106:1794–1803.

26. Kantarjian H, Oki Y, Garcia-Manero G, et al. Results of a randomized study of 3schedules of low-dose decitabine in higher-risk myelodysplastic syndrome andchronic myelomonocytic leukemia. Blood 2007;109:52–57.

27. List A, Kurtin S, Roe DJ, et al. Efficacy of lenalidomide in myelodysplastic syn-dromes. N Engl J Med 2005;352:549–557.

28. Raza A, Reeves JA, Feldman EJ, et al. Phase II study of lenalidomide in transfu-sion-dependent, low-risk, and intermediate-1 risk myelodysplastic syndromeswith karyotypes other than deletion 5q. Blood 2008;111:86–93.

29. Sekeres MA, Maciejewski JP, Giagounidis AA, et al. Relationship of treatment-related cytopenias and response to lenalidomide in patients with lower-risk mye-lodysplastic syndromes. J Clin Oncol 2008;26:5943–5949.

30. Cazzola M. Myelodysplastic syndrome with isolated 5q deletion (5q- syndrome).A clonal stem cell disorder characterized by defective ribosome biogenesis.Haematologica 2008;93:967–972.

31. Jadersten M, Saft L, Pellagatti A, et al. Clonal heterogeneity in the 5q-syndrome: p53 expressing progenitors prevail during lenalidomide treat-ment and expand at disease progression. Haematologica 2009;94:1762–1766.

32. Mesa RA, Tefferi A, Li CY, Steensma DP. Hematologic and cytogeneticresponse to lenalidomide monotherapy in acute myeloid leukemia arising fromJAK2(V617F) positive, del(5)(q13q33) myelodysplastic syndrome. Leukemia2006;20:2063–2064.

33. Lancet JE, List AF, Moscinski LC. Treatment of deletion 5q acute myeloid leuke-mia with lenalidomide. Leukemia 2007;21:586–588.

34. Borthakur G, Garcia-Manero G, Faderl S, et al. Lenalidomide in high-risk myelo-dysplastic syndrome and acute myelogenous leukemia associated with chromo-some 5 abnormalities. ASH Annu Meet Abstr 2007;110:1459.

35. Fehniger TA, Byrd JC, Marcucci G, et al. Single-agent lenalidomide inducescomplete remission of acute myeloid leukemia in patients with isolated trisomy13. Blood 2009;113:1002–1005.

36. Ades L, Boehrer S, Prebet T, et al. Efficacy and safety of lenalidomide in inter-mediate-2 or high-risk myelodysplastic syndromes with 5q deletion: Results of aphase 2 study. Blood 2009;113:3947–3952.

37. Sloand EM, Rezvani K. The role of the immune system in myelodysplasia: Impli-cations for therapy. Semin Hematol 2008;45:39–48.

38. Molldrem JJ, Leifer E, Bahceci E, et al. Antithymocyte globulin for treatment ofthe bone marrow failure associated with myelodysplastic syndromes. Ann InternMed 2002;137:156–163.

39. Stadler M, Germing U, Kliche KO, et al. A prospective, randomised, phase IIstudy of horse antithymocyte globulin vs rabbit antithymocyte globulin asimmune-modulating therapy in patients with low-risk myelodysplastic syn-dromes. Leukemia 2004;18:460–465.

40. Steensma DP, Dispenzieri A, Moore SB, et al. Antithymocyte globulin has limitedefficacy and substantial toxicity in unselected anemic patients with myelodys-plastic syndrome. Blood 2003;101:2156–2158.

41. Saunthararajah Y, Nakamura R, Wesley R, et al. A simple method to predictresponse to immunosuppressive therapy in patients with myelodysplastic syn-drome. Blood 2003;102:3025–3027.

42. Lim ZY, Killick S, Germing U, et al. Low IPSS score and bone marrow hypocellu-larity in MDS patients predict hematological responses to antithymocyte globu-lin. Leukemia 2007;21:1436–1441.

43. Ma X, Does M, Raza A, Mayne ST. Myelodysplastic syndromes: Incidence andsurvival in the United States. Cancer 2007;109:1536–1542.

44. Yue G, Hao S, Fadare O, et al. Hypocellularity in myelodysplastic syndrome isan independent factor which predicts a favorable outcome. Leuk Res 2008;32:553–558.

letters

American Journal of Hematology 201

Page 13: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

ASH 2009 meeting report—Top 10 clinically oriented abstractsin coagulation medicine and platelet disorders

Rajiv K. Pruthi*

The 2009 American Cancer Society of Hematology (ASH) annual meet-

ing took place on December 5–8, in New Orleans, LA. There were over

20,000 attendees with more than 4000 scientific presentations. Among

the latter, more than 300 were in the category of Basic and Clinical Sci-

ence of Coagulation Medicine and Platelet Disorders. Herein, are sum-

marized and discussed the top 10 abstracts in Coagulation Medicine

and Platelet disorders that are likely to have a practical impact on

patient care. All 10 abstracts are published in full in the November 20,

2009 issue of Blood (Volume 114, No. 22) and are identified here by

their abstract numbers.

Dabigatran Etexilate Versus Warfarin in the Treatment of

Venous Thromboembolism (Abstract #1)

Dabigatran etexilate is an oral formulation of the class of anticoagulants

termed direct thrombin inhibitors [1], with a potentially favorable therapeutic

index when compared with warfarin. In this randomized double blind trial,

2,539 patients (pt.) with acute venous thromboembolism (VTE), after initial

treatment with standard unfractionated heparin (UFH) or low-molecular-

weight heparin (LMWH), were randomized to receive a fixed dose of

oral dabigatran etexilate (150 mg twice daily) or adjusted dose warfarin

targeting an International Normalized Ratio (INR) of between 2.0 and 3.0,

for 6 months.

The rates of occurrence of the primary outcome of symptomatic, objec-

tively confirmed recurrent VTE was not significantly different between the

two arms. Toxicity of major and any bleeding also was not significantly differ-

ent. Rates of death, acute coronary syndromes, and liver function test

abnormalities were similar in the two groups with dyspepsia being more fre-

quent in pt. on dabigatran etilexate.

Thus, the efficacy and safety of fixed-dose dabigatran etexilate was similar

to warfarin in the 6 months of treatment of acute VTE.

Discussion

In this relatively short-term study, for which further details have recently

been published [2], there appeared to be no significant difference in efficacy

and safety between warfarin and dabigatran etexilate. However, given the

experience with a previous direct thrombin inhibitor [3], longer term data on

safety, especially hepatotoxicity is needed.

Once-Daily Oral Rivaroxaban Versus Placebo in the

Long-Term Prevention of Recurrent Symptomatic

Venous Thromboembolism. The Einstein-Extension

Study (Late Breaking Abstract #2)

In this international, randomized, double-blind placebo controlled trial,

patients who had completed a 6 to 12 month period of anticoagulation for an

acute venous thromboembolism (VTE) were randomized to either placebo (n

5 594) or a fixed daily dose (20 mg) of an oral factor Xa inhibitor (anti-Xa),

rivaroxaban (n 5 602) for secondary prophylaxis of VTE for an additional 6–

12 months. Excluded were patients in whom there was a clear indication for

ongoing anticoagulation. After a mean of 190 days on treatment, 7.1% and

1.3% of placebo and rivaroxaban treated patients respectively developed

recurrent VTE. There was no clinically significant difference in major bleeding

or mortality but an excess of clinically relevant nonmajor bleeding in the rivar-

oxaban arm was noted. Elevation of liver enzymes (ALT > three times and

total bilirubin > two times upper limit of normal) were not observed.

Discussion

Recurrent VTE results in high morbidity and mortality. It has been demon-

strated that long-term anticoagulation reduces the incidence of recurrent

VTE [4]; however, this comes at the cost of increasing risk of major and

minor hemorrhage. This trial demonstrates that in the short term, a fixed

daily dose of rivaroxaban is superior to placebo and likely safe. However,

more long-term safety data is needed.

High Incidence of Arterial Thrombosis in Young Patients

Treated for Multiple Myeloma: Results of a Prospective Cohort

Study (Abstract #149)

Patients (pt.) with multiple myeloma (MM) receiving thalidomide based

regimens are at high risk for venous thromboembolism (VTE). There is an

increasing recognition of development of arterial thromboembolism (ATE) in

this group of pt. The authors prospectively documented the incidence of

ATE in consecutive 195 newly diagnosed pt. with MM (age < 66 years). All

pt. were enrolled onto prospective, randomized Phase III trials of agents that

included various combinations of doxorubicin, dexamethasone, thalidomide,

vincristine, high-dose melphalan, interferon-alpha2 and bortezomib. Inci-

dence of ATE was compared with control populations of pt. (Framingham

Heart Study and the general Dutch population).

The overall incidence of ATE (myocardial infarction and ischemic

stroke) was 5.6% over a follow-up period of 522.4 patient-years with

most events occurring within one year after start of treatment. When com-

pared with the control population, this incidence was increased. After

adjustment for age and ISS stage, hypertension and current smoking

resulted in a increase in relative risk of 11.67 and 15.17, respectively.

Progressively increasing coagulation factor VIII levels conferred a higher

risk of ATE.

Discussion

A consensus on prevention and management of VTE in patients with mul-

tiple myeloma has recently been published [5], now there is an increasing

recognition of the risk of arterial thrombosis. Pending availability of interven-

tional guidelines, every effort should be made to address modifiable risk fac-

tors such as control of hypertension and smoking cessation. Further study is

needed to address the role of prophylactic antiplatelet or anticoagulant

agents in reducing risk of ATE.

Extended Follow-up of the Multicenter Multinational

Prospective Cohort Study That Derived the ‘‘Men Continue

and HERDOO2’’ Clinical Decision Rule Which Identifies

Low Risk Patients Who May be Able to Discontinue Oral

Anticoagulants 5–7 Months After Treatment for Unprovoked

Venous Thromboembolism (Abtract #451)

A clinical decision rule applied to patients with idiopathic venous throm-

boembolism (VTE) was found to be predictive of risk of recurrence [6]. Men

and women with �2 of the following (1) hyperpigmentation, edema, or red-

ness (HER) on examination in either leg, (2) Vidas D-Dimer (D) >250, (3)

Obesity(O): BMI > 30, or (4) Older age (O) over 65 were found to have a

high risk of recurrence after cessation of 6 months of anticoagulation. This

is a confirmatory, longer follow-up of 646 participants (mean age 53 years;

range 17–95) of whom 49% were women, from 11 centers experiencing a

first idiopathic VTE. During follow-up, mean 3.1 years (range 0.01–6.5),

overall the annual risk of recurrent VTE was 6.7%. Men had a 9.9% annual

risk whereas high risk women (�2 ‘‘HERDOO’’ points) had an annual risk of

recurrent VTE of 8.3%. Low-risk women (�1 ‘‘HERDOO’’ points) had 1.3%

annual risk of recurrent VTE.

Discussion

For secondary prophylaxis of VTE, the risk of OAC-related major hemor-

rhage (�1% to 3% annually) needs to be balanced against the risk of recur-

rent VTE off anticoagulation. The authors have previously reported on clini-

cal predictors of recurrent VTE [6]. This abstract provides longer term fol-

low-up confirming the validity of the clinical predictors. It should be noted

that patients with a known high risk thrombophilia, who likely are at higher

risk of recurrent VTE, were excluded from this study. Findings from such a

study provide a framework based on which decisions can be made regard-

ing duration of anticoagulation. Thus, men and high risk women with unpro-

voked VTE should be considered for long-term OAC therapy (with periodic

reassessment for safety) given a high risk of recurrence off anticoagulation.

letters

202 American Journal of Hematology

Page 14: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

Low-risk women (low HERDOO2 score) may be able to safely discontinue

anticoagulants, but should be educated on risks, symptoms, and signs of

recurrent VTE.

Low-Molecular-Weight Heparin Thromboprophylaxis in

Ambulatory Cancer Patients: A Systematic Review and

Meta-Analysis of Randomized Controlled

Trials (Abstract #490)

In this meta-analysis of seven randomized controlled trials (RCTs) of

venous thromboembolism (VTE) prophylaxis in ambulatory cancer patients

with nonhematologic malignancies, (n 5 2,960) of which 1,685 received low-

molecular-weight heparin (LMWH) and 1,275 controls. 2.79% of patients

receiving LMWH and 5.80% of control subjects experienced VTE, represent-

ing an absolute risk reduction of 2.55%. The absolute risk increase for major

bleeding was slightly increased 1.27%.

Discussion

Given the low risk of development of VTE in ambulatory cancer patients

in general, the risk of use of anticoagulant prophylaxis would have to be

weighed against its benefits, and is not currently recommended [7]. Thus, at

the present time, additional research into identification of high-risk sub-

groups [5] of ambulatory cancer patients for whom anticoagulant prophylaxis

would provide the most favorable to risk benefit ratio is ongoing and will

likely result in a modification of thromboprophylaxis recommendations for

such subgroups of patients.

A Phase III Study of Enoxaparin vs. Aspirin vs. Low-Dose

Warfarin as Thromboprophylaxis for Newly Diagnosed

Myeloma Patients Treated with Thalidomide-Based

Regimens (Abstract #492)

In this prospective, multicenter Phase III trial 991 newly diagnosed multi-

ple myeloma (MM) patients were randomized to three different therapeutic

trials including various combinations of agents such as velcade, thalidomide,

dexamethasone, melphalan, and prednisone. In a substudy, patients were

randomized to low-molecular-weight heparin (LMWH) (n 5 223) or low-dose

aspirin (ASA) (n 5 227), or low-fixed-dose (1.25 mg daily) warfarin (WAR)

(n 5 223) as anticoagulant prophylaxis. Patients on velcade–melphalan–pre-

dnisone (VMP; n 5 257) served as control patients and received no throm-

boprophylaxis.

There was no significant difference in the incidence of VTE, time of onset

of VTE, cardiovascular events, or bleeding in between the three anticoagu-

lant arms.

Discussion

In this trial, there was no difference in efficacy of VTE prophylaxis

between the different modalities and the overall incidence of VTE (<10%) in

all groups was not greater than expected for patients with MM. However this

patient population has been identified as one at increased risk of VTE, for

which guidelines have been published [5]. With the information available,

use of ASA seems a simple approach to VTE prophylaxis.

First-Line Rituximab Efficacy and Safety in Patients with

Acquired Idiopathic Thrombotic Thrombocytopenic Purpura

Experiencing a Nonoptimal Response to Therapeutical Plasma

Exchange: Results of a Prospective Multicenter Phase II Study

From the French Reference Center for the Management of

Thrombotic Microangiopathies (Abstract #890)

The role of rituximab in the management of thrombotic thrombocytopenic

purpura (TTP) is being defined [8]. In this Phase II trial, patients with TTP

being treated with daily total plasma exchange (TPE), but who had a ‘‘non-

optimal response,’’ defined as (a) refractory disease at day 5 or (b) a flare-

up of the disease within the first 15 days of standard intensive TPE treat-

ment were assigned to receive four standard doses of rituximab (n 5 22).

Their outcomes were compared with a historical cohort of 57 patients.

The data demonstrate no significant difference in mortality, mean time to

platelet recovery, and mean plasma volume required to achieve remission.

Significantly more patients in the historical control group were thrombocyto-

penic at day 35; however, data on platelet counts are not provided and the

clinical significance of this difference is unknown. Although recovery of

ADAMTS-13 activity was more rapid in patients receiving rituximab, there

was no difference at 12 months, and most importantly, there was no differ-

ence in the long-term risk of TTP relapse.

Discussion

The role of rituximab in management of TTP as part of initial therapy or for

refractory patients remains to be addressed. Current data are limited to case

series [9], and will likely not address this void of knowledge. Given its potential

serious toxicity [10], the use of rituximab should be confined to randomized con-

trolled clinical trials such as the one offered by the Transfusion Medicine/Hemo-

stasis Clinical Trials Network (http://www.tmhnetwork. org/protocols.htm).

Long-Term Follow-up Analysis Following Front Line

Therapy with Dexamethasone or Dexamethasone

Plus Rituximab in Adults with Primary Immune

Thrombocytopenia (Abstract #2415)

This abstract is an update from the initially reported (ASH 2008 Abstract #1)

prospective randomized trial of dexamethasone (DXM) vs. DXM and rituximab

(R) in newly diagnosed patients with primary immune thrombocytopenia (ITP).

Patients refractory to DXM alone were permitted to cross over into the DXM 1

R arm. This report extends the follow-up period to 36 months. Of the 101 origi-

nal patients, 21 were lost to follow-up. Of the 80 remaining patients only 53

were evaluable at a median follow-up of �20 months (4–40 months).

There was no significant difference in the duration of remission, relapse and

the need for additional therapy for ITP between the three groups of patients.

Apart from a single case of reactivation of herpes zoster in a patient initially

allocated to the DXM with cross over to R, no additional toxicities were noted.

Discussion

It is to be noted that only about half the original patients enrolled were evalu-

able for long term outcomes. In spite of the initial reported apparently superior

outcomes reported in 2008, longer term follow-up demonstrates no significant

advantage to the addition of rituximab as initial therapy. Rituximab’s role in refrac-

tory patients remains confined to case series and needs to be defined [11].

Thromboembolic Events Observed in Eltrombopag Clinical

Trials in Chronic Immune Thrombocytopenic

Purpura (Abstract #2423)

In this study, the authors present the incidence of and risk factors for throm-

boembolic events (TEE) in chronic immune thrombocytopenic purpura (ITP)

patients who received eltrombopag. This incidence is compared to the inci-

dence noted in this group of patients prior to initiation of eltrombopag. Data

were extracted from 446 patients enrolled on to three placebo-controlled and

two open-label eltrombopag studies. The incidence of TEE prior to enrollment

was 3.2% (16/493). After enrollment, 3.8% (17/446) of patients experienced

TEEs (which were all venous thromboembolism); however, these events all

occurred in patients treated with eltrombopag. All patients had a coexisting

risk factor (IVIg, no thromboprophylaxis during hospitalization and use of oral

corticosteroids) and only two patients had platelet counts >400,000.

Discussion

The incidence of thromboembolic events in patients with ITP is reported

to range from 3 to 6%. This study suggests the possibility that patients

receiving eltrombopag may be at risk for venous thromboembolic events,

however further long-term follow-up data are needed. The platelet count

appears not to be clearly associated with risk of TEE. It is important to care-

fully monitor patients receiving eltrombopag for TEEs.

Aspirin and Aspirin Combined with Low-Molecular-Weight

Heparin in Women with Unexplained Recurrent Miscarriage:

A Randomized Controlled Multicenter Trial (ALIFE Study)

(Abstract #488)

In this multicenter, blinded, randomized, controlled trial, 364 women (ages

18 to 42 years) with at least two recurrent miscarriages (RM) and with no

identifiable etiology (normal parental karyotype, no uterine pathology or

antiphospholipid antibodies, and a normal homocysteine), were randomized

to receive aspirin (ASA) or placebo or a combination of ASA and low-molec-

ular-weight heparin (LMWH). LMWH was administered in an open label

fashion. A prescheduled interim analysis of 281 women who had reached an

endpoint demonstrated no difference in live birth between treatment arms.

Discussion

Use of ASA and heparin has been shown to result in an improved live

birth rate among women with RM and antiphospholipid antibodies [12]. This

strategy has been empirically applied to women with RM and no identifiable

antiphospholipid antibodies. This randomized trial demonstrates no benefit

of such an intervention in this group of women.

letters

American Journal of Hematology 203

Page 15: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

Concluding Remarks

The clinical challenges in coagulation medicine include prevention and

treatment of thrombosis while balancing the risks of type and duration anti-

coagulant therapy. On the other end of the spectrum, options for prevention

and treatment of bleeding disorders need to be balanced with long-term tox-

icity. Clinical research efforts at defining the epidemiology of venous throm-

boembolism have led to development of guidelines which, with further clini-

cal research, will continue to be refined. Availability of novel anticoagulants

have the potential to further enhance the safety of long term anticoagulation.

Efforts at understanding the biology of thrombocytopenic disorders have led

to significant advances in therapeutic options. Ultimately, these options need

to be subjected to well-designed randomized trials with long-term follow-up

to provide our patients with the best possible outcomes.

Division of Hematology, Mayo Clinic, Rochester, Minnesota*Correspondence to: Rajiv K. Pruthi

Division of Hematology, Department of MedicineMayo Clinic, 200 First Street SW, Rochester, MN 55905

*E-mail: [email protected] of interest: Nothing to report

Published online 6 January 2010 in Wiley InterScience

(www.interscience.wiley.com).

DOI: 10.1002/ajh.21632

References

1. Garcia D, Libby E, Crowther MA. The new oral anticoagulants. Blood2010;115:15.

2. Schulman S, Kearon C, Kakkar AK, Mismettk P, Schellong S, Eriksson H, Baan-stra D, Schnee J, Goldhaber SZ. N Engl J Med 2009;361:2342.

3. Schulman S, Wahlander K, Lundstrom T, et al. Secondary prevention of venous

thromboembolism with the oral direct thrombin inhibitor ximelagatran. N Engl J

Med 2003;349:1713–1721.

4. Kearon C, Gent M, Hirsh J, et al. A comparison of three months of anticoagula-

tion with extended anticoagulation for a first episode of idiopathic venous throm-

boembolism. N Engl J Med 1999;340:901–907.

5. Palumbo A, Rajkumar SV, Dimopoulos MA, et al. Prevention of thalidomide-and lenalidomide-associated thrombosis in myeloma. Leukemia 2008;22:414–423.

6. Rodger MA, Kahn SR, Wells PS, et al. Identifying unprovoked thromboembolism

patients at low risk for recurrence who can discontinue anticoagulant therapy.

CMAJ 2008;179:417–426.7. Khorana AA, Streiff MB, Farge D, et al. Venous thromboembolism prophylaxis

and treatment in cancer: A consensus statement of major guidelines panels and

call to action. J Clin Oncol 2009;27:4919–4926.

8. Ling HT, Field JJ, Blinder MA. Sustained response with rituximab in patients with

thrombotic thrombocytopenic purpura: A report of 13 cases and review of the lit-

erature. Am J Hematol 2009;84:418–421.

9. Elliott MA, Heit JA, Pruthi RK, et al. Rituximab for refractory and or relapsing

thrombotic thrombocytopenic purpura related to immune-mediated severe

ADAMTS13-deficiency: A report of four cases and a systematic review of the lit-

erature. Eur J Haematol 2009;83:365–372.

10. Carson KR, Evens AM, Richey EA, et al. Progressive multifocal leukoence-

phalopathy after rituximab therapy in HIV-negative patients: A report of 57 cases

from the research on adverse drug events and reports project. Blood

2009;113:4834–4840.11. Hasan A, Michel M, Patel V, et al. Repeated courses of rituximab in chronic ITP:

Three different regimens. Am J Hematol 2009;84:661–665.12. Empson M, Lassere M, Craig J, et al. Prevention of recurrent miscarriage for

women with antiphospholipid antibody or lupus anticoagulant. Cochrane data-

base of systematic reviews (online). 2005;2:CD002859.

ASH 2009 meeting report—Top 10 clinically oriented abstractsin sickle cell diseaseCarlo Brugnara*

At the American Society of Hematology 2009 meeting in New Orleans,

Louisiana, more than 100 abstracts were selected for either oral or

poster presentation in the ‘‘Hemoglobinopathies, excluding Thalasse-

mia’’ category. Here, I summarize and discuss the top 10 abstracts for

sickle cell disease (SCD), which present relevant information which

may affect clinical practice. All 10 abstracts are published in full in the

November 20, 2009 issue of Blood (volume 114, No. 22) and are identi-

fied here by their abstract numbers.

Pulmonary hypertension has been recognized as a severe and potentially

fatal complication of SCD [1–6]. Its pathogenesis has been linked to hemolysis

with increased levels of plasma hemoglobin and an altered regulation of NO

(nitric oxide) metabolism leading to NO deficiency. Doppler echocardiographic

studies measuring tricuspid regurgitant jet velocity (TRV) as a proxy marker of

pulmonary hypertension have yielded relatively high estimates (32–41%) of the

prevalence of pulmonary hypertension (defined by a TRV equal or greater than

2.5 m per second) [7–9] with lower estimates (7%) in Nigerian patients [10,11].

However, no systematic study has been conducted on the true prevalence of

pulmonary hypertension using the gold-standard method of cardiac catheteriza-

tion. On the therapeutic side, the NO deficient state of SCD and the role of

NO in the pathogenesis of pulmonary hypertension have prompted clinical

studies to assess various strategies to normalize NO availability, either by

providing NO as an inhaled gas, [12] or by pharmacological manipulation

of NO metabolism [13]. The first five abstract discussed here present

novel, clinically relevant information on pulmonary hypertension patho-

physiology, diagnosis, and therapy.

1. Prospective Multicentric Survey on Pulmonary

Hypertension (PH) in Adults with Sickle Cell Disease (#572)Abstract summary

Doppler echocardiography, pulmonary function tests and 6 min walk test

were performed in 403 consecutive outpatients with SCD in stable clinical con-

ditions. 96 patients (25%) had elevated TRV (>2.5 m/sec) and underwent con-

firmatory right heart catheterism (RHC), using a measured pulmonary artery

pressure of �25 mmHg as diagnostic threshold for pulmonary hypertension.

Elevated PAP was found in 24/96 patients. These patients were older, had ele-

vated levels of plasma LDH and NT-Pro BNP, and were able to walk for a

lower distance in the 6 min walk test. Pulmonary hypertension, when deter-

mined by the gold standard RHC method, seems to be rare in SCD patients,

with a prevalence of 6%. Among these patients, only a small subgroup exhibits

precapillary pulmonary arterial hypertension normal capillary wedge pressure,

increased vascular resistance; 1.6% of the total population.

Discussion

Data from this study question the previously published data showing a

much greater prevalence of pulmonary hypertension in SCD patients, while

they seem to confirm the increased mortality associated with this condition.

As RHC may not be readily available and indicated in all patients with ele-

vated TRV, future studies should be focused on improving the methodology

used to screen patients for this severe and fatal complication. TRV should

not be used to diagnose pulmonary hypertension in SCD patients in the

absence of confirmatory studies.

2. Safety and Efficacy of Sildenafil Therapy for

Doppler-Defined Pulmonary Hypertension in

Patients with Sickle Cell Disease: Preliminary

Results of the Walk-PHaSST Clinical Trial (# 571)Abstract summary

The Walk-PHaSST (treatment of Pulmonary Hypertension and SCD with

Sildenafil Therapy) was a multicenter, placebo-controlled, double-blind study

to assess safety and efficacy of oral sildenafil in the treatment of Doppler-

defined PH (TRV > 2.7 m/sec) in children (>12 years) and adults with SCD.

Seventy four subjects had been randomized into the study when the study

was stopped due to a statistically significant increase in serious adverse

event in the sildenafil arm. In addition, no improvements were seen in either

TRV or 6 min walk distance after 16 weeks of sildenafil therapy.

letters

204 American Journal of Hematology

Page 16: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

Discussion

The disappointing results of this study throw into question the NO hypoth-

esis for the pathogenesis of vascular complications of SCD [14]. The lack of

efficacy of sildenafil in TRV-defined PH in conjunction with the results of the

study discussed earlier may suggest that different methods are needed to

diagnose PH in SCD. However, a positive and worth pursuing outcome of

this study may lie in the identification of the role of NO in the pain mecha-

nisms associated with SCD. In addition, the screening of more than 700

subjects carried out for this study provides a unique repository of data and

clinical material for future studies.

3. Arginine Therapy for Vaso-Occlusive Pain Episodes in

Sickle Cell Disease (#573)Abstract summary

This single center, double-blind, placebo control study enrolled 56 patients

to assess safety and efficacy of IV or oral arginine (0.1 g/kg TID) for the

treatment of vaso-occlusive events (VOE) in SCD. Patients were >3 years

of age and needed to be diagnosed with VOE within 24 h to be enrolled in

the study. A standardized protocol for treatment and monitoring of VOE and

pain was used in all patients. Use of arginine was associated with a signifi-

cant (over 50%) reduction in the total usage of morphine during the hospital

stay (mean ± SEM: 1.8 ± 0.4 mg/kg; n 5 28 vs. 4.1 ± 0.8 mg/kg; n 5 26, p

5 0.01) with no changes in the length of hospitalization. There were no sig-

nificant changes in either laboratory parameters or adverse events associ-

ated with the use of arginine.

Discussion

The reduction in narcotic usage with arginine therapy is remarkable,

and worth further investigation. The results of this study are difficult to

reconcile with the negative result of the sildenafil study discussed earlier,

and the associated hypothesis that the increase adverse events may be

due to NO increasing pain perception. However, an important outcome

of these studies may be a renewed focus on understanding the patho-

physiology of pain in SCD and developing novel analgesic approaches.

4. Hemolysis-Associated Elevation in Tricuspid Regurgitation

Velocity Predicts Reduction in Six-Minute Walk Distance After

Two Years of Follow-up in Children and Adolescents with

Sickle Cell Disease (#574)Abstract summary

The Pulmonary Hypertension and the Hypoxic Response in SCD (PUSH)

study is a multicenter effort to define baseline TRV, hemolytic parameters,

and six-minute walk test and to observe the behavior of these parameters

longitudinally in a cohort of children and adolescents with SCD. 361 patients

have been enrolled so far and follow-up data are available on 193 patients.

21/193 patients had TRV of 2.6 m/sec or greater at baseline, with 15 of the

21 patients showing high hemolytic rate. The six-minute walk test showed a

significant decline (10%) at follow-up in the patients with elevated TRV.

Discussion

Elevated TRV in association with hemolysis is seen in approximately 8%

of children and adolescents with SCD. It remains to be determined if

pediatric patients with elevated TRV are more likely to exhibit increased

morbidity/mortality and develop pulmonary hypertension. Longitudinal

studies like the PUSH study are important to determine variation in TRV

and 6-min walk with growth and development in children with SCD, and to

identify factors predictive of pulmonary hypertension development in adulthood.

5. Retrospective Review of the Natural History of Pulmonary

Hypertension in Sickle Cell Disease Demonstrates that

Progressive Enlargement of the Left Atrium is a Strong

Predictor of Death (# 1529)Abstract summary

A systematic, single center, chart, and echocardiogram review of 362

patients was carried out to identify how TRV hospitalization values may help

in predict steady state progression of pulmonary hypertension. A mean

yearly rate of progression for TRV of 0.04 m/s (p < 0.001, 95% CI 0.02 to

0.05) was identified in patients with multiple measurements over time. Con-

trary to the initial hypothesis, high TRV values recorded during admissions

did not seem to predict the rate of progression of pulmonary hypertension

during steady state. In addition, no correlates were found between rate of

progression of TRV at baseline conditions and mortality, although inpatient

TRV progression seemed to be higher for patient who subsequently died.

Left atrial enlargement was also identified as a novel predictor of death.

Discussion

As highlighted in the earlier studies, TRV has substantial limitation in pre-

dicting both pulmonary hypertension and disease related complications and

progression in SCD. Better and larger longitudinal studies are required to

identify the subset of SCD patient more at risk of developing pulmonary

hypertension and associated morbidity/mortality.

6. The Effect of Short-Term Simvastatin On Markers of

Vascular Dysfunction in Patients with Sickle Cell

Disease (SCD) (# 260)Abstract summary

The use of multiple chemotherapeutic approaches has been advocated to

treat and prevent complications associated with the vasculopathy of SCD

[15,16]. Statins have been shown to exert multiple anti-inflammatory/vascu-

lar effects and are thus potential candidates for SCD therapy. In addition,

some of these effects may be mediated via NO-dependent mechanism,

which are known to be affected in SCD. In this study, a small number of

patients with SCD [12] are being treated with low-dose oral simvastatin (20

mg/day) with the intent of describing changes in multiple biomarkers of

inflammation and vasculopathy after short-term (21 days) therapy. Prelimi-

nary data show a 24% increase in mean plasma NOx level (p 5 0.02), with

a concomitant decrease in both hs-CRP (p 5 0.01) and IL-6 levels (p 5

0.05) and no changes in plasma levels of tissue factor (TF), vascular endo-

thelial growth factor (VEGF) and vascular cell adhesion molecule (VCAM1).

Discussion

Additional long-term studies will be needed to properly assess efficacy

and safety of statins in the polychemotherapy of SCD vasculopathy.

7. Genetic Predictors of Hydroxyurea Response in

Children with Sickle Cell Disease (# 820)Abstract summary

Hydroxyurea therapy has shown to be remarkably effective in reducing

SCD complications and overall SCD-related morbidity/mortality [17]. Given

the high cost of care for patients with SCD, [18] optimization of therapy

and identification of nonresponders are crucial to prevent morbidity and

mortality. Previous studies have identified important laboratory predictors

of response to hydroxyurea, such as baseline %HbF, Hb, reticulocyte, and

WBC counts, and DNA levels [19,20]. In this study, candidate genes pre-

sumably involved in hydroxyurea pharmacokinetics (PK) or pharmacody-

namics (PD) were sequenced for a group of pediatric SCD patients partici-

pating in the prospective Hydroxyurea Study of Long-term Effects

(HUSTLE, NCT00305175). A variety of known and novel SNPs were asso-

ciated with hydroxyurea responsiveness. They involved known Hb F regu-

latory genes as well as the urea transporter UTB.

Discussion

Studies in large cohort of hydroxyurea-treated pediatric and adult SCD

patients are needed to identify all relevant genetic polymorphisms associ-

ated with treatment response and/or toxicity to hydroxyurea. These studies

should serve as prototypes for future drug candidates for SCD.

8. Prevalence of Nocturnal Hypoxia and Its Association with

Disease Severity in Adults with Sickle Cell Disease (# 261)Abstract summary

Pulmonary complications are an important cause of morbidity and mortal-

ity in SCD adult patients. Pulmonary function studies have identified a sub-

stantial decline in performance with age in adult patients with SCD [21].

Pediatric studies have identified subset of children with severe nocturnal

desaturation and hypercapnea associated with obstructive sleep apnea

syndrome (OSA) [22]. Therapeutic interventions to correct nocturnal desa-

turation in SCD children have shown promising results [23]. However, few

studies have addressed these issues in SCD adult patients. With the use

of a questionnaire screening, 22 patients were studied with overnight oxy-

metry. Of these, 11 showed clear signs of OSA and an additional six

showed nocturnal hypoxia without OSA. Nocturnal hypoxia was correlated

with some measures of organ damage, such as glomerular filtration rate

and priapism. A low daytime saturation seemed to be a good predictor of

night-time hypoxia.

letters

American Journal of Hematology 205

Page 17: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

Discussion

This study highlights the central role of pulmonary function in the patho-

physiology of SCD complications. A systematic approach focused on pulmo-

nary function studies, asthma prevention/treatment, and therapeutic inter-

vention to avoid nocturnal hypoxia/desaturation seems to be warranted for

both pediatric and for adult patients with SCD [24].

9. Increased Severity of Pandemic H1N1 Influenza in Children

and Young Adults with Sickle Cell Disease (# 264)Abstract summary

Children with SCD require admission for influenza at a frequency 50-fold

greater than the general population. Ninety-nine patients with sickle cell ane-

mia and influenza were identified for the period 1993–2009. Among these

patients, 89 had seasonal influenza A or B, and 10 had pandemic H1N1

infection confirmed using a real-time reverse transcription polymerase chain

reaction assay. There were no differences in clinical symptoms between the

two groups. However, patients infected with HIN1 virus were more likely to

have acute chest syndrome (3-fold increased risk), require intensive care,

and mechanical ventilation (9-fold increased risk).

Discussion

The data suggest the need for prophylaxis with vaccination against sea-

sonal and HIN1 influenza for both pediatric patients with SCD and family

members. However, most of the patients in the seasonal influenza group

were included via a retrospective survey going back to 1993. For the

2008–2009 season, only eight patients had seasonal influenza, a number

that is too small to allow meaningful comparisons over the same time

period with the pandemic H1N1 influenza. It remains to be determined if

the baseline clinical severity and care intensity are different between the

two groups and what are the overall vaccination and infection rate for both

seasonal and pandemic influenza in an unselected population of patients

with SCD.

10. Effects of Hydroxyurea (HU) and Magnesium Pidolate (Mg)

in Hemoglobin SC Disease (HbSC): The ‘‘CHAMPS’’

Trial (# 809)Abstract summary

The CHAMPS Trial is the first prospective, randomized, double-blinded,

multi-center Phase II study to be carried out in patients with Hb SC dis-

ease. The objective of this study was to determine the ability of hydrox-

yurea (HU) and Mg pidolate, alone and in combination, to reduced the

density (assessed as cell hemoglobin concentration) in children and adults

with HbSC disease. The study was designed as a four arms study with a

target enrollment of 188 subjects with Hb SC disease, 5 years of age and

older, who had experienced at least one VOE (pain, acute chest syn-

drome) in the previous 12 months. Due to slow enrollment, the study was

prematurely terminated. A total of 40 patients could be evaluated with 36

having reached the primary endpoint at week 8, and 22 patients having

completed 11 months of treatment. Significant changes in MCV, MCH, and

Hb F were observed with HU, while MG had no effect. Hb level, hyper-

dense red cells, erythrocyte Na, K, and Mg, KCl cotransport and Gardos

channel activity, plasma magnesium, serum LDH, red cell PS exposure,

and adhesion to endothelium showed no significant differences among the

four groups.

Discussion

The premature termination of the study did not allow accrual of a number

of a sufficient number of patients to assess effects on any of the primary

and secondary endpoints. However, the study confirmed prior, smaller, open

label studies, which had shown cellular and Hb F changes with hydroxyurea

treatment in patients with SC disease [25,26]. The lack of any Mg effect is

disappointing, as there were expectation that SC patients would exhibit simi-

lar responses to those seen in patients with SS disease [27,28]. It remains

to be determined if doses higher than 0.6 mEq/Kg/day, as the 0.9 mEq/Kg/

day determined to be safe and effective in patients with SS disease [29],

could induce cellular changes in SC erythrocytes.

Disclosure

Dr. Carlo Brugnara is one of the inventors for US Patent 6,331,557 ‘‘Use

of divalent cations for inhibiting erythrocyte dehydration in vivo.’’ issued on

December 18, 2001.

Department of Laboratory Medicine, Children’s Hospital Boston, Boston,Massachusetts

*Correspondence to: Carlo Brugnara, Department of Laboratory Medicine,Children’s Hospital Boston, 300 Longwood Avenue, Bader 760, Boston,

MA 02115. E-mail: [email protected] for publication 14 December 2009; Accepted 14 December 2009

Conflict of interest: Nothing to report.Published online 8 January 2010 in Wiley InterScience

(www.interscience.wiley.com).DOI: 10.1002/ajh.21636

References

1. Gladwin MT, Vichinsky E. Pulmonary complications of sickle cell disease. NEngl J Med 2008;359:2254–2265.

2. Klings ES. Pulmonary hypertension of sickle cell disease: More than justanother lung disease. Am J Hematol 2008;83:4–5.

3. Klings ES, Anton Bland D, Rosenman D, et al. Pulmonary arterial hypertension andleft-sided heart disease in sickle cell disease: Clinical characteristics and associa-tion with soluble adhesion molecule expression. Am J Hematol 2008;83:547–553.

4. Taylor JG VI, Ackah D, Cobb C, et al. Mutations and polymorphisms in hemoglo-bin genes and the risk of pulmonary hypertension and death in sickle cell dis-ease. Am J Hematol 2008;83:6–14.

5. Gordeuk VR, Sachdev V, Taylor JG, et al. Relative systemic hypertension inpatients with sickle cell disease is associated with risk of pulmonary hyperten-sion and renal insufficiency. Am J Hematol 2008;83:15–18.

6. Inamo J, Connes P, BarthElEmy J-C, et al. Pulmonary hypertension does notaffect the autonomic nervous system dysfunction of sickle cell disease. Am JHematol 2009;84:311–312.

7. Gladwin MT, Sachdev V, Jison ML, et al. Pulmonary hypertension as a risk factorfor death in patients with sickle cell disease. N Engl J Med 2004;350:886–895.

8. De Castro LM, Jonassaint JC, Graham FL, et al. Pulmonary hypertension asso-ciated with sickle cell disease: Clinical and laboratory endpoints and diseaseoutcomes. Am J Hematol 2008;83:19–25.

9. van Beers EJ, Nur E, Schaefer-Prokop CM, et al. Cardiopulmonary imaging,functional and laboratory studies in sickle cell disease associated pulmonaryhypertension. Am J Hematol 2008;83:850–854.

10. Aliyu ZY, Gordeuk V, Sachdev V, et al. Prevalence and risk factors for pulmonaryartery systolic hypertension among sickle cell disease patients in Nigeria. Am JHematol 2008;83:485–490.

11. Aliyu ZY, Kato GJ, Taylor IV, et al. Sickle cell disease and pulmonary hyperten-

sion in Africa: A global perspective and review of epidemiology, pathophysiology,

and management. Am J Hematol 2008;83:63–70.

12. Weiner DL, Hibberd PL, Betit P, et al. Preliminary assessment of inhaled nitric

oxide for acute vaso-occlusive crisis in pediatric patients with sickle cell disease.

JAMA 2003;289:1136–1142.13. Morris CR, Kuypers FA, Larkin S, et al. Arginine therapy: A novel strategy to induce

nitric oxide production in sickle cell disease. Br J Haematol 2000;111:498–500.

14. Morris CR, Kato GJ, Poljakovic M, et al. Dysregulated arginine metabolism,

hemolysis-associated pulmonary hypertension, and mortality in sickle cell dis-

ease. JAMA 2005;294:81–90.

15. Hebbel RP. The systems biology-based argument for taking a bold step in che-

moprophylaxis of sickle vasculopathy. Am J Hematol 2009;84:543–545.16. Kato GJ, Hebbel RP, Steinberg MH, Gladwin MT. Vasculopathy in sickle cell dis-

ease: Biology, pathophysiology, genetics, translational medicine, and new

research directions. Am J Hematol 2009;84:618–625.

17. Steinberg MH, Barton F, Castro O, et al. Effect of Hydroxyurea on Mortality andMorbidity in Adult Sickle Cell Anemia: Risks and Benefits Up to 9 Years of Treat-ment. JAMA 2003;289:1645–1651.

18. Kauf TL, Coates TD, Huazhi L, et al. The cost of health care for children andadults with sickle cell disease. Am J Hematol 2009;84:323–327.

19. Ware RE, Eggleston B, Redding-Lallinger R, et al. Predictors of fetal hemoglobinresponse in children with sickle cell anemia receiving hydroxyurea therapy.Blood 2002;99:10–14.

20. Ulug P, Vasavda N, Kumar R, et al. Hydroxyurea therapy lowers circulating DNAlevels in sickle cell anemia. Am J Hematol 2008;83:714–716.

21. Field JJ, Glassberg J, Gilmore A, et al. Longitudinal analysis of pulmonary func-tion in adults with sickle cell disease. Am J Hematol 2008;83:574–576.

22. Kaleyias J, Mostofi N, Grant M, et al. Severity of obstructive sleep apnea inchildren with sickle cell disease. J Pediatr Hematol Oncol 2008;30:659–665.

23. Marshall MJ, Bucks RS, Hogan AM, et al. Auto-adjusting positive airway pres-sure in children with sickle cell anemia: Results of a phase I randomized con-trolled trial. Haematologica 2009;94:1006–1010.

24. Morris CR. Asthma management: Reinventing the wheel in sickle cell disease.Am J Hematol 2009;84:234–241.

25. Steinberg MH, Nagel RL, Brugnara C. Cellular effects of hydroxyurea is Hb SCdisease. Br J Haematol 1997;98:838–844.

26. Iyer R, Baliga R, Nagel RL, et al. Maximum urine concentrating ability in childrenwith Hb SC disease: Effects of hydroxyurea. Am J Hematol 2000;64:47–52.

27. De Franceschi L, Bachir D, Galacteros F, et al. Oral magnesium supplementsreduce erythrocyte dehydration in patients with sickle cell disease. J Clin Invest1997;100:1847–1852.

28. De Franceschi L, Bachir D, Galacteros F, et al. Oral magnesium pidolate: Effectsof long-term administration in patients with sickle cell disease. Br J Haematol2000;108:248–289.

29. Hankins JS, Wynn LW, Brugnara C, et al. Phase I study of magnesium pidolatein combination with hydroxycarbamide for children with sickle cell anaemia. Br JHaematol 2008;140:80–85.

letters

206 American Journal of Hematology

Page 18: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

ASH 2009 meeting report—Top 10 clinically oriented abstractsin chronic lymphocytic leukemiaSusan O’Brien*

At the American Society of Hematology 2009 meeting in New Orleans,

Louisiana, a large number of abstracts were selected for either oral or

poster presentation in the various ‘‘Chronic lymphocytic leukemia (CLL)’’

categories. Here, I summarize and discuss the top 10 abstracts, which

present relevant information which may affect clinical practice. All 10

abstracts are published in full in the November 20, 2009 issue of Blood

(volume 114, No. 22) and are identified here by their abstract numbers.

1. First Line Treatment with Fludarabine (F), Cyclophospha-

mide �, and Rituximab (R) (FCR) Improves Overall Survival

(OS) in Previously Untreated Patients (pts) with Advanced

CLL: Results of a Randomized Phase III Trial on Behalf of an

International Group of Investigators and the German CLL

Study Group. (#535)

Abstract summary

The German CLL Study Group presented follow-up on a clinical trial that

was initially presented in ASH 2008. This was a large randomized trial of

over 800 patients who were treatment naıve that went on to receive either

fludarabine and cyclophosphamide (FC) chemotherapy or the same chemo-

therapy with the addition of one dose of rituximab (FCR). Cycles were

administered monthly for a total of 6 months. FCR induced a higher overall

response rate than FC (95.1 vs 88.4%) and more complete remissions (44.1

vs 21.8%; P < 0.001). In addition, the median PFS was longer for FCR at

51.8 months, compared to FC at 32.8 months (P < 0.001). Although more

neutropenia was seen on the FCR arm this did not translate into a greater

number of grade 3–4 infections; the incidence was not significantly different.

What was remarkable and what was presented this year was that this

randomized trial is now showing a statistically significant survival difference

in favor of FCR. The overall survival rate at 37.7 months was 84.1% in the

FCR arm vs 79.0% in the FC arm (P 5 0.01).

Discussion

This is the first randomized trial that has ever shown a survival advant-

age for a frontline therapy in CLL and speaks to the dramatic improve-

ment in complete remission rates that are achieved with adding one dose

of rituximab to FC chemotherapy every month. Although as a single agent

rituximab has minimal activity in CLL it has definitely transformed the

face of treatment with chemoimmunotherapy now being the new standard

of care.

2. Long-Term Survival Analysis of the North American

Intergroup Study C9011 Comparing Fludarabine and

Chlorambucil in Previously Untreated Patients

with CLL. (#536)Abstract summary

In keeping with the theme of better survival with improved frontline therapy

this is a followup on a much older trial that started about 20 years ago,

which randomized previously untreated patients with symptomatic CLL to flu-

darabine or chlorambucil or the combination of fludarabine and chlorambucil.

This was an Intergroup study that was published in the New England Jour-

nal of Medicine in 2000 [1]. It showed that fludarabine provided significantly

higher overall response rates, CR rates and longer progression free survival

(P < 0.001 for all endpoints). The combination arm was stopped early

because of increased morbidity and mortality. When this data was originally

published in 2000, there was no difference in overall survival among the

three groups. At this year’s meeting with nearly 10 more years of followup

(ending in January 2009) the difference has become significant. Of 509

patients, 85% have now died. As noted previously, fludarabine treatment

resulted in significantly longer PFS than did chlorambucil. Improved survival

benefit with fludarabine emerged after 5–6 years and the 8 year survival

with fludarabine was 31% vs 19% with chlorambucil vs 26% with the combi-

nation.

Discussion

Discussion of the prior abstract confirms that chemoimmunotherapy is the

new standard of care; single agent therapy is rarely used except in elderly

patients. Nevertheless for the first time ever 2 ASH presentations on CLL

showed improved survival based on choice of initial therapy. Although the fol-

low-up in the German CLL study of FCR vs FC was 37.7 months, the survivals

at this point in time were significantly better in both the FCR and the FC arm

than the survivals at 4 years with either fludarabine or chlorambucil, again vali-

dating that better therapies lead to improved survival.

3. Bendamustine Combined with Rituximab (BR) in First Line

Therapy of Advanced CLL: A Multicenter Phase II Trial of the

German CLL Study Group. (#205)Abstract summary

Bendamustine is a novel alkylating agent that has shown considerable activ-

ity as monotherapy for lymphoid malignancies including lymphoma and CLL

[2–4]. Preclinical data has suggested synergy when bendamustine is com-

bined with rituximab and evaluated against primary CLL cells [5]. Last year at

ASH the Germans presented data from the phase 2 trial using BR for patients

with relapsed CLL and showed a good overall response rate of 77% with 15%

CR. Based on this encouraging activity in relapsed patients, the German CLL

Study Group then went on to do a frontline trial with BR and presented data

on 117 patients with previously untreated CLL who received bendamustine at

a dose of 90 mg/m2 on days 1 and 2 combined with 375 mg/m2 rituximab for

the first cycle and 500 mg/m2 for subsequent cycles. The BR therapy was

given every 28 days for up to six cycles. The most frequent side effects were

myelosuppression and infection: grade 3–4 neutropenia was seen in 6.5% of

all cycles and grade 3–4 thrombocytopenia was seen in 6.1% of all cycles.

Twenty-nine episodes of infections �3 were documented (5.1% of all cycles).

The overall response rate was 90.9% with 32.7% complete remission, nodular

partial remission in 2.7%, and PR in 55.5%. After 18 months 75.8% of the

patients were still in remission and the median PFS had not been reached.

Impressively, an MRD level below 1024 was observed after completion of ther-

apy in 29 of 50 evaluable patients in the peripheral blood while 7 of 25

patients achieved MRD negativity in the bone marrow. In the high risk group

with 17p deletions, 3 of 7 patients showed a partial response (ORR:42.9%).

Discussion

Bendamustine and rituximab comprise an effective and relatively safe regi-

men for the treatment of CLL. As with most chemotherapeutic agents, the

major side effects are related to myelosuppression and infection. The myelo-

suppression rate appears mild although analyzing this data as percentage of

all cycles can underestimate the toxicity to an individual patient. That is

because patients who are having difficulties or problems with infections often

come off therapy early and patients doing well continue so that as time goes

on the toxicities tends to lessen when looked at as a percentage of all

cycles. Nevertheless the Germans consider this data impressive enough

that the current frontline randomized trial of the German CLL Study Group is

BR vs FCR.

4. Ofatumumab Combined with Fludarabine and

Cyclophosphamide (OFC) Shows High Activity in Patients

with Previously Untreated CLL: Results from a Randomized

Multicenter International Two Dose Parallel Group Phase II

Trial. (#207)Abstract summary

This is a fully humanized monoclonal antibody that targets a unique small

loop epitope on CD20, which is not bound by rituximab. Preclinically this

agent elicits very rapid and efficient compliment dependent cytotoxicity as

well as ADCC. In cells with low antigen density, which is particularly relevant

to CLL as opposed to lymphoma, this agent produced significantly more

CDC than did comparable doses of rituximab [6,7]. The trial presented at

letters

American Journal of Hematology 207

Page 19: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

ASH this year was conducted in previously untreated patients with CLL.

Sixty-one patients were randomized to receive ofatumumab 500 mg (Group

A) or 1000 mg (Group B) on day 1 combined with standard FC chemother-

apy (fludarabine 25 mg/m2 IV daily and cyclophosphamide 250 mg/IV daily,

both for 3 days). Up to six courses could be given every 4 weeks. In both

groups the first dose of ofatumumab was 300 mg; this dose was lower to

minimize infusion reactions. The primary endpoint of the trial was complete

remission rate using the 1996 NCI/WG criteria, 71% and 57% of patients in

Groups A and B, respectively, completed all six courses of OFC treatment.

The CR rate as assessed by an independent review committee was 32% for

Group A and 50% for Group B; the overall response rate was 77% and

73%, respectively. The follow-up was short so the median PFS had not

been reached. No CTC grade 3–4 infusion reactions were seen; all patients

were premedicated with acetaminophen, antihistamine, and additionally, glu-

cocorticoids before infusions 1 and 2. The most common adverse events

were infections in 11 patients (Group A N 5 4; Group B, N 5 7).

Discussion

The trial that was presented last year at ASH examined the efficacy of the sin-

gle agent given for 8 weeks followed by four monthly doses in highly refractory

patients, namely those who were doubly refractory to both fludarabine and alem-

tuzumab or who were refractory to fludarabine and had bulky adenopathy, so

not thought to be good candidates for alemtuzumab therapy. The overall

response rate was approximately 50% in this highly refractory group and this led

to the recent FDA approval of this agent. Using the higher dose of 1 g (as

opposed to the FDA approved single agent dose of 2 g) the CR rate with OFC

appears very comparable to the CR rate seen with FCR in the German random-

ized trial mentioned in Abstract 1. At this point in time the data do not appear to

be better than what can be achieved with FCR although longer follow-up to eval-

uate PFS, etc. is needed. Although a remarkably active single agent the data

thus far do not show any enhanced benefit from use of ofatumumab with chemo-

therapy over that achieved with rituximab; admittedly the data is very limited and

other combination trials using ofatumumab are ongoing.

5. Combination Therapy with Lenalidomide and Rituximab in

Patients with Relapsed CLL. (#206)Abstract summary

Lenalidomide is a immunomodulatory agent that has previously been shown

to have significant activity in the treatment of relapsed CLL [8,9]. Overall

response rates range from 32 to 40%. As discussed previously, rituximab has

minimal activity as monotherapy but significant synergistic interaction with multi-

ple chemotherapeutic agents. It had previously been shown that the addition of

rituximab to lenalidomide resulted in responses in a small number of patients

with CLL that had progressed while on lenalidomide alone. Because lenalido-

mide stimulates NK cell proliferation there is reason to think that lenalidomide

would enhance the activity of rituximab. In this phase II trial 60 patients received

the combination of lenalidomide and rituximab. All patients received rituximab

375 mg/m2 intravenously on days 1, 8, 15, and 22 of cycle 1 and then once a

month from cycles 3–12. Lenalidomide was given orally at a dose of 10 mg daily

starting on day 9 of cycle 1 and continued daily for 12 cycles. Each cycle con-

sisted of 28 days of treatment. When this presentation was made 37 patients

had received treatment for at least six cycles and were evaluated for response

and toxicity. The median number of previous treatments was 2 and 24% of the

patients were refractory to fludarabine; all patients had received prior rituximab.

Twenty-four percent of the patients had a chromosome 17p deletion. After six

cycles of treatment the overall response rate was 68% with 51% PR and 16%

nodular PR. Six patients (16%) had achieved only stable disease or some

improvement that did not qualify as PR but are continuing on treatment. Six

patients failed to respond including one death that occurred on day 34 related to

infection. The most common grade 3–4 treatment related side effects were neu-

tropenia in 43%, fatigue in 16%, and thrombocytopenia in 11%. Although all

patients received allopurinol, one patient developed grade 3 tumor lysis syn-

drome. Infectious complications were seen in 24% of patients. The lenalidomide

associated tumor flare reaction was seen in nine patients (25%) and was limited

to grade 1 in eight patients and grade 2 in one patient. Of note is that six of nine

patients with deletion 17p (67%) responded to treatment.

Discussion

When we compare this to historical data using lenalidomide as a single

agent, it would appear that the response rate almost doubles with the addi-

tion of rituximab, with little increase in toxicity. In fact the low incidence of

tumor lysis and the minimal flare reactions are likely related to the fact that

rituximab is initiated first and the lenalidomide does not begin until a week

later, allowing for some tumor debulking to occur. This certainly is in keeping

with the theme that adding rituximab to other agents appears to produce at

least additive if not synergistic activity in lymphoid malignancies and now we

can expand this conclusion to the combination of rituximab and imids.

Although the number of patients with 17p deletions was small the response

rate was impressive in this group.

6. Front Line Combined Chemoimmunotherapy with

Fludarabine, Cyclophosphamide, Alemtuzumab, and

Rituximab (CFAR) in High Risk CLL. (#208)Abstract summary

We have already seen that chemoimmunotherapy with FCR produces

excellent results when used as frontline therapy for patients with CLL. In the

MD Anderson experience there were a subset of patients treated with FCR

who had lower CR rates and shorter time to progression and overall sur-

vival. This group was characterized by having a serum b2 microglobulin >/

5 4 mg/L [10,11]. This group with a high b2 microglobulin also encompasse

many of those patients with 17p deletions. Thus the rationale behind the cur-

rent regimen was to intensify FCR by adding alemtuzumab. All patients

were less than 70-years-old and had a b2 microglobulin >/5 4 mg/L. CFAR

consisted of C 200 mg/m2 days 3–5, F 20 mg/m2 days 3–5, A 30 mg IV

days 1, 3, and 5 and R 375 mg/m2 on cycle 1 and 500 mg/m2 on subse-

quent cycles. In other words, this is standard FCR with the addition of 3

doses of IV alemtuzumab to each cycle. Cycles were repeated every 28

days for a total of six courses and all patients received pegylated filgrastim

6 mg SC with each course of therapy. Antibiotic prophylaxis with TMP/SMX

DS and valaciclovir or valganciclovir was also given to all patients. Sixty

patients were enrolled with a median age of 59. Median b2 microglobulin

was 5.1 mg/L (4–11.6) and median white count was 100,000 (5–665). Fifty-

one percent of the patients had Rai stage 3–4 disease. The median number

of courses was four and the main reason for not completing six cycles was

delayed recovery of counts and/or infection. In 14 patients with 17p deletion

the overall response rate was 78% and the CR rate was 57%, which was the

highest CR rate ever reported for such patients. Nonetheless the median time

to progression in these patients was a disappointing 18 months. When compar-

ing the results to a historical cohort of patients treated with FCR the incidence

of grade 3/4 neutropenia was the same at 31%, albeit patients receiving FCR

did not all receive growth factor support. Grade 3–4 thrombocytopenia was not

significantly different. Major infections were seen in 17% of patients receiving

CFAR versus 15% of the historical control patients. CMV reactivation occurred

in seven patients, all of whom were on valaciclovir prophylaxis.

Discussion

Although CFAR has significant activity there is not yet any convincing evi-

dence that this is more efficacious than FCR, whereas it appears to be sig-

nificantly more myelosuppressive given that even with the use of prophylac-

tic growth factor the grade 3–4 neutropenia rate was 31%. In addition CMV

is rarely seen with FCR so clearly it is related to the addition of alemtuzu-

mab to the FCR chemotherapy. Although the patients with deletion 17p fared

better than might be expected with other regimens their median time to pro-

gression was significantly shorter than all other patients and new agents are

still sorely needed for patients with 17p deletion.

The last abstracts will focus on investigational agents that are still rela-

tively early in clinical trials but appear highly promising.

7. Evidence of Clinical Activity in the Phase I Study of

CAL-101 an Oral P110g Isoform Selective Inhibitor of

Phosphatidylinositol3-Kinase in Patients with Relapsed

or B-Cell Malignancies. (#922)Abstract summary

The PI3K p110g isoform is primarily expressed in cells of hematopoetic

origin and plays an important role in normal B-cell maturation and function.

CAL-101 is an oral inhibitor of PI3KP110g with 40–300 fold selectivity com-

pared to other PI3K isoforms. This selectivity might provide a better thera-

peutic index relative to pan-PI3K inhibitors. In this Phase I trial sequential

cohorts of three patients with either relapsed CLL or NHL were enrolled to

determine the DLT. During subsequent cohort expansion approximately 12

patients with CLL, indolent NHL, aggressive NHL, and AML were to be

letters

208 American Journal of Hematology

Page 20: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

enrolled. CAL-101 was administered orally twice daily continuously for 28

days per cycle. Forty-three patients have been enrolled including 17 patients

with CLL. The median number of prior regimens was 5. Dose escalation

ranged from 50 to 350 mg b.i.d. In the cohort expansion 31 patients were

enrolled to either 200 mg (N 5 17) or 350 mg (N 5 14) b.i.d. DLTs were

observed in five patients with increases in LFTs, which resolved following

discontinuation of the drug. No grade 4 hematologic toxicity was seen; infec-

tions did occur in nine patients. Of 41 evaluable patients for response, 17

had CLL and six achieved a partial remission with seven other patients

showing >50% reduction in lymph nodes but concurrent increase in periph-

eral blood lymphocytes to >50% of baseline suggesting a compartment

shift. This lymphocytosis was maximal during the first two cycles and

decreased thereafter. This effect of dislodging CLL cells from the tissue

microenvironment suggests that using CAL-101 in treatment in combination

with other cytotoxic agents might be particularly attractive.

Discussion

This is an interesting new agent that has a mechanism of action totally dif-

ferent from other drugs used to treat CLL. The DLT was liver toxicity but it

was also shown that this was reversible and, in fact, some patients were

resumed on therapy at a lower dose without subsequent liver abnormalities.

Of importance is that the drug is not myelosuppressive. This is very attrac-

tive, particularly for combination trials, as most of the active chemotherapeu-

tic agents that we have for CLL cause myelosuppression with resulting infec-

tion and combining drugs that will not contribute further to myelosuppression

is always an attractive concept.

8. An Ongoing Phase I/IIA Study of ABT-263;

Pharmacokinetics, Safety, and Antitumor Activity in

Patients with Relapsed or Refractory CLL. (#883)Abstract summary

ABT-263 is an orally bioavailable BH3 mimetic, which inhibits multiple antia-

poptotic BCL-2 family proteins including BCL-w, BCL-2, and BCL-XL [12]. Dos-

ing in this study was daily initially, for 14 of every 21 days and then to a ameliolo-

rate the impact of thrombocytopenia due to BCL-XL inhibition induced platelet

apoptosis, a 100 mg lead-in dose for 7 days followed by continuous 21/21 day

dosing up to 300 mg/day was investigated. The study enrolled 29 patients, 15 on

the 14/21 day schedule and 14 on subsequent schedules. Platelet nadirs were

transient and usually occurred on day 3–5 followed by partial recovery due to

compensatory increased megakarypoiesis during the continued dosing. The

lead-in dosing did reduce the early platelet nadir. At doses ranging from 10 to

250 mg on the 14/21 day schedule circulating platelet counts dropped on cycle 1

by an average of 12 to 70% at the highest dose. On the 21/21 day schedule at

doses of 125–300 mg the drop was between 52 and 68%. Two patients with CLL

had radiologically confirmed PRs with reduction in lymph node bulk of 99% and

79% and three others have had unconfirmed nodal responses of 100%, 71%,

and 55%. The overall response rate was 33%. The median PFS had not been

reached but the median time on study is 9 months. Other than thrombocytopenia

the most common adverse event that was dose limiting was diarrhea. Nausea,

vomiting, and fatigue also occurred. The recommended phase II dose is 100 mg

for 7 days followed by 250 mg continuous dosing.

Discussion

The first BCL-2 inhibitor in clinical trials was Genasense, a BCL-2 anti-

sense. However, several BCL-2 family members are over expressed in CLL

and there are many redundant pathways, suggesting that inhibiting more

than one protein would be highly desirable [13–14]. There are several drugs

now in development that inhibit multiple BCL-2 family members. ABT-263 is

the first one that’s in development that’s orally available. The response rate

is impressive in a refractory population given that this is a phase I/II study.

One caution with this drug will be the thrombocytopenia that occurs, so that

most patients, as in this trial, will be required to have adequate platelets to

receive this agent. The likely optimal use of any of these BH3 mimetics is in

combination with chemotherapy as decreasing BCL-2 family members should

markedly enhance the response to chemotherapy. Some caution will have to

be used in designing the combinations because of the thrombocytopenia.

9. Phase I Study of RO5072759 (GA101) in

Relapsed/Refractory CLL. (#884)Abstract summary

RO5072759(GA101) is the first humanized and glycoengineered type 2

monoclonal anti CD20 antibody to enter clinical trials [15]. In preclinical studies

GA101 showed increased ADCC as well as direct apoptosis of cells as com-

pared to rituximab. In this phase I study a flat dose of 400–2000 mg was given

on days 1, 8, and 22 and then every 3 weeks for a total of nine infusions. Thir-

teen patients with CLL have been entered; nine of the 13 had a 17p deletion or

an 11q deletion and seven of 10 evaluable had an unmutated IGVH status. The

median number of prior regimens was 3; all patients had received prior fludara-

bine. GA101 was well tolerated with no dose limiting toxicity seen and no dose

reductions required. Similarly to rituximab, the most common side effect was

grade 1 or 2 infusion reactions, usually with the initial infusion. However, transi-

ent neutropenia was seen in nine patients. The overall response rate was 62%

(8/13) with one CRi and 7 PRs. There was no clear dose relationship estab-

lished. The responses were ongoing and ranged from 3.51 to 81months.

Discussion

Although the number of patients is small in this analysis, the response rate

is very impressive given the extent of prior therapy and the high risk character-

istics that these patients had. It is particularly impressive if one considers that

this is a drug with minimal side effects, although it would appear that neutro-

penia may be more prominent with this agent than is seen with rituximab.

10. A Phase I trial of TRU-016, an Anti CD37 Small Modular

Immunopharmaceutical (SMIT) Protein in Relapsed and

Refractory CLL: Early Promising Clinical Activity. (#3424)Abstract summary

CD37 is a tetraspan family member expressed predominantly on normal and

transformed B-cells across a wide range of maturational stages [16]. TRU-016 is a

humanized anti-CD37 SMIP protein. Preclinical studies demonstrated CD37 SMIP

protein mediated significantly greater direct killing of CLL cells than rituximab as

well as greater NK cell mediated killing of CLL cells as compared to either alemtu-

zumab or rituximab [17]. In this phase I study there were two dosing regimens

evaluated either weekly for 4 weeks or three times the first week and then once

weekly for 3 weeks. Patients were allowed to receive up to two additional cycles if

clinical benefit was seen with the first cycle. Thirty-three patients were treated;

70% of them had Rai stage 3–4 disease, 36% of them had a 17p deletion, 18%

had an 11q deletion, and 6% had both. Doses range from 0.03 mg/kg to 10 mg/

kg. No MTD was reached. Mild infusion related toxicity was observed and there

was biological activity beginning with the 0.03 mg/kg dose. Five patients have

achieved PR and most patients had significant reduction in lymphocyte counts.

The protocol has been amended to explore higher saturating doses of CD37.

Discussion

This is another new family of molecules, the SMIPs, and TRU-016 is the

first in development for CLL. Again, given that this is a phase I study, that

an MTD has not been reached, and that patients are extensively pretreated,

5 responses of 33 is quite significant. The lack of toxicity other than mild

infusion reactions is also very encouraging and as was noted for some of

the other new molecules the lack of myelosuppression should make this an

easy agent to develop in combination.

Conclusion

FCR has now been reported to be the most efficacious frontline regimen in

CLL. Not only are response rates high with this regimen and PFS prolonged,

but a survival advantage has been shown when compared to chemotherapy

alone. Another very promising regimen is bendamustine and rituximab, which

also provides high overall response rates. This regimen is being compared to

FCR in a frontline trial by the German CLL Study Group. Data with lenalido-

mide and rituximab was presented for relapsed patients and it would appear

that just as when rituximab is added to chemotherapy, adding rituximab to an

imid markedly improves response rates. A new monoclonal antibody targeting

CD20, ofatumumab, was recently approved by the FDA for fludarabine and

alemtuzumab refractory CLL. The combination of ofatumumab and fludarabine

and cyclophosphamide appears comparable to that of FCR but with short fol-

low-up. Several promising new agents are in clinical trials. Particularly encour-

aging is the fact that all of these drugs appear to have different mechanisms of

action and in most cases, minimal to no myelosuppression, which should ena-

ble them to be developed in combination going forward.

1Department of Leukemia, UT MD Anderson Cancer Center, Houston, Texas*Correspondence to: Susan O’Brien, UT MD Anderson Cancer Center

Department of Leukemia, 1515 Holcombe Blvd. Unit 428Houston, Texas 77030. E-mail: [email protected]

Published online 8 January 2010 in Wiley InterScience(www.interscience.wiley.com).

DOI: 10.1002/ajh.21640Conflict of interest: Nothing to report.

letters

American Journal of Hematology 209

Page 21: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

References

1. Rai KR, Peterson BL, Appelbaum FR, et al. Fludarabine compared with chlor-

ambucil as primary therapy for chronic lymphocytic leukemia. N Engl J Med

2000;343:1750–1757.

2. Leoni LM, Bailey B, Reifert J, et al. Bendamustine (Treanda) displays a distinct

pattern of cytotoxicity and unique mechanistic features compared with other

alkylating agents. Clin Cancer Res 2008;14:309–317.

3. Knauf WU, Lissichkov T, Aldaoud A, et al. Phase III randomized study of benda-

mustine compared with chlorambucil in previously untreated patients with

chronic lymphocytic leukemia. J Clin Oncol 2009;27:4378–4384.

4. Cheson BD, Rummel MJ. Bendamustine: Rebirth of an old drug. J Clin Oncol

2009;27:1492–1501.

5. Rummel MJ, Chow KU, Hoelzer D, et al: In vitro studies with bendamus-

tine: Enhanced activity in combination with rituximab. Semin Oncol 2002;

29:12–14.

6. Teeling JL, French RR, Cragg MS, et al. Characterization of new human CD20

monoclonal antibodies with potent cytolytic activity against non-Hodgkin lym-

phomas. Blood 2004;104:1793–1800.

7. Teeling JL, Mackus WJM,Wiegman LJJM, et al. The biological activity of human

CD20 monoclonal antibodies is linked to unique epitopes on CD20. J Immunol

2006;177:362–371.

8. Chanan-Khan A, Miller KC, Musial L, et al. Clinical efficacy of lenalidomide in

patients with relapsed or refractory chronic lymphocytic leukemia: results of a

phase II study. J Clin Oncol 2006;24:5343–5349.

9. Ferrajoli A, Lee B-N, Schlette EJ, et al. Lenalidomide induces complete and par-tial remissions in patients with relapsed and refractory chronic lymphocytic leu-kemia. Blood 2008;111:5291–5297.

10. Keating MJ, O’Brien S, Albitar M, et al. Early results of a chemoimmunotherapyregimen of fludarabine, cyclophosphamide, and rituximab as initial therapy forchronic lymphocytic leukemia. J Clin Oncol 2005;23:4079–4088.

11. Tam CS, O’Brien S, Wierda W, et al. Long-term results of the fludarabine, cyclo-phosphamide, and rituximab regimen as initial therapy of chronic lymphocyticleukemia. Blood 2008;112:975–980.

12. Certo M, Del Gaizo Moore V, Nishino M, et al. Mitochondria primed by death sig-nals determine cellular addiction to antiapoptotic BCL-2 family members. Can-cer Cell 2006;9:351–365.

13. Chen L, Willis SN, Wei A, et al. Differential targeting of prosurvival Bcl-2 proteinsby their BH3-only ligands allows complementary apoptotic function. Mol Cell2005;17:393–403.

14. Willis SN, Fletcher JI, Kaufmann T, et al. Apoptosis initiated when BH3 ligandsengage multiple Bcl-2 homologs not bax or bak. Science 2007;315:856–859.

15. Robak T. GA-101, a third-generation, humanized and glycol-engineered anti-CD20 mAb for the treatment of B-cell lymphoid malignancies. Curr Opin InvestigDrugs 2009;10:588–596.

16. Knobeloch KP, Wright MD, Ochsenbein AF, et al. Targeted inactivation of the tet-raspanin CD37 impairs T-cell-dependent B-cell response under suboptimal cos-timulatory conditions. Mol Cell Biol 2000;20:5363–5369.

17. Zhao X, Lapalombella R, Joshi T, et al. Targeting CD27-positive lymphoid malig-nancies with a novel engineered small modular immunopharmaceutical. Blood2007;110:2569–2577.

ASH 2009 meeting report—Top 10 clinically oriented abstractsin multiple myeloma

Shaji Kumar* and S. Vincent Rajkumar

Treatment of multiple myeloma continues to evolve rapidly as a result of

clinical trials exploring the optimum approach for utilizing IMiDs and

bortezomib as well as clinical trials exploring new agents for this dis-

ease. The annual ASH meeting at New Orleans showcased over 800

abstracts related to various aspects of multiple myeloma, highlighting

the significant amount of research effort devoted to improving outcome

of patients with myeloma, still an incurable disease despite the recent

advances. We have selected ten most clinically relevant abstracts from

among those presented at ASH, while acknowledging the important

contribution of the other studies towards advancing the field.

Induction with Velcade(R)/Dexamethasone Partially

Overcomes the Poor Prognosis of t(4;14), but Not

That of Del(17 p), in Young Patients with Multiple

Myeloma (Abstract 957)

One of the most important advances in myeloma in the past decade has

been the appreciation of the genetic heterogeneity in multiple myeloma and

the use of genetic abnormalities to risk stratify patients. While the prognostic

impact of these agents were examined and validated in the setting of alkylator

based therapies and stem cell transplant, a better understanding of the differ-

ential impact of the new agents in patients with different abnormalities has led

to early attempts at defining therapy based on prospective risk stratification

[1]. Previous studies had suggested bortezomib can improve the outcome in

patients with high risk genetic abnormalities especially deletion 13, and t

(4;14). However, most of the available studies did not specifically look at the

impact of del (17 p) in patients treated with bortezomib, an abnormality that

has been associated with poor outcome in the context of most of the treat-

ments currently used including stem cell transplantation [2]. Avet Loiseau et al.

examined the impact of various cytogenetic abnormalities among patients

enrolled in the IFM 2005-01 clinical trial that randomized patients to either Bor-

tezomib dexamethasone or VAD (vincristine/doxorubicin/dexamethasone) as

induction prior to autologous stem cell transplant (SCT) [3]. While the pres-

ence of t (4;14) retained prognostic value among patients receiving bortezo-

mib, the event free (EFS) and overall (OS) survival of patients with t (4;14)

was significantly better for those treated with bortezomib (2.3 vs. 1.4 years

and NR vs. 2.9 years, respectively) compared to VAD. However, among those

patients with del (17 p) abnormality, no beneficial effect was seen with the use

of bortezomib compared to VAD, either in terms of EFS or OS. These results

taken in the context of other studies examining the impact of del (17 p) with

lenalidomide [4,5] highlights the poor outcome of patients with this abnormality

with the currently available therapies and the need to consider these patients

for clinical trials evaluating novel combinations or new therapies for myeloma.

Multicenter, Randomized, Open-Label, Phase III Trial of

Lenalidomide-Dexamethasone (Len/dex) Versus Therapeutic

Abstention in Smoldering Multiple Myeloma at High Risk of

Progression to Symptomatic MM: Results of the First Interim

Analysis (Abstract 614)

The conventional approach to management of multiple myeloma has been

to initiate therapy at the onset of CRAB features or any other symptomatic

manifestation of the underlying gammopathy. In contrast to monoclonal

gammopathy of undetermined significance (MGUS) where the risk of pro-

gression is about 1% every year, smoldering multiple myeloma (SMM) has a

high risk of progression to symptomatic myeloma (estimated at 50% during

the first 5 years). The availability of well tolerated and effective drugs as well

as the ability to identify patients with SMM at a high risk of progression has

led to interest in exploring early intervention approaches for these patients.

Mateos and colleagues randomized patients with high risk SMM (defined as

bone marrow plasmacytosis >10% and a serum M pike >3 gm/dl OR either

of these features along with evidence of immunoparesis) to observation, the

current standard of care, or lenalidomide/dexamethasone therapy [6]. Patients

in the experimental arm received nine 4-week cycles of lenalidomide (25 mg/

day for 21/28 days) and dexamethasone (20 mg days 1–4 and 12–15) followed

by lenalidomide 10 mg daily for 3 weeks every 2 months. After a median of four

cycles of therapy, a VGPR or better was seen in 30% of patients (n 5 40)

receiving active treatment. More importantly, 16 patients in the observation

arm had disease progression (median TTP 5 19.3 months) compared to two

patients in the treatment arm, both of who had stopped treatment early.

Clearly, it is an important study in terms of evaluating approaches to early

intervention in patients with myeloma. However, it should be viewed as an

early versus delayed approach to myeloma therapy given that the treatment

arm is essentially receiving therapy currently used for symptomatic patients.

letters

210 American Journal of Hematology

Page 22: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

Improvement in TTP would be expected given the known activity of lenalido-

mide in newly diagnosed myeloma and the appropriate endpoint should be an

improvement in the overall survival or at the minimum an improvement in the

quality of life related to myeloma related symptoms/complications in the face

of comparable survival. The results of this trial while very intriguing, should not

be interpreted as changing the current standard of care, which is careful and

close observation for evidence of progression.

Novel Three- and Four-Drug Combinations of Bortezomib,

Dexamethasone, Cyclophosphamide, and Lenalidomide, for

Newly Diagnosed Multiple Myeloma: Encouraging Results

from the Multi-Center, Randomized, Phase 2 EVOLUTION

Study (Abstract 127)

The initial success with the novel agents spurred interest in maximizing

the potential therapeutic benefit by judicious combination of the novel agents

and conventional agents has been done successfully for other malignancies

like lymphoma and childhood leukemia. The combination of bortezomib and

dexamethasone with either lenalidomide or alkylators have resulted in high

response rates in previously untreated MM. The EVOLUTION trial was

designed to explore the possibility of combining four of the most effective

medications to develop an active regimen with durable response [7]. Follow-

ing an initial Phase 1 portion that determined the dose of cyclophosphamide

(C) that can be safely combined with bortezomib (V), lenalidomide (R), and

dexamethasone (D), the Phase 2 portion randomized newly diagnosed

patients to receive up to eight 21-d cycles of VDR (V 1.3 mg/m2 d 1, 4, 8,

11; Dex 40 mg d 1, 8, 15; Rev 25 mg d 1–14) or VDC (VD as in VDR, plus

Cy 500 mg/m2 d 1, 8) or VDCR (VDC plus Rev 15 mg d 1–14) as induction

therapy. This was followed by V 1.3 mg/m2 (d 1, 8, 15, 22) for four 42-d

maintenance cycles in all treatment arms. The VDC arm was modified to

add cyclophosphamide on day 15 following the interim analysis (mod-VDC

arm). Patients were allowed to pursue SCT after four cycles. The overall

response rates (93, 93, 91, and 93%) and VGPR rates (59, 55, 47, and

60%), respectively for the VDCR, VDR, VDC, and mod-VDC arms were

comparable. The regimens were well tolerated with hematological toxicities

being the most common and neutropenia more common with the regimens

containing cyclophosphamide. Peripheral neuropathy rates were comparable

to other bortezomib studies. In essence, the trial demonstrates comparable

response rates when bortezomib is combined with lenalidomide or cyclo-

phosphamide or both, and highlights the need for clinical trials comparing

combination regimens against sequential use of the new drugs. In particular,

the mod-VDC arm had significant activity and compares with the data

obtained from other trials that evaluated this combination, albeit with differ-

ent doses and schedules. In this meeting, Einsele et al. reported on 400 pts

with untreated MM receiving three 3-week cycles of induction treatment with

V 1.3 mg/m2 iv d 1, 4, 8, 11; D 40 mg/d orally d 1, 2, 4, 5, 8, 9, 11, 12; and

C 900 mg/m2 iv d 1 before scheduled high dose melphalan and SCT. The

overall response was over 80% in an ITT analysis [8]. Also at this meeting

Reeder et al. presented results from a Phase 2 trial that used weekly borte-

zomib and obtained comparable response rates as twice weekly bortezomib,

but with less toxicity [9].

High Complete and Very Good Partial Response Rates with

Bortezomib–Dexamethasone as Induction Prior to ASCT in

Newly Diagnosed Patients with High-Risk Myeloma: Results

of the IFM2005-01 Phase 3 Trial (Abstract 353)

Given the important role of SCT in management of myeloma in transplant

eligible patients, developing effective regimens for initial therapy prior to SCT

has been an area of intense investigation with numerous Phase 3 trials in the

past decade. Incorporation of the new agents in the upfront treatment of mye-

loma has been the common theme for these trials and mature data from these

earlier trials are now becoming available. The IFM-2005-01 trial randomized

patients with newly diagnosed MM to receive 4 3 4-week cycles of VAD with-

out (n 5 121) or with (n 5 121) 2 3 4-week cycles of DCEP consolidation, OR

4 3 3-week cycles of VD without (n 5 121) or with (n 5 119) DCEP [10].

Patients achieving <VGPR post-first HDT-ASCT could receive a second autol-

ogous or an allogeneic SCT. The proportion of patients obtaining a VGPR or

better was higher with VD compared to VAD (68% vs. 47%) at all stages and at

the end of all planned treatment including a second SCT for those with <VGPR

after first SCT. As expected, more patients in the VAD arm received a tandem

SCT (54% vs. 38%) and the PFS was marginally improved with VD (36 vs. 30

months; P 5 0.06). However, the PFS was significantly better with VD among

ISS stage 2–3 patients. In addition, a significant PFS benefit was seen with

attainment of VGPR or better among patients with ISS 2–3 and in those with

high risk cytogenetics underscoring the fact that only subgroups of patients

may benefit from deep responses, such as high risk patients. In another Phase

3 study presented by Cavo et al., patients were randomized to three 21-d

cycles of either bortezomib-thalidomide-dexamethasone (VTD) (V, 1.3 mg/m2

twice-weekly; T, 200 mg/d through d 1–63; D, 320 mg/cycle) or thalidomide-

dexamethasone (TD) (same dose and schedule as VTD) as induction therapy

in preparation for ASCT [11]. Two 35-d cycles of either VTD or TD were given

as consolidation therapy following ASCT (V, 1.3 mg/m2 once-weekly; T, 100

mg/d through d 1–70; D, 320 mg/cycle). VGPR or better after SCT was more

likely with VTD (88% vs. 72%) compared to TD, translating into better PFS (76

vs. 58% at 30 months) for VTD patients. While these trials demonstrate high

response rates for the new regimens in the context of SCTand improved PFS,

OS data is still lacking.

A Prospective, Multicenter, Randomized, Trial of Bortezomib/

Melphalan/Prednisone (VMP) Versus Bortezomib/Thalidomide/

Prednisone (VTP) as Induction Therapy Followed by

Maintenance Treatment with Bortezomib/Thalidomide (VT)

Versus Bortezomib/Prednisone (VP) in Elderly Untreated

Patients with Multiple Myeloma Older Than 65 Years (Abstract 3)

Patients over 65 years at diagnosis, a group often considered ineligible for

SCT and had seen little progress since introduction of MP, have been the

focus of several large studies evaluating the role of novel agents combined

with MP. The VISTA trial demonstrated a survival advantage for these

patients with addition of bortezomib to MP. However, this progress came at

the coast of increased toxicity in this frail population, especially peripheral

neuropathy related to bortezomib. The Spanish myeloma group presented

the results of their Phase 3 trial that asked the question whether efficacy

can be maintained while reducing toxicity with less intensive Bortezomib

dosing [12]. Patients were randomized to receive six cycles of VMP versus

VTP as induction therapy followed by maintenance with VT versus VP for

up to 3 years. In the VMP arm bortezomib was given at 1.3 mg/m2 twice

weekly (days 1, 4, 8, and 11; 22, 25, 29, and 32) for one 6-week cycle, followed

by once weekly (days 1, 8, 15, and 22) for five 5-week cycles in combination

with oral melphalan 9 mg/m2 and prednisone 60 mg/m2 once daily on days 1–

4 of each cycle. In the VTP arm bortezomib and prednisone were same, but

instead of melphalan they received thalidomide at a dose of 100 mg daily. Fol-

lowing the six cycles of induction, maintenance consisted of bortezomib, 1.3

mg/m2 twice weekly (days 1, 4, 8, and 11) administered every three months in

combination with either continuous thalidomide, 50 mg daily (VT) or predni-

sone, 50 mg on alternate days (VP) up to 3 years. The overall response rates

to induction were similar for the two regimens (VMP vs. VTP) and the propor-

tion of patients improving their response to CR with maintenance was similar

for the two regimens evaluated (VT vs. VP). However, the serious adverse

events and proportion of patients discontinuing therapy for SAE was higher

with VTP compared to VMP. When the survival outcomes were compared

among the four groups (four combinations of induction/maintenance regi-

mens), the PFS was significantly better for VMP followed by VT compared to

VTP followed by VP. The OS was similar across all groups. This study serves

to highlight the low toxicity of weekly bortezomib, without compromising on the

efficacy and confirms the utility of combining bortezomib with an alkylating

agent as well as that of prolonged treatment at lower doses that are better tol-

erated. The overall survival at 2 years was 85% for the entire group, underscor-

ing the improvements that have been made in this patient population. In addi-

tion, the study also demonstrated a superior PFS for patients obtaining a mini-

mal residual disease negative status by flow cytometry of marrow.

Bortezomib, Melphalan, Prednisone, and Thalidomide (VMPT)

Followed by Maintenance with Bortezomib and Thalidomide

for Initial Treatment of Elderly Multiple Myeloma Patients

(Abstract 128)

Several Phase 3 trials have compared MP with and without thalidomide,

and a metanalysis of these trials presented at this meeting showed a sig-

nificant improvement in the PFS and possibly the OS with addition of tha-

lidomide to MP [13]. The Italian myeloma group randomized patients over

65 years to VMPT followed by maintenance with bortezomib and thalido-

mide or VMP. Patients were treated with nine 6-week cycles of VMPT

letters

American Journal of Hematology 211

Page 23: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

(Induction: V 1.3 mg/m2 days 1, 4, 8, 11, 22, 25, 29, and 32 in cycles

1–4 and days 1, 8, 22, and 29 in cycles 5–9; M 9 mg/m2 days 1–4; P

60 mg/m2 days 1–4 and T 50 mg days 1–42; Maintenance: V 1.3 mg/m2

every 15 days and T 50 mg/day) or VMP (Same doses and schedules

as in VMPT, but without maintenance) [14]. The protocol was subse-

quently amended with both VMPT and VMP induction schedules changed

to nine 5-week cycles and bortezomib schedule modified to weekly

administration (1.3 mg/m2 days 1, 8, 15, and 22). VMPT followed by VT

maintenance resulted in higher response rates (VGPR or better of 59%

vs. 50%) translating to an improved PFS of 60% vs. 42% at 3 years. OS

still remains comparable between the arms. While the rate of any grade

and grade 3/4 peripheral neuropathy on VMP with twice weekly bortezo-

mib was similar to that of VISTA trial (43% and 14%, respectively), the

rates were only 21% and 2% for the weekly schedule. This reduction in

the dose of bortezomib had no impact on the PFS when the two sched-

ules were compared and most importantly, the actual delivered cumulative

dose of bortezomib was identical between the two schedules because of

more frequent dose reductions in the twice weekly schedule. These

results make a convincing case for adoption of weekly bortezomib, espe-

cially in combination with alkylating agents. These results also are in con-

cordance with the findings of the Mayo study of bortezomib in combina-

tion with cyclophosphamide.

A Phase III Study to Determine the Efficacy and Safety of

Lenalidomide in Combination with Melphalan and Prednisone

(MPR) in Elderly Patients with Newly Diagnosed Multiple

Myeloma (Abstract 613)

Combination regimens combining thalidomide or bortezomib with MP

are superior to MP alone and have become the standard of care for non-

transplant eligible patients. Phase 2 studies have suggested excellent

activity for lenalidomide combined with MP (MPR). Palumbo et al.

randomized non-transplant eligible patients to receive MPR followed by

lenalidomide maintenance (MPR-R) or MPR followed by placebo (MPR)

or MP followed by placebo (MP) [15]. The doses were: M-0.18 mg/kg,

days 1–4, P- 2 mg/kg, days 1–4 and R- 10 mg/day PO, days 1–21 of

28 day cycles. Following nine cycles, patients in the MPR-R arm started

maintenance with R 10 mg/day, days 1–21/28. The overall response

rates were higher for the MPR-R compared to MP (77 vs. 49%) as was

VGPR or better (32 vs. 11%). The median PFS was 13 months for MP

while not reached for MPR-R. Interestingly, the PFS was identical

between MP and MPR without maintenance, even though the response

rates appeared to be higher with the MPR suggesting that the effect on

PFS is entirely explained by the use of maintenance lenalidomide.

These results, especially the lack of even a PFS advantage for MPR

compared to MP, highlights the need for more mature data and demon-

stration of overall survival improvements before accepting this combina-

tion as standard of care.

Pomalidomide (CC4047) Plus Low Dose Dexamethasone (Pom/

dex) Is Active and Well Tolerated in Lenalidomide Refractory

Multiple Myeloma (MM) (Abstract 429)

Myeloma remains incurable with the currently available options and the

search continues for better drugs for continued control of relapsed dis-

ease. Pomalidomide, a new IMiD was shown to have significant activity in

relapsed disease with response rates of 50–60% including responses in

lenalidomide refractory patients in a Phase 2 study from Mayo Clinic [16].

This follow-up Phase 2 study targeted patients refractory to lenalidomide,

defined as progression on lenalidomide or within 60 days of stopping

lenalidomide [17]. Thirty four patients with relapsed disease, with median

of four prior therapies were enrolled, including eight patients with high

risk cytogenetic features. Patients received pomalidomide: 2 mg p.o. daily

days 1–28 and dexamethasone: 40 mg p.o. days 1, 8, 15, and 22. After

two cycles, pomalidomide dose could be increased to 4 mg/day if no

response was seen or if progression. The regimen was well tolerated with

an overall response rate (PR or better) of 32%. An additional 18% had

minimal response. One patient achived a PR after dose was increased to

4 mg. At a median follow up of 6 months, 16 patients (41%) remained

progression free and 6 patients (18%) had died. In another presentation

at the meeting, Richardson et al. reported on a Phase 1 study that

sought to determine the maximum tolerated dose of pomalidomide alone

or with dexamethasone on a 21/28 day schedule [18]. Pomalidomide was

given QD on Days 1–21 of 28-day cycle: four dose levels (2, 3, 4, 5 mg)

were studied with option to add dex at 40 mg/wk after four cycles for

lack of response or progressive disease. The MTD was determined to be

4 mg given for 21/28 days. PR or better was seen in 28% and an MR

was seen in another 24%. Dex was added in 15 patients with 53% hav-

ing an improved response. Clearly, these trials demonstrate the activity of

this drug in relapsed myeloma and in particular, the activity in the lenali-

domide refractory population demonstrate lack of cross resistance

between the agents.

Updated Results of Bortezomib-Naive Patients in PX-171-004,

An Ongoing Open-Label, Phase II Study of

Single-Agent Carfilzomib (CFZ) in Patients with Relapsed

or Refractory Myeloma (MM)/PX-171-004, An Ongoing

Open-Label, Phase II Study of Single-Agent Carfilzomib (CFZ)

in Patients with Relapsed or Refractory Myeloma (MM);

Updated Results From the Bortezomib-Treated

Cohort (Abstract 302/303)

The efficacy of bortezomib in the treatment of myeloma has clearly dem-

onstrated the role of proteasome inhibitors in this disease. Several new pro-

teasome inhibitors are being evaluated in clinical trials and carfilzomib

appear to be very promising based on results so far. Results of a Phase 2

trial that evaluated the efficacy of carfilzomib in patients with relapsed dis-

ease was presented at the meeting, one in a group of patients who were

bortezomib naıve and another that had previous treatment with bortezomib

[19,20]. CFZ 20 mg/m2 IV was administered on Days 1, 2, 8, 9, 15, and 16

every 28 days, for up to 12 cycles. A partial response or better was seen

among 46% of 54 patients in the bortezomib cohort and an MR was seen in

another 15%. In contrast, among the 35 patients previously treated with bor-

tezomib, PR or better was seen in 18% of patients and another 12% had

MR. The median TTP was 7.6 months in the BTZ naıve group and 5.4

months in the BTZ treated group. The most common adverse events

included fatigue, GI symptoms and infections in both group of patients.

Overall the treatment was well tolerated with over a third of the patients

receiving more than nine cycles. Grade 3/4 neuropathy was relatively

uncommon as was discontinuations due to neuropathy, presented in more

detail in an associated abstract [21]. Clearly, this drug has promising activity

and ongoing trials are geared to move it forward into the clinic.

Phase 1/2 Study of Elotuzumab in Combination with

Lenalidomide and Low Dose Dexamethasone in Relapsed or

Refractory Multiple Myeloma: Interim Results (Abstract 432)

Monoclonal antibodies have had significant success in the treatment of

malignacies as demonstrated by the success of rituximab in lymphoma. How-

ever, this class of drugs have not been very successful in myeloma due to the

phenotypic heterogeneity seen in MM. Elotuzumab is a humanized monoclo-

nal IgG1 antibody directed against CS1, a cell surface glycoprotein, which is

highly and uniformly expressed in MM. It induces significant antibody-depend-

ant cytotoxicity against myeloma cells in vitro. This Phase 1 study was done

to evaluate the maximum tolerated dose of elotuzumab in combination with

lenalidomide and low dose dexamethasone in patients with relapsed MM [22].

Lenalidomide was dosed at 25 mg on days 1–21 of a 28-day cycle with dex 40

mg weekly. Elotuzumab in three escalating dose cohorts (5, 10, and 20 mg/

kg) is administered by IV infusion on Days 1, 8, 15, and 22 of the 28-day cycle

in the first two cycles and then on Days 1 and 15 of each subsequent cycle.

Twenty eight patients were enrolled including 22 with no prior exposure to

lenalidomide. A PR or better was seen in 82% of patients, and was 92%

among the 22 lenalidomide naıve patients. This included 18 and 23% patients

with VGPR respectively. At 4.5 months of follow-up, the median TTP was not

reached. No dose limiting toxicity was seen up to the highest dose tested. The

most common adverse events included fatigue, neutropenia, and GI symp-

toms. These results are very encouraging and the ongoing and planned stud-

ies will provide a better estimate of its efficacy.

Conclusions

Clinical trials are paving the way for continued progress in the treatment

of myeloma and we are making progress in understanding the most effective

ways to use the available treatments. Incorporation of multiple effective

letters

212 American Journal of Hematology

Page 24: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

drugs in combination regimens improves the response rates and depth of

responses leading to improved duration of disease control. However, long-

term follow up is needed to assess the impact of combination approaches

on the natural history of disease compared to a more sequential approach.

This remains one of the most important questions today as one search for

the optimum balance between efficacy and toxicity. Encouraging improve-

ment has been seen in the outcome of older patients with myeloma with the

new drugs and clearly these patients should be treated initially with an alky-

lator-novel drug combination such as MPT or MPV. However, it is sobering

to see the continued poor outcome among patients with del 17 p abnormal-

ities, highlighting an area of need. Finally, the effective drugs currently going

through clinical trials will continue to redefine the treatment and outcome of

myeloma patients in the near future.

Division of Hematology, Mayo Clinic, Rochester, Minnesota*Correspondence to: Shaji Kumar, Division of HematologyDepartment of Medicine, Mayo Clinic, 200 First street SW

Rochester, MN 55905E-mail: [email protected]

Published online 8 January 2010 in Wiley InterScience(www.interscience.wiley.com).

DOI: 10.1002/ajh.21637Conflict of interest: Research funding from Celgene, Millennium

References

1. Kumar SK, Mikhael JR, Buadi FK, et al. Management of newly diagnosed sympto-matic multiple myeloma: Updated mayo stratification of myeloma and risk-adaptedtherapy (mSMART) consensus guidelines. Mayo Clin Proc 2009;84:1095–1110.

2. Gertz MA, Lacy MQ, Dispenzieri A, et al. Clinical implications of t(11;14)(q13;q32), t(4;14)(p16.3;q32), and 217p13 in myeloma patients treated withhigh-dose therapy. Blood 2005;106:2837–2840.

3. Avet Loiseau H, Moreau P, Mathiot C, et al. Induction with Velcade(R)/Dexame-thasone partially overcomes the poor prognosis of t(4;14), but not that ofDel(17p), in young patients with multiple myeloma. ASH Annu Meet Abstr2009;114:957.

4. Reece D, Song KW, Fu T, et al. Influence of cytogenetics in patients withrelapsed or refractory multiple myeloma treated with lenalidomide plus dexame-thasone: Adverse effect of deletion 17p13. Blood 2009;114:522–525.

5. Kapoor P, Kumar S, Fonseca R, et al. Impact of risk stratification on outcomeamong patients with multiple myeloma receiving initial therapy with lenalidomideand dexamethasone. Blood 2009;114:518–521.

6. Mateos M-V, Lopez-Corral L, Hernandez MT, et al. Multicenter, randomized,open-label, Phase III trial of Lenalidomide-Dexamethasone (Len/dex) Vs thera-peutic abstention in smoldering multiple myeloma at high risk of progression tosymptomatic MM: Results of the first interim analysis. ASH Annu Meet Abstr2009;114:614.

7. Kumar S, Flinn IW, Hari PN, et al. Novel three- and four-drug combinations ofbortezomib, dexamethasone, cyclophosphamide, and lenalidomide, fornewly diagnosed multiple myeloma: Encouraging results from the multi-cen-ter, randomized, Phase 2 EVOLUTION Study. ASH Annu Meet Abstr 2009;114:127.

8. Einsele H, Liebisch P, Langer C, et al. Intravenous cyclophosphamide and dexa-methasone (VCD) induction for previously untreated multiple myeloma (GermanDSMM XIa Trial). ASH Annu Meet Abstr 2009;114:131.

9. Reeder CB, Reece DE, Kukreti V, et al. A Phase II trial comparison of once ver-sus twice weekly bortezomib in CYBORD chemotherapy for newly diagnosedmyeloma: Identical high response rates and less toxicity. ASH Annu Meet Abstr2009;114:616.

10. Harousseau J-L, Avet-Loiseau H, Attal M, et al. High complete and very goodpartial response rates with bortezomib–dexamethasone as induction prior toASCT in newly diagnosed patients with high-risk myeloma: Results of theIFM2005–01 Phase 3 trial. ASH Annu Meet Abstr 2009;114:353.

11. Cavo M, Tacchetti P, Patriarca F, et al. A Phase III study of double auto-transplantation incorporating bortezomib-thalidomide-dexamethasone (VTD) orthalidomide-dexamethasone (TD) for multiple myeloma: Superior clinical out-comes with VTD compared to TD. ASH Annu Meet Abstr 2009;114:351.

12. Mateos M-V, Oriol A, Martinez J, et al. A prospective, multicenter, randomized,trial of bortezomib/melphalan/prednisone (VMP) versus bortezomib/thalido-mide/prednisone (VTP) as induction therapy followed by maintenance treatmentwith bortezomib/thalidomide (VT) versus bortezomib/prednisone (VP) in elderlyuntreated patients with multiple myeloma older than 65 years. ASH Annu MeetAbstr 2009;114:3.

13. Kapoor P, Rajkumar SV, Dispenzieri A, et al. Melphalan and prednisone (MP)versus melphalan, prednisone and thalidomide (MPT) as initial therapy for previ-ously untreated elderly and/or transplant ineligible patients with multiple mye-loma: A meta-analysis of randomized controlled trials. ASH Annu Meet Abstr2009;114:615.

14. Palumbo A, Bringhen S, Rossi D, et al. Bortezomib, melphalan, prednisone andthalidomide (VMPT) followed by maintenance with bortezomib and thalidomidefor initial treatment of elderly multiple myeloma patients. ASH Annu Meet Abstr2009;114:128.

15. Palumbo A, Cavallo F, Yehuda DB, et al. A prospective, randomized study ofmelphalan, prednisone, lenalidomide (MPR) versus melphalan (200 Mg/M2)and autologous transplantation (Mel200) in newly diagnosed myeloma patients:An interim analysis. ASH Annu Meet Abstr 2009;114:350.

16. Lacy MQ, Hayman SR, Gertz MA, et al. Pomalidomide (CC4047) plus low-dosedexamethasone as therapy for relapsed multiple myeloma. J Clin Oncol2009;27:5008–5014.

17. Lacy MQ, Gertz MA, Hayman SR, et al. Pomalidomide (CC4047) plus low dosedexamethasone (Pom/dex) is active and well tolerated in lenalidomide refractorymultiple myeloma (MM). ASH Annu Meet Abstr 2009;114:429.

18. Richardson P, Siegel D, Baz R, et al. A Phase 1/2 multi-center, randomized,open label dose escalation study to determine the maximum tolerated dose,safety, and efficacy of pomalidomide alone or in combination with low-dose dex-amethasone in patients with relapsed and refractory multiple myeloma whohave received prior treatment that includes lenalidomide and bortezomib. ASHAnnu Meet Abstr 2009;114:301.

19. Siegel D, Wang L, Orlowski RZ, et al. The multiple myeloma research consor-tium PX-171–004, an ongoing open-label, Phase II study of single-agent carfil-zomib (CFZ) in patients with relapsed or refractory myeloma (MM); updatedresults from the bortezomib-treated cohort. ASH Annu Meet Abstr 2009;114:303.

20. Wang L, Siegel D, Kaufman JL, et al. The multiple myeloma research consor-tium updated results of bortezomib-naive patients in PX-171–004, an ongoingopen-label, Phase II study of single-agent carfilzomib (CFZ) in patients withrelapsed or refractory myeloma (MM). ASH Annu Meet Abstr 2009;114:302.

21. Vij R, Wang L, Orlowski RZ, et al. The multiple myeloma research consortiumcarfilzomib (CFZ), a novel proteasome inhibitor for relapsed or refractory multi-ple myeloma, is associated with minimal peripheral neuropathic effects. ASHAnnu Meet Abstr 2009;114:430.

22. Lonial S, Vij R, Harousseau J-L, et al. Phase 1/2 study of elotuzumab in combi-nation with lenalidomide and low dose dexamethasone in relapsed or refractorymultiple myeloma: Interim results. ASH Annu Meet Abstr 2009;114:432.

Genetic polymorphisms in cytochrome P450s, GSTs, NATs,alcohol consumption and risk of non-Hodgkin lymphoma

Yonghong Li,1,2 Tongzhang Zheng,2 Briseis A. Kilfoy,3 Qing Lan,3 Theodore Holford,2 Xuesong Han,2

Ping Zhao,4 Min Dai,4 Brian Leaderer,2 Nat Rothman,3 and Yawei Zhang2*

The aim of this study was to investigate whether genetic polymorphisms

in cytochrome P450s (CYPs), glutathione S-transferases (GSTs), and N-

acetyltransferases (NATs) genes modify the relationship between alcohol

consumption and risk of non-Hodgkin’s lymphoma (NHL) in a population-

based, case-control study including 1,115 Connecticut women. Although

we did not find strong evidence that the genetic polymorphisms modify

the relationship between alcohol consumption and risk of NHL, we identi-

fied significant interactions for multiple GSTs and NATs and alcohol intake

among persons with DLBCL. Our results confer support investigation of

the gene–environment interaction in a larger study population of DLBCL.

Numerous epidemiologic studies have investigated the effect of alcohol

consumption on non-Hodgkin lymphoma (NHL) risk but the results have

been inconsistent [1–12]. Alcohol is a risk factor of interest as it is oxidized

to acetaldehyde, which is genotoxic, and it is also thought to induce the

activity of metabolic enzymes involved in its excretion and the detoxification

of potentially harmful xenobiotic compounds. Polymorphisms in genes that

code various types of cytochrome P450s (CYPs), glutathione S-transferases

(GSTs), and N-acetyltransferases (NATs) manifest as decreased or lack of

enzyme activity [2], prompting the hypothesis that allelic variants may be

associated with an impaired detoxification capacity and subsequently an

letters

American Journal of Hematology 213

Page 25: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

overall increased susceptibility to cancer. Although a limited number of stud-

ies have explored the relationship between genetic polymorphisms in CYPs,

GSTs and NATs and NHL risk [13–18], no study has been conducted to

investigate whether genetic variation in xenobiotic metabolic genes modifies

the association between alcohol consumption and NHL risk. We subse-

quently analyzed data from a population-based, case-control study in Con-

necticut women to explore the relationship between genetic polymorphisms

in CYPs, GSTs, and NATs genes, alcohol consumption, and risk of NHL.

Methods

The study population has been reported in detail in the previously pub-

lished papers [6,19]. Briefly, a total of 835 histologically confirmed incident

female NHL cases from 1996 to 2000 in Connecticut that were aged 21–84-

years old, still alive at the time of interview and had never been diagnosed of

cancer except for nonmelanoma skin cancer, were identified by the Yale Can-

cer Center’s Rapid Case Ascertainment Shared Resource. Among these eligi-

ble cases, 601 completed in-person interviews and 461 cases provided blood

samples. All cases were histologically confirmed and classified into NHL sub-

types according to the WHO classification system [20]. Population-based con-

trols were frequency matched on age (±5-year age group) and were identified

via random-digit dialing (for those aged less than 65 years) or random selec-

tion from Centers for Medicare and Medicaid Services (CMS) records (for

those aged 65 years or older). Of 1,264 eligible controls, 718 women com-

pleted in-person interviews and 535 women provided blood samples.

DNA was extracted from blood samples using phenol-chloroform extrac-

tion. A total of 20 single nucleotide polymorphisms (SNPs) in 11 xenobiotic

genes, including CYP21A2 (rs6474), CYP1A1 (rs1048943), CYP1A2

(rs762551), CYP1B1 (rs1056836), CYP2C9 (rs1799853), CYP2E1 (rs2070673),

GSTM3 (rs1799735), GSTP1 (rs1695, rs1138272), GSTT1 (rs17856199), NAT1

(rs4987076, rs13249533, rs1057126, and rs15561), and NAT2 (rs1041983,

rs1801280, rs1799929, rs1799930, rs1208, and rs1799931), were selected and

genotyped in blood-based DNA samples. Genotyping was conducted using real-

time PCR on an ABI 7900HT sequence detection system as described on the web-

site (http://snp500cancer.nci.nih.gov) at the NCI Core Genotyping Facility [21].

Duplicate samples from 100 study subjects and 40 replicate samples from each of

two blood donors were interspersed throughout the plates used for genotype analy-

sis. The concordance rates for quality control samples were 100% for all assays.

All genotyping frequencies among control populations were in Hardy-Weinberg

equilibrium (P > 0.05). The SNP data were used to assign the most likely NAT1

and NAT2 alleles and NAT2 acetylation phenotypes at the University of Louisville

(by Hein) [22].

Unconditional logistic regression models were used to estimate the odds

ratio (OR), the 95% confidence interval (CI) and P value for associations

between drinking of any beverage, wine, beer, or liquor, polymorphisms in

CYPs, GSTs and NATs, and risk of NHL, adjusting for age (<50, 50–70,

>70 years), education (high school or less, some college, college graduate,

or more), smoke (nonsmoker, 1–7, 8–14,15–33, �34 pack-years), and family

history of cancer (any cancer, none). All tests were two-sided with signifi-

cance level of 0.05 using SAS 9.1 (SAS Institute, Cary, NC). We also con-

ducted the same analysis for non-Hispanic Caucasians only.

No significant association was identified between genetic polymorphisms

in the xenobiotic genes of interest and risk of NHL overall among either non-

drinkers or drinkers of any type of alcoholic beverage or for drinkers of spe-

cific alcoholic beverages including wine, beer, or liquor (Supporting Informa-

tion Table S1).

When we looked at the association by the DLBCL histologic type (Sup-

porting Information Table S2), we found that a polymorphism in GSTM3

(rs1799735) modified the association between alcohol consumption and risk

of DLBCL (P for interaction 5 0.023). Among never drinkers, individuals with

null alleles of GSTM3 (rs1799735) had a significantly decreased risk of

DLBCL (OR 5 0.39, 95% CI: 0.16–0.94) compared with individuals with the

alleles, whereas among drinkers, a slightly increased risk of DLBCL was

found for women with null alleles (OR 5 1.26, 95% CI: 0.77–2.04).

We also observed a twofold increased risk of DLBCL (OR 5 2.04, 95%

CI: 1.00–4.13) for women who carried GSTP1 (rs1695) AG/GG genotypes

compared with those with AA genotype among nondrinkers, but not among

drinkers. An �60% reduced risk of DLBCL (OR 5 0.42, 95% CI: 0.25–0.72)

was found for women with NAT1*10 genotype compared with women without

NAT1*10 genotype among drinkers but not among nondrinkers. However,

there was no significant interaction was found for these genotypes.

When we looked at the associations of DLBCL and specific types of bev-

erages (Supporting Information Table S2), we found that polymorphisms in

CYP1A1 (rs1048943), CYP2C9 (rs1799853), GSTM3 (rs1799735), and

NAT1 modified the association between liquor consumption and risk of

DLBCL (P for interaction 5 0.04, 0.02, 0.03, and 0.02, respectively). Among

liquor drinkers, individuals with variant or null alleles of CYP1A1

(rs1048943), CYP2C9 (rs1799853), or GSTM3 (rs1799735) had an

increased risk of DLBCL (OR 5 3.22, 95% CI: 1.23–8.44; 1.77:0.95, 3.32;

1.32:0.72, 2.44, respectively) compared with individuals with the wildtype

alleles. Among each of the three kinds of beverage drinkers, a similar

decreased risk of DLBCL was observed (0.48:0.27, 0.88; 0.40:018, 0.90;

0.30 : 0.15, 0.60 for wine, beer and liquor drinkers, respectively).

When we looked at the association for the marginal zone and T-cell histo-

logic types (Supporting Information Table S2), no significant association was

identified between genetic polymorphisms in xenobiotic genes of interest

among either nondrinkers or drinkers of any type of alcoholic beverage or

for drinkers of specific alcoholic beverages including wine, beer, or liquor.

However, for the marginal zone histologic type (Supporting Information Table

S2), significant interactions were observed for the GSTP1 (rs1695) genotype

and wine drinking (P-interaction 5 0.02). For the T-cell histologic type (Sup-

porting Information Table S2), significant interactions were observed for the

NAT1*10 genotype and drinking (P-interaction 5 0.02) and for liquor drinking

(P-interaction 5 0.02). However, for the rarer subtypes, the results were

based on small numbers of cases, some <5.

When we restricted the sample to non-Hispanic Caucasians, the results

were unchanged.

We evaluated SNPs that were drawn from 11 key genes that play a role in

the mediation of carcinogen metabolism. Overall, our results do not confer evi-

dence that the relationship between NHL and alcohol intake is modified by

common genetic variation in CYP, GST, and NAT genes. However, when we

evaluated the DLBCL histologic type, we identified significant interactions for

multiple GSTs and NATs according to alcohol intake. Because of limited case

numbers for the DLBCL as well as other subtypes of interest, we recommend

that these findings be pursued in a larger study population in the future.

Many of the cytochrome P450 (CYP) enzymes are known to oxidize etha-

nol into acetaldehyde, and the expression of most of CYPs, especially

CYP1A1 and CYP2C9, is inducible by ethanol [23–25]. As such, the three-

fold increased risk of DLBCL for those with a heterozygous or wildtype

CYP1A1 genotype in liquor drinkers may be due to the higher ethanol con-

tent of liquor (361 mg/ml) compared with wine (79 mg/ml) and beer (36 mg/

ml) and higher CYP expression [2]. CYP1A1 is also associated with the

metabolism of many other potential carcinogens, such as nitrosamines,

some components of tobacco smoke and many organic chlorinated and non-

chlorinated solvents, including benzene [26], which may increase the risk of

DLBCL [19,27]. A potential explanation for our finding is that alcohol con-

sumption may induce the over-expression of CYPs that accelerates the

metabolism of toxicants derived from other exposures, thus increasing the

risk of cancer. The nonsignificant decreased risk in nondrinkers compared

with the increased observed risk in those receiving a heavy dose of ethanol

suggests that the alcohol may play an important role in inducing the capacity

of those with the impaired genotype.

Glutathione S-transferases (GSTs) consist of a family of isoenzymes that

also play an important role in the detoxification of endogenous compounds

as they catalyze the conjugation of these compounds to facilitate excretion

from the body. In our study, we found a twofold increased risk of DLBCL

among nondrinkers with variant alleles of GSTP1 suggesting that the low

catalytic efficiency may impair the detoxification of other harmful substances.

However, our results also showed a significant interaction as never drinkers

with a null GSTM3 genotype had a significantly decreased risk of DLBCL

compared with individuals with the alleles, whereas a slightly increased risk

of DLBCL was found for drinkers with the null genotype. It has been shown

previously that the GSTM3 null allele has an increased transcription poten-

tial and enhances the detoxification activity of GSTM3-encoded protein [28].

A potential explanation for our finding is subsequently that the alcohol-

induced expression of oxidases (such as CYPs) creates more reactive inter-

mediates which add to the burden of detoxification.

NATs are also important in the metabolism of toxicants as they catalyze

the conjugation of compounds to prepare them for excretion. The rate of

acetylation is thought to be related to the toxicity of a compound as it may

letters

214 American Journal of Hematology

Page 26: ASH 2009 meeting report—Top 10 clinically oriented abstracts in chronic lymphocytic leukemia

affect how quickly a chemical is excreted. In our study, we found a signifi-

cantly decreased risk of DLBCL among drinkers with a NAT1*10 genotype,

though this was not observed among nondrinking women. The NAT1*10

allele has been associated with a rapid acetylator phenotype both in vitro

[29] and in vivo [30]. Our results suggest that the expression of NAT1*10

genotype is alcohol-inducible, and it reduces the risk of DLBCL by increas-

ing the rate at which environmental and cancer-causing agents are acety-

lated and excreted from the body.

Our study has several strengths. It is a population-based, case-control

study with both incident cases that are histologically confirmed and highly

accurate genotyping data. The primary limitation of our study is that the

sample size is modest and the number of cases in several histologic sub-

groups was small. This resulted in reduced power to detect associations for

SNPs with low allele frequencies. It was limited to women and may be non-

generalizable to the entire population. Information bias, resulting from expo-

sure misclassification is likely to have been nondifferential, thus biasing our

risk estimates toward the null. Furthermore, our findings were based on

small numbers and could be due to chance. In addition, because of the

large number of comparisons, we cannot rule out chance findings due to

multiple comparisons. As such, the positive findings in our report require

replication in larger studies with greater power, which will be particularly val-

uable if tagged SNPs with full genomic coverage of the most promising can-

didate genes are used.

In sum, our study suggested that the polymorphisms in key metabolic

pathway genes may be related to the risk of DLBCL, and this association

may be modified by alcohol consumption. We did not find this to be true for

NHL overall or for the other histologic subtypes. Our results confer support

for the need for this hypothesis to be pursued in a larger study population,

with a particular focus on DLBCL.

Acknowledgments

The manuscript contents are solely the responsibility of the authors and

do not necessarily represent the official view of NCRR. This research was

approved by the DPH HIC. Certain data used in this study were obtained

from the Connecticut Department of Public Health. The authors assume full

responsibility for analyzes and interpretation of these data.

1Department of Epidemiology, National Institute of Environmental Health andRelated Product Safety, China CDC, Beijing, People’s Republic of China

2Department of Epidemiology and Public Health,Yale University School of Public Health, New Haven, Connecticut

3Division of Cancer Epidemiology and Genetics, National Cancer Institute,National Institutes of Health, Rockville, Maryland

4Chinese Academy of Medical Sciences, Cancer Institute/Hospital, Beijing,People’s Republic of China

*Correspondence to: Yawei Zhang, Yale University School of Public Health, 60College Street LEPH 440, New Haven, Connecticut 06520.

E-mail: [email protected] grant sponsor: National Cancer Institute (Intramural Research Program);

Grant numbers: CA62006Grant sponsor: National Institute of Health; Contractgrant numbers: 1D43TW008323-01, 1D43TW007864-01Grant sponsor: National

Center for Research Resources (NCRR); Grant numbers: UL1 RR024139Additional Supporting Information may be found in the online version of this article

Published online 4 December 2009 in Wiley InterScience(www.interscience.wiley.com).

DOI: 10.1002/ajh.21608Conflict of interest: Nothing to report.

References

1. Chiu BC, Cerhan JR, Gapstur SM, et al. Alcohol consumption and non-Hodgkin lymphoma in a cohort of older women. Br J Cancer 1999;80:1476–1482.

2. Morton LM, Zheng T, Holford TR, et al. Alcohol consumption and risk of non-Hodgkin lymphoma: A pooled analysis. Lancet Oncol 2005;6:469–476.

3. Chiu BC, Weisenburger DD, Cantor KP,et al. Alcohol consumption, family historyof hematolymphoproliferative cancer, and the risk of non-Hodgkin’s lymphomain men. Ann Epidemiol 2002;12:309–315.

4. Tavani A, Gallus S, La Vecchia C, Franceschi S. Alcohol drinking and risk ofnon-Hodgkin’s lymphoma. Eur J Clin Nutr 2001;55:824–826.

5. Chang ET, Smedby KE, Zhang SM, et al. Alcohol intake and risk of non-Hodgkinlymphoma inmen andwomen. Cancer Causes Control 2004;15:1067–1076.

6. Morton LM, Holford TR, Leaderer B, et al. Alcohol use and risk of non-Hodgkin’slymphoma among Connecticut women (United States). Cancer Causes Control2003;14:687–694.

7. Lim U, Morton LM, Subar AF, et al. Alcohol, smoking, and body size in relationto incident Hodgkin’s and non-Hodgkin’s lymphoma risk. Am J Epidemiol2007;166:697–708.

8. Casey R, Piazzon-Fevre K, Raverdy N, et al. Case-control study of lymphoidneoplasm in three French areas: Description, alcohol and tobacco consumption.Eur J Cancer Prev 2007;16:142–150.

9. Monnereau A, Orsi L, Troussard X, et al. Cigarette smoking, alcohol drinking,and risk of lymphoid neoplasms: Results of a French case-control study. CancerCauses Control 2008;19:1147–1160.

10. Polesel J, Dal Maso L, La Vecchia C, et al. Dietary folate, alcohol consumption,and risk of non-Hodgkin lymphoma. Nutr Cancer 2007;57:146–150.

11. Talamini R, Polesel J, Spina M, et al. The impact of tobacco smoking and alcoholdrinking on survival of patients with non-Hodgkin lymphoma. Int J Cancer2008;122:1624–1629.

12. Besson H, Brennan P, Becker N, et al. Tobacco smoking, alcohol drinking andnon-Hodgkin’s lymphoma: A European multicenter case-control study (Epi-lymph). Int J Cancer 2006;119:901–908.

13. Sarmanova J,Benesova K, Gut I, et al. Genetic polymorphisms of biotransfor-mation enzymes in patients with Hodgkin’s and non-Hodgkin’s lymphomas.Hum Mol Genet 2001;10:1265–1273.

14. Gra OA, Glotov AS, Nikitin EA, et al. Polymorphisms in xenobiotic-metabolizinggenes and the risk of chronic lymphocytic leukemia and non-Hodgkin’s lym-phoma in adult Russian patients. Am J Hematol 2008;83:279–287.

15. Skibola CF, Lightfoot T, Agana L, et al. Polymorphisms in cytochrome P45017A1 and risk of non-Hodgkin lymphoma. Br J Haematol 2005;129:618–621.

16. Kerridge I, Lincz L, Scorgie F, et al. Association between xenobiotic gene poly-morphisms and non-Hodgkin’s lymphoma risk. Br J Haematol 2002;118:477–481.

17. De Roos JA, Gold G, Wang S, et al. Metabolic gene variants and risk ofnon-Hodgkin’s lymphoma. Cancer Epidemiol Biomarkers Prev 2006;15:1647–1653.

18. Brian C-H, Chiu CK, Gapstur SM, et al. Association of NAT and GST polymor-phisms with non-Hodgkin’s lymphoma: A population-based case-control study.Br J Haematol 2005;128:610–615.

19. Kilfoy BA, Zheng T, Lan Q, et al. Genetic polymorphisms in glutathione S-trans-ferases and cytochrome P450s, tobacco smoking, and risk of non-Hodgkin lym-phoma. Am J Hematol 2009;84:279–282.

20. Jaffe ES, Harris NL, Stein H, et al.Tumours of Haematopoietic and LymphoidTissues. Lyon:IARC press; 2001.

21. Packer BR, Yeager M, Staats B, et al. SNP500Cancer: A public resource forsequence validation and assay development for genetic variation in candidategenes. Nucleic Acids Res 2004;32:D528–D532.

22. Hein DW. Molecular genetics and function of NAT1 and NAT2: Role in aromaticamine metabolism and carcinogenesis. Mutat Res 2002;506–507:65–77.

23. Cowpland C, Su GM, Murray M, et al. Effect of alcohol on cytochrome p450arachidonic acid metabolism and blood pressure in rats and its modulation byred wine polyphenolics. Clin Exp Pharmacol Physiol 2006;33:183–188.

24. Badger TM, Huang J, Ronis M, Lumpkin CK. Induction of cytochrome P450 2E1during chronic ethanol exposure occurs via transcription of the CYP 2E1 genewhen blood alcohol concentrations are high. Biochem Biophys Res Commun1993;190:780–785.

25. Ronis MJ, Huang J, Crouch J, et al. Cytochrome P450 CYP 2E1 induction dur-ing chronic alcohol exposure occurs by a two-step mechanism associated withblood alcohol concentrations in rats. J Pharmacol Exp Ther 1993;264:944–950.

26. Ishikawa H, Miyatsu Y, Kurihara K, Yokoyama K. Gene-environmental interac-tions between alcohol-drinking behavior and ALDH2 and CYP2E1 polymor-phisms and their impact on micronuclei frequency in human lymphocytes. MutatRes 2006;594:1–9.

27. Thier R, Bruning T, Roos PH, Bolt HM. Cytochrome P450 1B1, a new keystonein gene-environment interactions related to human head and neck cancer? ArchToxicol 2002;76:249–256.

28. Yengi L, InskipA, Gilford J, et al. Polymorphism at the glutathione S-transferaselocus GSTM3: Interactions with cytochrome P450 and glutathione S-transferasegenotypes as risk factors for multiple cutaneous basal cell carcinoma. CancerRes 1996;56:1974–1977.

29. Bell DA, Badawi AF, Lang NP, et al. Polymorphism in theN-acetyltransferase 1(NAT1) polyadenylation signal: Association of NAT1*10 allelewith higherN-acety-lation activity in bladder and colon tissue. Cancer Res 1995;55:5226–5229.

30. Hein DW, McQueen CA, Grant DM, et al. Pharmacogenetics of the arylamineN-acetyltransferases: A symposium in honor of Wendell W. Weber. Drug MetabDispos 2000;28:1425–1432.

letters

American Journal of Hematology 215