You are on page 1of 18

2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.

sgm LaTeX2e(2002/01/18) P1: GBC


10.1146/annurev.med.56.082103.104704

Annu. Rev. Med. 2005. 56:1–16


doi: 10.1146/annurev.med.56.082103.104704
Copyright 
c 2005 by Annual Reviews. All rights reserved

MYELODYSPLASTIC SYNDROME
Wolf-K. Hofmann1 and H. Phillip Koeffler2
1
Department of Hematology and Oncology and Transfusion Medicine, University Hospital
“Benjamin Franklin”, 12200 Berlin, Germany; email: W.K.Hofmann@charite.de
2
Division of Hematology and Oncology, Cedars Sinai Medical Center, UCLA School
of Medicine, Los Angeles, California 90048
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

Key Words classification, clinical features, diagnosis, molecular alterations,


treatment
■ Abstract During the past 15 years, important progress has been made in the
understanding of the biology and prognosis of myelodysplastic syndrome (MDS).
MDS is a clonal disorder characterized by ineffective hematopoiesis, which can lead
to either fatal cytopenias or acute myelogenous leukemia (AML). Risk-adapted treat-
ment strategies were established because of the high median age (60–75 years) of
the MDS patients and the individual history of the disease (number of cytopenias,
cytogenetic changes, transfusion requirements). Allogeneic bone marrow transplanta-
tion currently offers the only potentially curative treatment, but this form of therapy
is not available for the typical MDS patient, who is >60 years of age. Therapy with
erythropoietin and G-CSF has improved the quality of life of selected patients. The
development of small molecules directed against specific molecular targets with min-
imal adverse effects is the hope for the future. Innovative uses of immunomodulatory
agents and the optimizing of cytotoxic treatment should continue to help in the treatment
of MDS.

CLASSIFICATION
Myelodysplastic syndrome (MDS) is a clonal disorder characterized initially by in-
effective hematopoiesis and subsequently by frequent development of acute myel-
ogenous leukemia (AML). Peripheral blood cytopenias in combination with a
hypercellular bone marrow exhibiting dysplastic changes are the hallmark of
MDS. In 1982, the French-American-British Cooperative Group classified five
subentities of MDS (1): refractory anemia, refractory anemia with excess of blasts
(RAEB), refractory anemia with excess of blasts in transformation (RAEB-T),
refractory anemia with ringed sideroblasts (RARS), and chronic myelomonocytic
leukemia (CMML). This classification based on morphological criteria was re-
cently revised, resulting in the World Health Organization (WHO) classification
(2–6) (Table 1). Because of the better prognosis of patients with an isolated cy-
togenetic aberration at 5q, the WHO classification includes the 5q– syndrome as
0066-4219/05/0218-0001$14.00 1
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

2 HOFMANN  KOEFFLER


TABLE 1 World Health Organization classification of myelodysplastic syndromes

Category Peripheral blood Bone marrow

RA Anemia Blasts <5%


Blasts <1% Ringed sid. <15%
Mono <1000/mm3
RARS Anemia Blasts <5%
Blasts <1% Ringed sid. ≥15%
Mono <1000/mm3
RCMD Cytopenias (bi- or pancytopenia) Dysplasia ≥10% of cells
Blasts <1% Blasts <5%
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org

Mono <1000/mm3 Ringed sid. <15%


by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

RCMD-RS Cytopenias (bi- or pancytopenia) Dysplasia ≥10% of cells


Blasts <1% Blasts <5%
Mono <1000/mm3 Ringed sid. ≥15%
RAEB-I Cytopenias Multilineage dysplasia
Blasts <5% Blasts 5%–9%
Mono <1000/mm3
RAEB-II Cytopenias Multilineage dysplasia
Blasts 5%–19% Blasts 10%–19%
Mono <1000/mm3 Auer rods ±
Auer rods ±
MDS–U Cytopenias Unilineage dysplasia
Blasts <1% Blasts <5%
MDS, isolated Anemia Blasts <5%
del(5q) Blasts <5% Isolated del(5q)
Platelets normal or increased Normal or increased mega.

Abbreviations: RA, refractory anemia; RARS, RA with ringed sideroblasts; RCMD, refractory cytopenia with
multilineage dysplasia; RCMD–RS, RCMD with ringed sideroblasts; RAEB, RA with excess blasts; MDS–U,
myelodysplastic syndrome unclassifiable; Mono, monocytes; Sid., sideroblasts; mega., megakaryocytes.

a separate entity. The initial chromosomal aberration, the age of the patient, and
the number and severity of the cytopenias are important to evaluate the prognosis
of MDS, as summarized in the International Prognostic Scoring System, IPSS (7)
(Table 2). The median survival of MDS patients according to this classification
ranges from 6 years for low-risk to 6 months for high-risk patients.

Secondary MDS
Secondary MDS is a term that we use to emphasize that the MDS results from expo-
sure to a mutagen. This could occur as a consequence of therapy for another disease
(treatment-related MDS) or exposure to a toxic material, such as benzene. MDS has
been described following therapy of malignancies (e.g., Hodgkin’s disease, non-
Hodgkin lymphomas, multiple myeloma), including autologous transplantation
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

MYELODYSPLASTIC SYNDROMES 3

TABLE 2 International prognostic scoring system for myelodysplastic syndromes

Score Bone marrow Cytopenias


valuea blasts (%) Karyotype (lineages affected)

0 <5 Normal, sole: –Y, del 5q, del 20q 0 to 1


0.5 5–10 Others 2 to 3
1.0 Complexb and/or
chromosome 7 anomalies
1.5 11–20
2.0 21–30
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

a
The prognostic score is determined by the sum of the single scoring values. The risk groups are determined as
follows (brackets: median survival): Low = 0 points (5.7 years); Intermediate–1 = 0.5–1.0 points (3.5 years);
Intermediate–2 = 1.5–2.0 points (1.2 years); High ≥2.5 points (6 months).
b
≥3 chromosomal abnormalities.

(8–10). Alkylating chemotherapy most frequently is implicated as the causative


class of agents (11). The highest incidence of secondary MDS/AML occurs 3–
8 years after treatment of the antecedent malignancy (12).
After therapy with DNA topoisomerase II inhibitors, secondary MDS/AML
develops within a period of about 2–3 years (12). Recently, several single reports
have indicated that the NUP98 gene on chromosome 11p15 is one of the targets for
chromosomal translocations in patients with therapy-related MDS after exposure
to poly-chemotherapy including topoisomerase II inhibitors (13).

5q– Syndrome
This clinically distinct entity involves a deletion of the long arm of chromosome
5 and is characterized by a female preponderance, thrombocytosis, prominent
megaloblastoid erythroid hyperplasia in the bone marrow, and megakaryocytes
that are unusually large (often >30 µm in diameter) and often have a single,
eccentric, round nucleus (14). Interestingly, the rate of leukemic transformation
in individuals with 5q– syndrome is only 25% after an observation period of
15 years, as compared to about 40% in patients with a normal karyotype (7).
The underlying molecular targets in 5q– syndrome are still unknown. The criti-
cal segment may be at chromosome 5q31 (15, 16). The human GRAF (GTPase
regulator associated with focal adhesion kinase) gene located in this region can
fuse to the MLL (mixed lineage leukemia) gene, disrupting both alleles (17), but
the significance of this gene in the 5q– syndrome is unclear. Also, the loss of one
allele on 5q (haploinsufficiency) could contribute to the 5q– abnormality, similar
to what occurs in the familial platelet disorder with predisposition to AML (see
below).
In striking contrast to this syndrome is the loss of chromosome 5 or
5q– in therapy-related MDS, which presages an almost inevitable leukemic
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

4 HOFMANN  KOEFFLER

transformation. This may be associated with loss of regions of 5q other than the
region lost in the 5q– syndrome.

Hypoplastic MDS
Usually, patients with MDS have a hypercellular bone marrow. In ∼10% of cases,
peripheral blood cytopenia is associated with a hypoplastic bone marrow. In such
cases distinguishing MDS from aplastic anemia may be difficult, but the presence
of cytogenetic changes can help to identify malignant clonal cells, supporting the
diagnosis of MDS. The pathophysiology of hypoplastic MDS is not known; autore-
active and clonal-involved T-cells are believed to suppress normal hematopoietic
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org

cells by secretion of inhibitory cytokines. This subtype is most likely to respond


by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

to immunosuppressive agents.

CLINICAL FEATURES
MDS occurs primarily in the elderly population, with a median age between 60
and 75 years and a surprisingly high incidence of about 15–50 cases per 100,000
individuals over the age of 70 per year (18, 19). Despite this predominance of
elderly patients, MDS occasionally occurs in younger individuals and even in
children. In recent years, the incidence of MDS appears to be rising, perhaps
partly owing to increased diagnostic awareness. The sex distribution is balanced
(18), despite earlier reports that indicated a male preponderance. The exception is
CMML, which has a clear male preponderance (20).
Fatigue and exertional dyspnea may develop over a prolonged period, often
exceeding 6–12 months. These symptoms may be misinterpreted as either cardiac
failure or pulmonary disease, particularly in elderly patients. Approximately half
of the individuals are asymptomatic at the time of initial diagnosis and are usually
diagnosed after a routine blood count. Progressive hematopoietic failure leading
to anemia, thrombocytopenia, and leukopenia, either alone or in any combination,
is the dominant finding in MDS. Anemia is an almost universal characteristic at
the time of initial diagnosis; more than 80% of patients present with a hemoglobin
concentration below 10 g/dl. The reticulocyte count usually is reduced.
The peripheral blood leukocyte count is low in ∼25%–30% of individuals with
MDS, but because the blast cells replace more of the bone marrow, 40% of RAEB
cases and 80% of RAEB-T cases can have leukopenia. Granulocytes may exhibit a
reduced segmentation and either diminished or absent granulation. Approximately
one third of the individuals have recurrent infections. These occur not only because
of granulocytopenia but also as a result of defects of neutrophil function, including
impaired chemotaxis and reduced phagocytic activity.
Signs of bleeding, mainly petechiae, gingival bleeding, or hematoma following
trivial injuries, are surprisingly uncommon, given the frequency of thrombocy-
topenia. Fewer than 10% of patients will present initially with serious bleeding,
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

MYELODYSPLASTIC SYNDROMES 5

e.g., gastrointestinal hemorrhage, macrohematuria, or retinal or central nervous


system hemorrhage.

CYTOGENETICS AND MOLECULAR ALTERATIONS

The cytogenetic changes found in MDS are not unique to the disease. Both struc-
tural and numerical cytogenetic changes may occur (reviewed in References 21
and 22). The most frequent chromosomal abnormalities in MDS involve deletions
of chromosomes 5, 7, 11, 12, and 20 and/or trisomy 8 (Table 3). The incidence
of chromosomal abnormalities is about 30%–50% in primary MDS and 80% in
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

mutagen-related MDS. The latter often features complex changes that frequently
involve deletions of chromosomes 5 and/or 7. Translocations are rare in MDS.
MDS-related chromosomal deletions imply alterations in tumor suppressor genes
or DNA repair genes. Usually these changes require two hits: mutation of the target
gene and loss of the second allele through deletion, duplication, or recombination.
The underlying causes of primary MDS are still being defined. Analysis of
cytogenetic abnormalities, G6PD isoenzymes, restriction-linked polymorphisms,
and X-linked DNA polymorphisms of the androgen receptor have shown that MDS
is a clonal abnormality of the hematopoietic stem cell characterized by defective
maturation and, in advanced stages, uncontrolled proliferation (23). Lymphocytes
are probably not involved in the clonal hematopoiesis in MDS (24), but precursors
of erythrocytes, platelets, neutrophils, monocytes, eosinophils, and basophils are
members of the abnormal clone. Figure 1 proposes a multistep pathogenesis of
MDS. After initial damage of the progenitor cell by a toxin or spontaneous mu-
tation, several additional alterations may affect these cells, providing them with a
growth advantage. These alterations can influence expression of cell cycle-related
genes, transcription factors, and tumor suppressor genes.

TABLE 3 Most frequent chromosomal aberrations in MDS


patients (frequency of chromosomal aberration is given in

brackets)

Numerical Translocations Deletions

+8 (19%) inv 3 (7%) del 5q (27%)


–7 (15%) t(1;7) (2%) del 11q (7%)
+21 (7%) t(1;3) (1%) del 12q (5%)
–5 (7%) t(3;3) (1%) del 20q (5%)
t(6;9) (<1%) del 7q (4%)
t(5;12) (<1%) del 13q (2%)

Symbols: –, loss of chromosome; +, additional chromosome; inv, inversion; t,
translocation; del, deletion.
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

6 HOFMANN  KOEFFLER
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

Figure 1 Multistep pathogenesis in MDS. The initial genetic insult to the hematopoi-
etic stem cell can be caused by chemicals, radiation, cytotoxic drugs, or random en-
dogenous mutations. The accumulation of several alterations that may affect cell cycle
control, transcription, or tumor suppressors results in the expansion of the MDS clone.
The progression to leukemia probably does not depend on the order of occurrence of
genetic alterations but depends on which genes are altered. The final step, leukemic
transformation, may be enhanced by alteration of additional genes, including tumor
suppressor genes, and/or by hypermethylation of critical targets. The later stage of
leukemogenesis is associated with decreased apoptosis.

Recent studies have suggested that microarray analyses can provide sufficient
data to detect genes or gene patterns associated with alterations of specific cellular
pathways or signal cascades in MDS (25). The prediction of prognosis or risk type
of MDS from gene expression data is a long way from practice (26, 27), but it may
have a strong impact in the future on classification and risk definition in MDS.

TREATMENT
Patients with MDS are mainly older individuals suffering from accompanying
diseases. Therefore, various strategies have been used to treat patients with MDS.
Rather than to offer a cure, the main therapeutic goals in patients with MDS
are often to improve hematopoiesis and to ensure an age-related quality of life.
Low-intensity therapies, defined as treatments that permit outpatient management,
are often directed at patients with low-risk MDS (IPSS low and intermediate-1).
Such strategies are not necessarily associated with either improved survival or
progression-free survival.
Patients with high-risk MDS (IPSS intermediate-2 and high) need high-intensity
therapies (aggressive antileukemic chemotherapy and/or stem cell transplantation)
to eliminate the expanded clonal cells and to induce hematological responses.
Because of the high median age of patients with MDS, only about one third of
high-risk MDS patients can receive intensive cytotoxic treatment. For patients
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

MYELODYSPLASTIC SYNDROMES 7

who do not qualify for intensive therapy, experimental treatments to suppress,


differentiate, or eradicate the malignant clone are under investigation.

Supportive Care
Supportive care generally is the mainstay of therapy. Patients should be treated
with erythrocytes for symptomatic anemia. To reduce the risk of iron overload
in patients who receive >10 erythrocyte transfusions per year, therapy with an
iron chelator should be used, e.g., by bolus subcutaneous administration (28). An
optimal hemoglobin level is not known, but usually symptoms such as fatigue and
dyspnea occur if the hemoglobin level drops below 8 g/dl, prompting erythrocyte
transfusions. Individuals with a history of coronary heart disease have to be trans-
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

fused more aggressively. To minimize the risk of myocardial infarction in these


patients, the hemoglobin level should be 10 g/dl or greater.
Platelet transfusions are restricted to bleeding complications because repeated
platelet transfusions are associated with alloimmunization, which results in refrac-
toriness to donor platelets. In cases of refractoriness to either pooled or single-donor
platelet transfusions, administration of antifibrinolytic agents or vasopressin may
stop minor bleeding, such as from the nose or gingiva.
In the event of documented infections, as well as in the case of fever of unknown
origin, antibiotics must be given early. The danger of sepsis from common infec-
tions is greater in neutropenic patients than in normal individuals, and intravenous
antibiotics are often necessary.

Allogeneic Transplantation
Allogeneic bone marrow transplantation (BMT) or peripheral blood progenitor cell
transplantation (PBPCT) is a curative therapy for MDS. The results of allogeneic
transplantation vary considerably depending on the subtype of disease and other
clinical factors (29). The overall long-term disease-free survival is 60% for IPSS
low, 40% for IPSS intermediate-2, and 20% for IPSS high individuals (30–32).
Several important factors can improve the outcome after allogeneic transplantation
(33), including age (increased survival in patients aged <37 years), low number
of cytopenias (<2 is associated with good prognosis after allotransplantation),
low IPSS, matched sibling donor, short disease duration (<3–6 months), bone
marrow blast cells <5%, and the achievement of complete remission (CR) before
transplantation (Table 4). The non-relapse mortality (NRM) after 3 y in this set-
ting ranges from 12% to 30% for both sibling and unrelated matched transplants,
but it is up to 50% in unrelated HLA-nonidentical transplants. This considerable
therapy-associated mortality is related to a high incidence of graft-versus-host dis-
ease (GVHD), which is often associated with severe bacterial or fungal infections
including disseminated aspergillosis (32).
“Mini-transplantation” may be an alternative treatment option in MDS pa-
tients aged >60 years (34). These individuals receive partial hematopoietic and
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

8 HOFMANN  KOEFFLER

TABLE 4 Prognostic parameters that affect outcome after


allogeneic stem cell transplantation in patients with MDS
(adapted from Reference 33)

Favorable Unfavorable

Younger age Age >60 years


HLA-matched sibling donor Mismatched donor
Good performance status Poor performance status
Early-stage disease Advanced or secondary MDS
Disease of short duration Prolonged course of disease
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

immune ablation, and the transplantation results in a chimerism to encourage a


graft-versus-MDS/AML effect. This approach is still investigational.
The use of peripheral blood progenitor cells for transplantation is associated
with a rapid hematopoietic recovery. This significantly reduces the number of
days of fever, as well as the need for parenteral antibiotics, antifungal therapy, and
transfusion of erythrocytes and/or platelets. Total duration of hospitalization is also
decreased, compared with that of a historical matched bone marrow transplantation
group (35).

Autologous Transplantation
The high toxicity of allogeneic transplantation and the advanced age of most pa-
tients have prompted many investigators to evaluate the feasibility of autologous
transplants in high-risk MDS. Early results of PBPCT indicate that survival rates,
but not disease-free survival, may approach those of allogeneic transplantation
because of the low mortality associated with the autotransplant itself (36). A ret-
rospective analysis of 79 patients autografted for MDS showed an overall survival
of 39% and a disease-free survival of 34% at 2 years (37, 38).

Hematopoietic Growth Factors


Eighty-five percent of MDS patients have elevated serum erythropoietin levels.
Despite this fact, the rationale to treat MDS patients with recombinant erythro-
poietin (Epo) is the possibility that pharmaceutical doses may enhance the de-
fective proliferation and differentiation of the erythroid precursors. In ∼25% of
low-risk MDS patients receiving Epo, erythrocyte mass increases and/or transfu-
sion requirements are reduced. Factors associated with response to Epo include
serum Epo level below 200 U/L, absence of transfusion requirement, and ab-
sence of ringed sideroblasts (39). Long-term Epo responders have been described
(40).
Interestingly, the combination of Epo with either granulocyte-colony stimulat-
ing factor or granulocyte macrophage-colony stimulating factor (G- or GM-CSF)
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

MYELODYSPLASTIC SYNDROMES 9

has a synergistic effect in patients with MDS (41–46). In particular, RARS pa-
tients who have almost no response to Epo have a 40% response rate (increased
erythrocyte mass) if they receive the combination of Epo and G-CSF. The clinical
role for thrombopoietin in MDS patients with low platelet counts (<20,000/µl)
must be further defined by careful, controlled clinical trials that evaluate bleed-
ing events, side effects, and particularly the risk of development of neutralizing
antibodies.

Demethylating Agents
Many genes have regions in their promoter (CpG islands) that can be methylated at
the 5 position of cytosine, which silences expression of these genes. Theoretically,
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

demethylation of methylated genes that are important in differentiation and/or


apoptosis could have clinical applications. For example, p15INK4b, a cell cycle
brake, is frequently methylated in MDS but not in normal myeloid cells (47).
Both 5-azacytidine (Azacitidine; VidazaTM, Pharmion Corp., Boulder, CO) and
5-aza 2 -deoxycytidine (Decitabine, SuperGen, Inc., Dublin, CA) reduce DNA
methyltransferase activity and therefore can cause DNA hypomethylation (48, 49).
Whereas Azacitidine is incorporated mostly into mRNA and tRNA, Decitabine is
incorporated solely into DNA, possibly accounting for the somewhat different
toxicity profile of these two drugs.
Initial pilot trials with low-dose Azacitidine (50) and low-dose Decitabine (51)
provided encouraging results. A multicenter study of Azacitidine (52) confirmed its
efficacy. Recently, results of a multicenter phase II trial with low-dose intravenous
5-aza 2 -deoxycytidine (45 mg/m2 for 3 days every 6 weeks) were reported for
66 mostly elderly patients with advanced MDS (IPSS: 24% intermediate-1, 38%
intermediate-2, 38% high) (53). The overall hematologic response rate was 49%,
which included a response of 64% for high-risk individuals. The actuarial median
response duration was 31 weeks, with a response duration of 39 weeks and 36
weeks for patients who reached a PR or CR, respectively.

Immunosuppressive Agents
Immunosuppressive therapy has been proposed for the past 15 years as a treatment
option for MDS to reverse bone marrow failure by inhibiting intramedullary se-
cretion of proapoptotic cytokines. In a pilot study, 42 transfusion-dependent MDS
patients received antithymocyte globuline (ATG, 40 mg/kg/d for 4 days) (54, 55).
Remarkedly, erythrocyte transfusion independence occurred in 16 individuals, and
platelets increased in 14 of them. Three individuals with refractory anemia had a
complete remission. The response rate was 64% in the low-risk individuals and
33% in those with high-risk MDS. The median response duration was 10 months,
ranging from 3 to 38 months. A recently published update of this nonrandomized,
single-treatment study (56) reported an overall frequency of transfusion indepen-
dence of 34% in a total of 61 patients with MDS. Parameters to predict a good
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

10 HOFMANN  KOEFFLER

response to ATG treatment are low IPSS and a short history of erythrocyte trans-
fusions (57).
Several small studies used cyclosporin A (CSA) for MDS patients (58–60) with
variable results. A predictive marker for a good response may be the expression
of the HLA-DRB1∗ 1501 allele (61). Of a total of 10 patients studied, 6 patients
responded to CSA, all of whom had the 1501 allele. Four patients did not show
any response and did not have this human leukocyte antigen. It remains unclear
how ATG and/or CSA improves hematopoiesis in MDS, and why many months of
therapy are needed before a response occurs. Several multicenter trials using ATG
and/or CSA for MDS are ongoing.
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

Intensive Cytotoxic Treatment


At the present time, long-term benefit for individuals with MDS can be achieved
only by eradication of the abnormal clone and restoration of normal hematopoiesis.
Recent studies in which intensive cytotoxic treatment was administered to younger
individuals with high-risk MDS have produced remission rates ranging from 22%
to 79%. As a consequence of the further improved supportive care in patients
receiving intensive cytotoxic treatment, the remission rate achieved in high-risk
MDS patients is comparable with those for patients with de novo AML. On the
other hand, in a recently published study that evaluated intensive chemotherapy
alone compared to chemotherapy followed by transplantation in a total of 269
patients with high-risk MDS, neither the chemotherapy nor the transplantation
showed a clear benefit (62).
The decision whether to pursue aggressive treatment should include stratifica-
tion according to the patient’s risk factors using the IPSS (7). The use of hematopoi-
etic growth factors permits more patients to receive intensive cytotoxic treatment.
Nevertheless, the duration of remission often is short, usually about 12 months (63).
Long-term remissions are associated with restoration of polyclonal hematopoiesis
(64), and the achievement of a partial remission after induction therapy may be of
clinical benefit for high-risk patients.

Arsenic Trioxide (As2O3)


Arsenic compounds have been used therapeutically for at least a millennium in
China. In the West, arsenic in the form of “Fowler’s solution” was used in the
middle of the nineteenth century to treat fever of unknown orign. Most recently,
As2O3 has produced very good response in acute promyelocytic leukemia (65–67).
Clinical studies to evaluate As2O3 in MDS are under way (68).

Future Experimental Approaches


Farnesyl transferase inhibitors showed in vivo efficacy in the treatment of pa-
tients with AML, conceivably by inhibiting the Ras pathway. The role of these
agents in the treatment of high-risk MDS patients who are not qualified to receive
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

MYELODYSPLASTIC SYNDROMES 11

intensive treatment is not yet established, but they may be a future option
(69).
The bone marrow of individuals with MDS contains an abnormally high num-
ber of blood vessels (70, 71). This has encouraged the investigation of inhibitors
of angiogenesis, such as SU5416 and thalidomide, and of inhibitors of vascu-
lar endothelial growth factor (VEGF) for individuals with either AML or MDS.
Thalidomide was initially developed as a “sleeping pill,” but it was found to
have activity in the treatment of patients with multiple myeloma. In MDS, treat-
ment with thalidomide as a single agent (72) resulted in 30%–40% of MDS pa-
tients showing a hematopoietic response, usually improved erythropoiesis. Further
studies of antiangiogenetic drugs, including the new thalidomide analogue CC-
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org

5013 (RevimidTM; Celgene, Warren, NJ) which may have immunomodulatory


by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

efficacy in patients with 5q– syndrome (A. List, personal communication), and
anti-TNFα therapeutics are ongoing. Ultimately, we will define the genetic lesion
of the patient and have in our armamentarium a therapy specific for that molecular
abnormality.

Overall Treatment Approach in MDS


The treatment decision should take into consideration the patient’s disease risk
according to IPSS, age, and performance status. Based on these factors, possible
treatment strategies are as follows:

1. Individuals up to the (biological) age of ∼55–60 years are candidates for al-
logeneic transplantation from an HLA-matched (sibling or unrelated) donor.
The patients should be carefully informed about the risks of the allogeneic
stem cell transplantation. The alternative treatment options (see below)
should be discussed.
2. Patients with IPSS low or intermediate-1 MDS who have no HLA-identical
donor, or are older than 60 years and have good clinical performance, should
receive either supportive care or, when necessary, a trial of erythropoietin. For
nonresponders to erythropoietin, the combination therapy of erythropoietin
with G-CSF may be effective. Alternatively, immunosuppressive therapy
or other experimental approaches should be considered. As their disease
progresses, various therapies might be evaluated in the context of an ongoing
clinical study.
3. Patients with IPSS high or intermediate-2 MDS, older than 60 years and
having a good clinical performance, are candidates for intensive cytotoxic
therapy, followed by consolidation therapy and perhaps autologous trans-
plantation.
4. Individuals who are elderly and/or in poor clinical condition should re-
ceive supportive care. If possible and desired by the patient, investigational,
outpatient-based therapy (e.g., demethylating drugs or thalidomide) should
be offered.
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

12 HOFMANN  KOEFFLER

The Annual Review of Medicine is online at http://med.annualreviews.org

LITERATURE CITED
1. Bennett JM, Catovsky D, Daniel MT, et al. an assessment of risk factors. Blood 95:
1982. Proposals for the classification of the 1588–93
myelodysplastic syndromes. Br. J. Haema- 10. Pedersen-Bjergaard J, Andersen MK,
tol. 51:189–99 Christiansen DH. 2000. Therapy-related
2. Bennett JM. 2000. World Health Organi- acute myeloid leukemia and myelodyspla-
zation classification of the acute leukemias sia after high-dose chemotherapy and au-
and myelodysplastic syndrome. Int. J. tologous stem cell transplantation. Blood
Hematol. 72:131–33 95:3273–79
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

3. Harris NL, Jaffe ES, Diebold J, et al. 11. Tester WJ, Kinsella TJ, Waller B, et al.
2000. The World Health Organization clas- 1984. Second malignant neoplasms com-
sification of neoplastic diseases of the plicating Hodgkin’s disease: the National
haematopoietic and lymphoid tissues: re- Cancer Institute experience. J. Clin. Oncol.
port of the Clinical Advisory Committee 2:762–69
meeting, Airlie House, Virginia, November 12. Leone G, Mele L, Pulsoni A, et al. 1999.
1997. Histopathology 36:69–86 The incidence of secondary leukemias.
4. Germing U, Gattermann N, Strupp C, et al. Haematologica 84:937–45
2000. Validation of the WHO proposals for 13. Ahuja HG, Felix CA, Aplan PD. 2000. Po-
a new classification of primary myelodys- tential role for DNA topoisomerase II poi-
plastic syndromes: a retrospective analy- sons in the generation of t(11;20)(p15;q11)
sis of 1600 patients. Leuk. Res. 24:983– translocations. Genes Chromosomes Can-
92 cer 29:96–105
5. Greenberg P, Anderson J, de Witte T, et al. 14. Thiede T, Engquist L, Billstrom R. 1988.
2000. Problematic WHO reclassification Application of megakaryocytic morphol-
of myelodysplastic syndromes. Members ogy in diagnosing 5q– syndrome. Eur. J.
of the International MDS Study Group. J. Haematol. 41:434–37
Clin. Oncol. 18:3447–52 15. Carter G, Ridge S, Padua RA. 1992. Ge-
6. Vardiman JW, Harris NL, Brunning RD. netic lesions in preleukemia. Crit. Rev.
2002. The World Health Organization Oncol. 3:339–64
(WHO) classification of the myeloid neo- 16. Willman CL, Sever CE, Pallavicini MG,
plasms. Blood 100:2292–302 et al. 1993. Deletion of IRF-1, mapping to
7. Greenberg P, Cox C, LeBeau MM, et al. chromosome 5q31.1, in human leukemia
1997. International scoring system for eval- and preleukemic myelodysplasia. Science
uating prognosis in myelodysplastic syn- 259:968–71
dromes. Blood 89:2079–88 17. Borkhardt A, Bojesen S, Haas OA, et al.
8. Pedersen-Bjergaard J, Pedersen M, Roul- 2000. The human GRAF gene is fused
ston D, et al. 1995. Different genetic to MLL in a unique t(5;11)(q31;q23) and
pathways in leukemogenesis for patients both alleles are disrupted in three cases of
presenting with therapy-related myelodys- myelodysplastic syndrome/acute myeloid
plasia and therapy-related acute myeloid leukemia with a deletion 5q. Proc. Natl.
leukemia. Blood 86:3542–52 Acad. Sci. USA 97:9168–73
9. Krishnan A, Bhatia S, Slovak ML, 18. Aul C, Bowen DT, Yoshida Y. 1998.
et al. 2000. Predictors of therapy-related Pathogenesis, etiology and epidemiology
leukemia and myelodysplasia following of myelodysplastic syndromes. Haemato-
autologous transplantation for lymphoma: logica 83:71–86
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

MYELODYSPLASTIC SYNDROMES 13

19. Aul C, Germing U, Gattermann N, et al. ysis according to IPSS score. Leukemia
1998. Increasing incidence of myelodys- 12(Suppl. 1):S25–29
plastic syndromes: real or fictitious? Leuk. 30. Anderson JE, Appelbaum FR, Schoch G,
Res. 22:93–100 et al. 1996. Allogeneic marrow transplan-
20. Ribera JM, Cervantes F, Rozman C. 1987. tation for refractory anemia: a compari-
A multivariate analysis of prognostic fac- son of two preparative regimens and anal-
tors in chronic myelomonocytic leukaemia ysis of prognostic factors. Blood 87:51–
according to the FAB criteria. Br. J. Haema- 58
tol. 65:307–11 31. Ballen KK, Gilliland DG, Guinan EC,
21. Mecucci C, La Starza R. 1999. Cytogenet- et al. 1997. Bone marrow transplanta-
ics of myelodysplastic syndromes. Forum tion for therapy-related myelodysplasia:
(Genova) 9:4–13 comparison with primary myelodyspla-
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

22. Fenaux P, Morel P, Lai JL. 1996. Cyto- sia. Bone Marrow Transplant. 20:737–
genetics of myelodysplastic syndromes. 43
Semin. Hematol. 33:127–38 32. Deeg HJ, Storer B, Slattery JT, et al. 2002.
23. Parker J, Mufti GJ. 1996. Ras and myelo- Conditioning with targeted busulfan and
dysplasia: lessons from the last decade. cyclophosphamide for hemopoietic stem
Semin. Hematol. 33:206–24 cell transplantation from related and unre-
24. Saitoh K, Miura I, Takahashi N, et al. lated donors in patients with myelodysplas-
1998. Fluorescence in situ hybridization of tic syndrome. Blood 100:1201–7
progenitor cells obtained by fluorescence- 33. Luger S, Sacks N. 2002. Bone mar-
activated cell sorting for the detection of row transplantation for myelodysplastic
cells affected by chromosome abnormality syndrome—who? when? and which? Bone
trisomy 8 in patients with myelodysplastic Marrow Transplant. 30:199–206
syndromes. Blood 92:2886–92 34. Slavin S, Nagler A, Naparstek E, et al.
25. Miyazato A, Ueno S, Ohmine K, et al. 1998. Nonmyeloablative stem cell trans-
2001. Identification of myelodysplastic plantation and cell therapy as an alterna-
syndrome-specific genes by DNA microar- tive to conventional bone marrow trans-
ray analysis with purified hematopoietic plantation with lethal cytoreduction for
stem cell fraction. Blood 98:422–27 the treatment of malignant and nonmalig-
26. Hofmann WK, de Vos S, Komor M, et al. nant hematologic diseases. Blood 91:756–
2002. Characterization of gene expression 63
of CD34+ cells from normal and myelodys- 35. Guardiola P, Runde V, Bacigalupo A,
plastic bone marrow. Blood 100:3553– et al. 2002. Retrospective comparison of
60 bone marrow and granulocyte colony-
27. Ueda M, Ota J, Yamashita Y, et al. stimulating factor-mobilized peripheral
2003. DNA microarray analysis of stage blood progenitor cells for allogeneic stem
progression mechanism in myelodysplas- cell transplantation using HLA identical
tic syndrome. Br. J. Haematol. 123:288– sibling donors in myelodysplastic syn-
96 dromes. Blood 99:4370–78
28. Franchini M, Gandini G, de Gironcoli M, 36. Laporte JP, Isnard F, Lesage S, et al.
et al. 2000. Safety and efficacy of subcuta- 1993. Autologous bone marrow transplan-
neous bolus injection of deferoxamine in tation with marrow purged by mafosfamide
adult patients with iron overload. Blood in seven patients with myelodysplas-
95:2776–79 tic syndromes in transformation (AML-
29. Appelbaum FR, Anderson J. 1998. Al- MDS): a pilot study. Leukemia 7:2030–
logeneic bone marrow transplantation for 33
myelodysplastic syndrome: outcomes anal- 37. de Witte T, van Biezen A, Hermans J,
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

14 HOFMANN  KOEFFLER

et al. 1997. Autologous bone marrow trans- colony-stimulating factor in the treatment
plantation for patients with myelodysplas- of myelodysplastic syndromes. Identifica-
tic syndrome (MDS) or acute myeloid tion of a subgroup of responders. The Span-
leukemia following MDS. Chronic and ish Erythropathology Group. Haematolog-
Acute Leukemia Working Parties of the ica 84:1058–64
European Group for Blood and Marrow 45. Thompson JA, Gilliland DG, Prchal JT,
Transplantation. Blood 90:3853–57 et al. 2000. Effect of recombinant human
38. de Witte T, Hermans J, Vossen J, et al. erythropoietin combined with granulocyte/
2000. Haematopoietic stem cell transplan- macrophage colony-stimulating factor in
tation for patients with myelo-dysplastic the treatment of patients with myelodys-
syndromes and secondary acute myeloid plastic syndrome. GM/EPO MDS Study
leukaemias: a report on behalf of the Group. Blood 95:1175–79
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

Chronic Leukaemia Working Party of the 46. Mantovani L, Lentini G, Hentschel B, et al.
European Group for Blood and Marrow 2000. Treatment of anaemia in myelodys-
Transplantation (EBMT). Br. J. Haematol. plastic syndromes with prolonged admin-
110:620–30 istration of recombinant human granulo-
39. Hellstrom-Lindberg E. 1995. Efficacy of cyte colony-stimulating factor and ery-
erythropoietin in the myelodysplastic syn- thropoietin. Br. J. Haematol. 109:367–
dromes: a meta-analysis of 205 patients 75
from 17 studies. Br. J. Haematol. 89:67– 47. Quesnel B, Guillerm G, Vereecque R, et al.
71 1998. Methylation of the p15(INK4b) gene
40. Hast R, Wallvik J, Folin A, et al. in myelodysplastic syndromes is frequent
2001. Long-term follow-up of 18 patients and acquired during disease progression.
with myelodysplastic syndromes respond- Blood 91:2985–90
ing to recombinant erythropoietin treat- 48. Cihak A. 1974. Biological effects of 5-aza-
ment. Leuk. Res. 25:13–18 cytidine in eukaryotes. Oncology 30:405–
41. Negrin RS, Stein R, Vardiman J, et al. 1993. 22
Treatment of the anemia of myelodys- 49. Jones PA, Taylor SM. 1980. Cellular dif-
plastic syndromes using recombinant hu- ferentiation, cytidine analogs and DNA
man granulocyte colony-stimulating factor methylation. Cell 20:85–93
in combination with erythropoietin. Blood 50. Silverman LR, Holland JF, Weinberg RS,
82:737–43 et al. 1993. Effects of treatment with 5-aza-
42. Negrin RS, Stein R, Doherty K, et al. cytidine on the in vivo and in vitro
1996. Maintenance treatment of the ane- hematopoiesis in patients with myelodys-
mia of myelodysplastic syndromes with plastic syndromes. Leukemia 7(Suppl. 1):
recombinant human granulocyte colony- 21–29
stimulating factor and erythropoietin: ev- 51. Zagonel V, Lo Re G, Marotta G, et al.
idence for in vivo synergy. Blood 87:4076– 1993. 5-Aza-2 -deoxycytidine (Decita-
81 bine) induces trilineage response in un-
43. Hellstrom-Lindberg E, Ahlgren T, Beguin favourable myelodysplastic syndromes.
Y, et al. 1998. Treatment of anemia in Leukemia 7:30–35
myelodysplastic syndromes with granulo- 52. Silverman LR, Demakos EP, Peterson BL,
cyte colony-stimulating factor plus erythro- et al. 2002. Randomized controlled trial of
poietin: results from a randomized phase II azacitidine in patients with the myelodys-
study and long-term follow-up of 71 pa- plastic syndrome: a study of the cancer and
tients. Blood 92:68–75 leukemia group B. J. Clin. Oncol. 20:2429–
44. Remacha AF, Arrizabalaga B, Villegas A, 40
et al. 1999. Erythropoietin plus granulocyte 53. Wijermans P, Lubbert M, Verhoef G, et al.
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

MYELODYSPLASTIC SYNDROMES 15

2000. Low-dose 5-aza-2 -deoxycytidine, 62. Oosterveld M, Muus P, Suciu S, et al.


a DNA hypomethylating agent, for the 2002. Chemotherapy only compared to
treatment of high-risk myelodysplastic chemotherapy followed by transplantation
syndrome: a multicenter phase II study in in high risk myelodysplastic syndrome
elderly patients. J. Clin. Oncol. 18:956– and secondary acute myeloid leukemia;
62 two parallel studies adjusted for vari-
54. Molldrem JJ, Caples M, Mavroudis D, ous prognostic factors. Leukemia 16:1615–
et al. 1997. Antithymocyte globulin for pa- 21
tients with myelodysplastic syndrome. Br. 63. de Witte T, Suciu S, Verhoef G, et al.
J. Haematol. 99:699–705 2001. Intensive chemotherapy followed by
55. Molldrem JJ, Jiang YZ, Stetler-Stevenson allogeneic or autologous stem cell trans-
M, et al. 1998. Haematological response of plantation for patients with myelodysplas-
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

patients with myelodysplastic syndrome to tic syndromes (MDSs) and acute myeloid
antithymocyte globulin is associated with a leukemia following MDS. Blood 98:2326–
loss of lymphocyte-mediated inhibition of 31
CFU-GM and alterations in T-cell recep- 64. Aivado M, Rong A, Germing U, et al.
tor Vβ profiles. Br. J. Haematol. 102:1314– 2000. Long-term remission after intensive
22 chemotherapy in advanced myelodysplas-
56. Molldrem JJ, Leifer E, Bahceci E, et al. tic syndromes is generally associated with
2002. Antithymocyte globulin for treat- restoration of polyclonal haemopoiesis. Br.
ment of the bone marrow failure associated J. Haematol. 110:884–86
with myelodysplastic syndromes. Ann. In- 65. Soignet SL, Maslak P, Wang ZG, et al.
tern. Med. 137:156–63 1998. Complete remission after treatment
57. Stadler M, Germing U, Kliche KO, et al. of acute promyelocytic leukemia with ar-
2004. A prospective, randomised, phase II senic trioxide. N. Engl. J. Med. 339:1341–
study of horse antithymocyte globulin vs 48
rabbit antithymocyte globulin as immune- 66. Shen ZX, Chen GQ, Ni JH, et al. 1997. Use
modulating therapy in patients with low- of arsenic trioxide (As2O3) in the treatment
risk myelodysplastic syndromes. Leukemia of acute promyelocytic leukemia (APL):
18:460–65 II. Clinical efficacy and pharmacokinet-
58. Porta F, Facchetti F, Tettoni K, et al. ics in relapsed patients. Blood 89:3354–
1998. Myelodysplastic syndrome in an in- 60
fant: induction of remission by cyclosporin. 67. Jing Y, Dai J, Chalmers-Redman RM, et al.
Lancet 352:1600–1 1999. Arsenic trioxide selectively induces
59. Biesma DH, van den Tweel JG, Ver- acute promyelocytic leukemia cell apop-
donck LF. 1997. Immunosuppressive ther- tosis via a hydrogen peroxide-dependent
apy for hypoplastic myelodysplastic syn- pathway. Blood 94:2102–11
drome. Cancer 79:1548–51 68. List A, Beran M, DiPersio J, et al. 2003. Op-
60. Jonasova A, Neuwirtova R, Cermak J, et al. portunities for Trisenox (arsenic trioxide)
1998. Cyclosporin A therapy in hypoplas- in the treatment of myelodysplastic syn-
tic MDS patients and certain refractory dromes. Leukemia 17:1499–507
anaemias without hypoplastic bone mar- 69. Reuter CW, Morgan MA, Bergmann L.
row. Br. J. Haematol. 100:304–9 2000. Targeting the Ras signaling path-
61. Okamoto T, Okada M, Yamada S, et al. way: a rational, mechanism-based treat-
2000. Good response to cyclosporine ther- ment for hematologic malignancies? Blood
apy in patients with myelodysplastic syn- 96:1655–69
dromes having the HLA-DRB1∗ 1501 al- 70. Aguayo A, Kantarjian H, Manshouri T,
lele. Leukemia 14:344–46 et al. 2000. Angiogenesis in acute and
2 Dec 2004 7:12 AR AR236-ME56-01.tex AR236-ME56-01.sgm LaTeX2e(2002/01/18) P1: GBC

16 HOFMANN  KOEFFLER

chronic leukemias and myelodysplastic 72. Raza A, Meyer P, Dutt D, et al. 2001.
syndromes. Blood 96:2240–45 Thalidomide produces transfusion in-
71. Bertolini F, Mancuso P, Gobbi A, et al. dependence in long-standing refractory
2000. The thin red line: angiogenesis in anemias of patients with myelodys-
normal and malignant hematopoiesis. Exp. plastic syndromes. Blood 98:958–
Hematol. 28:993–1000 65
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.
P1: JRX
December 9, 2004 15:2 Annual Reviews AR236-FM

Annual Review of Medicine


Volume 56, 2005

CONTENTS
MYELODYSPLASTIC SYNDROME, Wolf-K. Hofmann and H. Phillip Koeffler 1
G PROTEIN POLYMORPHISMS IN HYPERTENSION, ATHEROSCLEROSIS,
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org
by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

AND DIABETES, Winfried Siffert 17


POST-TRANSPLANT LYMPHOPROLIFERATIVE DISORDERS,
Stephen Gottschalk, Cliona M. Rooney, and Helen E. Heslop 29
METABOLIC SYNDROME: A CLINICAL AND MOLECULAR PERSPECTIVE,
David E. Moller and Keith D. Kaufman 45
NEW ANTICOAGULANT THERAPY, Lori-Ann Linkins and Jeffrey I. Weitz 63
ENDOTHELIAL PROGENITOR CELLS, Aarif Y. Khakoo and Toren Finkel 79
AROMATASE INHIBITORS: RATIONALE AND USE IN BREAST CANCER,
Cynthia Osborne and Debu Tripathy 103
ANDROPAUSE: IS ANDROGEN REPLACEMENT THERAPY INDICATED
FOR THE AGING MALE?, Rabih A. Hijazi and Glenn R. Cunningham 117
SURGICAL THERAPY FOR METASTATIC DISEASE TO THE LIVER,
David J. Bentrem, Ronald P. DeMatteo, and Leslie H. Blumgart 139
NEW STRATEGIES IN THE TREATMENT OF THE THALASSEMIAS,
Stanley L. Schrier and Emanuele Angelucci 157
NEW CONCEPTS IN VON WILLEBRAND DISEASE, J. Evan Sadler 173
GENETICS OF LONGEVITY AND AGING, Jan Vijg and Yousin Suh 193
PROGRESS TOWARD AN HIV VACCINE, Norman L. Letvin 213
THERAPEUTIC INTERVENTION AND TARGETS FOR SEPSIS, Todd W. Rice
and Gordon R. Bernard 225
MANAGEMENT OF PERIPHERAL VASCULAR DISEASE, I. Baumgartner,
R. Schainfeld, and L. Graziani 249
ROLE OF MAGNETIC RESONANCE IMAGING AND IMMUNOTHERAPY IN
TREATING MULTIPLE SCLEROSIS, Jingwu Zhang and George Hutton 273
DEFINITION AND CLINICAL IMPORTANCE OF HAPLOTYPES,
Dana C. Crawford and Deborah A. Nickerson 303
APPROACHES TO THERAPY OF PRION DISEASES, Charles Weissmann
and Adriano Aguzzi 321

v
P1: JRX
December 9, 2004 15:2 Annual Reviews AR236-FM

vi CONTENTS

ENDOMETRIOSIS: NEW GENETIC APPROACHES AND THERAPY,


David H. Barlow and Stephen Kennedy 345
SEVERE ACUTE RESPIRATORY SYNDROME (SARS): A YEAR IN
REVIEW, Danuta M. Skowronski, Caroline Astell, Robert C. Brunham,
Donald E. Low, Martin Petric, Rachel L. Roper, Pierre J. Talbot,
Theresa Tam, and Lorne Babiuk 357
GENE-ENVIRONMENT INTERACTIONS IN ASTHMA AND OTHER
RESPIRATORY DISEASES, Steven R. Kleeberger and David Peden 383
THE SILENT REVOLUTION: RNA INTERFERENCE AS BASIC BIOLOGY,
RESEARCH TOOL, AND THERAPEUTIC, Derek M. Dykxhoorn
Annu. Rev. Med. 2005.56:1-16. Downloaded from arjournals.annualreviews.org

and Judy Lieberman 401


by "UNIV. OF WISCONSIN, MADISON" on 03/11/05. For personal use only.

MANAGEMENT OF ADULT IDIOPATHIC THROMBOCYTOPENIC PURPURA,


Douglas B. Cines and Robert McMillan 425
MONOGENIC OBESITY IN HUMANS, I. Sadaf Farooqi
and Stephen O’Rahilly 443
APPLICATION OF MICROBIAL GENOMIC SCIENCE TO ADVANCED
THERAPEUTICS, Claire M. Fraser and Rino Rappuoli 459
ATRIAL FIBRILLATION: MODERN CONCEPTS AND MANAGEMENT,
Ashish Agarwal, Meghan York, Bharat K. Kantharia,
and Michael Ezekowitz 475
DNA REPAIR DEFECTS IN STEM CELL FUNCTION AND AGING,
Youngji Park and Stanton L. Gerson 495
HEMATOPOIETIC STEM AND PROGENITOR CELLS: CLINICAL AND
PRECLINICAL REGENERATION OF THE HEMATOLYMPHOID SYSTEM,
Judith A. Shizuru, Robert S. Negrin, and Irving L. Weissman 509
INHERITED SUSCEPTIBILITY TO COLORECTAL CANCER, Peter T. Rowley 539
APTAMERS: AN EMERGING CLASS OF THERAPEUTICS,
Shahid M. Nimjee, Christopher P. Rusconi, and Bruce A. Sullenger 555
GENE THERAPY FOR SEVERE COMBINED IMMUNODEFICIENCY,
Marina Cavazzana-Calvo, Chantal Lagresle, Salima Hacein-Bey-Abina,
and Alain Fischer 585

INDEXES
Subject Index 603
Cumulative Index of Contributing Authors, Volumes 52–56 637
Cumulative Index of Chapter Titles, Volumes 52–56 640
ERRATA
An online log of corrections to Annual Review of Medicine
chapters may be found at http://med.annualreviews.org/errata.shtml

You might also like