Vous êtes sur la page 1sur 16

Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.

com 973
CHRONIC MYELOID LEUKEMIA SYMPOSIUM ON ONCOLOGY PRACTICE: HEMATOLOGICAL MALIGNANCIES
From the Department of Leukemia, The University of Texas, M. D. Anderson
Cancer Center, Houston, Tex.
Dr Cortes has research grant support from Novartis, BMS, The Vaccine Com-
pany, ChemGenex Pharmaceuticals, Ltd, and Breakthrough Therapeutics.
Address correspondence to J orge E. Cortes, MD, Department of Leukemia,
The University of Texas, M. D. Anderson Cancer Center, 1515 Holcombe Blvd,
Unit 428, Houston, TX 77030 (e-mail: jcortes@mdanderson.org). Individual
reprints of this article and a bound booklet of the entire Symposium on
Oncology Practice: Hematological Malignancies will be available for purchase
from our Web site www.mayoclinicproceedings.com.
2006 Mayo Foundation for Medical Education and Research
Chronic Myeloid Leukemia: Diagnosis and Treatment
ALFONSO QUINTS-CARDAMA, MD, AND JORGE E. CORTES, MD
ALL = acut e l ymphobl ast i c l eukemi a; AML = acut e myel oi d l eukemi a;
AP = accel erat ed phase; ATP = adenosi ne t ri phosphat e; BP = bl ast
phase; CCGR = compl et e cyt ogenet i c response; CHR = compl et e hema-
t ol ogi c response; CML = chroni c myel oi d l eukemi a; CP = chroni c phase;
Grb2 = growt h fact or recept orbound prot ei n 2; IC
50
= i nhi bi t ory concen-
t rat ion t hat result s in a 50% reduct ion in proliferat ion; MCGR= major
cyt ogenet i c response; MMR = maj or mol ecul ar response; MUD =
mat ched unrelat ed donor; Ph = Philadelphia; PI3K = phosphat idylinosit ol
3-kinase; RT-PCR = reverse t ranscript asepolymerase chain react ion;
SCT = st em cell t ransplant at ion; STAT = signal t ransducer and act ivat or
of t ranscript ion; TKI = t yrosine kinase inhibit or; WBC = whit e blood cell
Chroni c myel oi d l eukemi a (CML) has become a model i n research
and management among mal i gnant di sorders. Si nce t he di scovery
of t he presence of a uni que and const ant chromosomal abnormal -
i t y sl i ght l y more t han 40 years ago, subst ant i al progress has been
made i n t he underst andi ng of t he bi ol ogy of t he di sease. Thi s
progress has t ransl at ed i nt o si gni fi cant i mprovement i n t he l ong-
t erm prognosi s of pat i ent s wi t h t hi s di sease. Thi s change came
fi rst wi t h t he use of st em cel l t ranspl ant at i on and i nt erferon al fa,
but recent l y i t has opened t he era of mol ecul arl y t arget ed t hera-
pi es. Imat i ni b, a pot ent and sel ect i ve t yrosi ne ki nase i nhi bi t or,
may be t he best exampl e of our at t empt s t o i dent i fy mol ecul ar
abnormal i t i es and devel op drugs di rect ed speci fi cal l y at t hem.
Furt hermore, t he underst andi ng of at l east some of t he mecha-
ni sms of resi st ance t o i mat i ni b has l ed t o rapi d devel opment of
new agent s t hat may overcome t hi s resi st ance. The out l ook t oday
for pat i ent s wi t h CML i s much bri ght er t han j ust a few years ago.
It i s our hope t hat t hi s fasci nat i ng j ourney i n CML can be repl i -
cat ed i n ot her mal i gnanci es. In t hi s art i cl e, we revi ew our current
underst andi ng of t hi s di sease.
Mayo Cl i n Proc. 2006;81(7):973-988
C
hronic myeloid leukemia (CML) is a clonal myelopro-
liferative disorder of a pluripotent stem cell
1
first de-
scribed by John Hughes Bennett in 1845 at The Royal
Infirmary of Edinburgh.
2
It has been classified as a myelo-
proliferative disorder along with agnogenic myeloid meta-
plasia, polycythemia vera, and essential thrombocythe-
mia.
3
Chronic myeloid leukemia was the first malignancy
that had a specific chromosomal abnormality uniquely
linked to it after the discovery of a minute chromosome
now known as the Philadelphia (Ph) chromosome,
4
later
defined to result from a t(9;22) reciprocal chromosomal
translocation.
5
Of critical importance was the demonstra-
tion that this translocation involved the ABL1 (Abelson)
protooncogene in chromosome 9 and the BCR (breakpoint
cluster region) gene in chromosome 22.
6,7
Since then, a
constant stream of clinical and basic advances has made
CML one of the most extensively studied human malignan-
cies. The introduction of rationally designed small mol-
ecules that target the tyrosine kinase activity of Bcr-Abl in
the treatment of CML has pioneered the development of
targeted therapies in cancer medicine. In this review, we
summarize the current knowledge of the molecular biology
of CML and discuss the current treatment modalities, in-
cluding the important historical role of recombinant inter-
feron alfa, the development of imatinib mesylate and the
new tyrosine kinase inhibitors (TKIs), stem cell transplan-
tation (SCT), and other novel therapeutic approaches still
in preclinical or early clinical development.
EPIDEMIOLOGY AND ETIOLOGY
Chronic myeloid leukemia is a rare disease worldwide.
Approximately 4600 new cases of CML were diagnosed in
2004 in the United States, among 33,400 new cases of
leukemia. Hence, CML represents 14% of all leukemias
and 20% of adult leukemias. The annual incidence is 1.6
cases per 100,000 adults,
8
with a slight male preponderance
(male-female ratio, 1.4:1). The median age at diagnosis
is 65 years. In contrast, CML is exceedingly rare in chil-
dren, and its incidence increases with age.
9
There are no
known hereditary, familial, geographic, ethnic, or eco-
nomic associations with CML; therefore, the disease is nei-
ther preventable nor inherited. In most patients, the factor
responsible for the induction of the Ph chromosome is un-
known, although CML has been observed with increased
frequency in individuals exposed to the atom bomb explo-
sions in Japan in 1945, in radiologists, and in patients with
ankylosing spondylitis treated with radiation therapy.
10
CLINICAL PRESENTATION AND NATURAL HISTORY
Classically, 3 phases of disease progression are recognized
in CML: chronic phase (CP), accelerated phase (AP), and
blast phase (BP) (Table 1). Currently, nearly 90% of pa-
tients with CML have their conditions diagnosed in the CP.
Frequently, the diagnosis is made incidentally after a rou-
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 974
CHRONIC MYELOID LEUKEMIA
tine complete blood cell count is performed for unrelated
reasons. This is due to the fact that patients with CML-CP
have a competent immune system and may remain asymp-
tomatic for prolonged periods. When symptoms appear, they
are generally related to the expansion of CML cells and
consist typically of malaise, weight loss, and discomfort
caused by splenomegaly.
14
Leukocytosis is a common fea-
ture of CP, and the white blood cell (WBC) count can be as
high as 1000 10
9
/L, leading in rare instances to signs and
symptoms of hyperviscosity, such as retinal hemorrhage,
priapism, cerebrovascular accidents, tinnitus, confusion,
and stupor. After a median of 3 to 5 years, untreated pa-
tients with CML-CP inevitably progress to CML-BP, an
aggressive form of acute leukemia highly refractory to
chemotherapy that is rapidly fatal.
15
The risk of transforma-
tion to CML-BP is estimated at 3% to 4% per year.
15,16
Chronic myeloid leukemia-AP is characterized by an in-
creasing arrest of maturation that usually heralds transfor-
mation to CML-BP.
Different criteria have been used to define CML-AP.
One commonly used classification defines AP as the pres-
ence of at least 1 of the following hematologic features:
15% or more blasts, 30% or more blasts plus promyelo-
cytes, 20% or more basophils, or platelet counts lower than
100 10
9
/L unrelated to therapy.
17
Of note, most classifica-
tion systems for CML-AP and CML-BP proposed through-
out the years in the literature were developed before the
introduction of imatinib, and some, such as the recent
World Health Organization proposal,
13
have not been vali-
dated in clinical studies and therefore may lack clinical
importance.
18
Most discrepancies among different classi-
fications relate to the use of slightly different cutoffs for
hematologic values and other additional minor criteria.
Although patients may become symptomatic, the transition
from CML-CP to CML-AP is usually subclinical, and labo-
ratory monitoring is necessary for detection of disease
progression. The median survival for patients with CML-
AP is 1 to 2 years.
19
Most patients CML will remain in AP
for 4 to 6 months before progressing to BP.
20,21
Classic
criteria define CML-BP by the demonstration of at least
30% of blasts in the peripheral blood or the bone marrow or
the presence of extramedullary blastic foci. The World
Health Organization classification proposes a reduction
from 30% or more to 20% or more blasts for the diagnosis
of CML-BP. Clinically, most patients with CML-BP expe-
rience signs and symptoms related to increasing tumor
burden, including inability to control WBC counts with
previously stable doses of medication, marked constitu-
tional symptoms (fever, night sweats, anorexia, malaise,
weight loss), splenic infarcts due to massive splenomegaly,
bone pain, and increased risk of infections and bleeding.
The presence of cytogenetic abnormalities in addition to
the Ph chromosome (ie, clonal evolution) involves most
frequently trisomy 8, isochromosome 17, and duplicate Ph
chromosome and has been considered a criterion of CML-
AP. However, patients who are classified as having CML-
AP based solely on the presence of clonal evolution appear
to fare better than those whose conditions were diagnosed
on the basis of other clinical features. Immunopheno-
typically, CML-BP can be characterized by myeloid or
lymphoid phenotype. In rare instances, blast cells can be
biphenotypic with a mixed lymphoblastic-myeloblastic
lineage.
19,22
Lymphoid CML-BP occurs in 20% to 30%
of patients, myeloid in 50%, and undifferentiated in
25%.
22,23
Median survival of patients with lymphoid CML-
BP is 3 to 6 months, with patients with lymphoid CML-BP
having a slightly better prognosis than those with myeloid
phenotype.
TABLE 1. Cri t eri a for Accel erat ed Phase Accordi ng t o 3
Frequent l y Used Cl assi fi cat i ons*
MDACC
11
IBMTR
12
WHO
13
Blasts (%) 15 10 10-19
Blasts and
promyelocytes (%) 30 20 NA
Basophils (%) 20 20 20
Platelets ( 10
9
/L) <100 Unresponsive increase or <100 or >1000 unresponsive
persistent decrease to treatment
Cytogenetics CE CE CE not at the time of diagnosis
White blood cell NA Difficult to control or NA
doubling in <5 d
Anemia NA Unresponsive NA
Splenomegaly NA Increasing NA
Other NA Chloromas, myelofibrosis Megakaryocyte proliferation,
fibrosis
*CE = clonal evolution; IBMTR = International Bone Marrow Transplant Registry; MDACC = M. D. Ander-
son Cancer Center; NA = not applicable; WHO = World Health Organization.
Blast phase of 20% or more blasts (!30% for MDACC and IBMTR).
Basophils and eosinophils.
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 975
CHRONIC MYELOID LEUKEMIA
LABORATORY AND PATHOLOGICAL FEATURES
Laboratory findings in CML include leukocytosis with a
remarkable left shift, basophilia, and eosinophilia. Platelet
count may be either high or low, and mild anemia is com-
monly observed. The leukocyte alkaline phosphatase ac-
tivity is reduced, although phagocytic function remains
essentially normal.
24
Low values may also be observed in
agnogenic myeloid metaplasia, whereas activity may in-
crease with infection, clinical remission, or at the onset of
BP. The increase in the size of the WBC pool is reflected by
marked elevation of serum B
12
levels and unsaturated B
12
-
binding capacity. During transformation to CML-BP, cir-
culating basophil levels, histamine levels, or both often
increase, but the causative factors and importance of these
phenomena remain unknown.
25
The bone marrow of patients with CML is notoriously
hypercellular and devoid of fat.
26
In fact, it can be hypoth-
esized that the primary biologic defect in CML is not
unregulated proliferation of leukemic stem cells but rather
discordant maturation, wherein a slight delay in cell matu-
ration within the myeloid compartment results in increased
myeloid mass.
27
All stages of myeloid maturation are
present, with predominance of myelocytes. In CP, the sum
of myeloblasts and promyelocytes usually accounts for less
than 10% of the marrow cellularity. A decrease in apopto-
sis of myeloid cells contributes to the expansion of this
compartment and defective adherence of immature CML
cells to marrow stromal cells in vitro,
28
with subsequent
early release into the circulation. Megakaryocytes may be
increased, and Gaucher-like cells can be observed in 10% of
cases.
29
As CML progresses, varying degrees of reticulin
fibrosis may be seen,
30
likely a result of the interaction
between the proliferative clone of megakaryocytes and
cytokines, such as platelet-derived growth factor, transform-
ing growth factor , and basic fibroblastic growth factor,
whose plasma levels are significantly increased in CML.
31
Of note, these cytokines play an important role modulating
angiogenesis, and characteristically the bone marrow from
patients with CML shows the highest number of blood ves-
sels and largest vascular area of all leukemias.
31
Blastic cells from patients with lymphoid CML-BP ex-
press terminal deoxynucleotidyl transferase, an enzyme
that catalyzes the polymerization of deoxynucleoside tri-
phosphates and is mainly found in poorly differentiated B
and T cells.
32
In most cases of lymphoid CML-BP, blasts
exhibit a B-cell immunophenotype, expressing CD10,
CD19, and CD22. In contrast, myeloid CML-BP closely
resembles acute myeloid leukemia (AML), and blasts stain
with myeloperoxidase and express myeloid markers such
as CD13, CD33, and CD117. Frequently, myeloid markers
such as CD13 and CD33 may be expressed in patients with
lymphoid CML-BP. Rare cases of megakaryoblastic,
33
erythroblastic,
34
and mastoblastic
35
transformation have
been reported.
PROGNOSTIC SYSTEMS
Because of the difference in clinical aggressiveness be-
tween the initial CP and advanced phases of CML and
because the prognosis in CML can vary significantly even
among patients in the same phase of the disease, several
risk stratification strategies have been developed in an
attempt to prognosticate and guide patient management.
The system most extensively used is the Sokal score, which
was derived from 813 patients with CML collected from 6
European and American series in the 1960s and 1970s.
36
In
this system, the hazard ratio for death is calculated from
baseline patient and disease characteristics using the fol-
lowing formula:
i
(+)/
o
(t) = Exp 0.0116 (age 43.4) +
0.0345 (spleen 7.51) + 0.188 [(platelets/700)
2
0.563] +
0.0887 (blasts 2.10). This scoring system establishes 3
prognostic groups with hazard ratios of less than 0.8, 0.8 to
1.2, and more than 1.2 and median survivals of approxi-
mately 2.5, 3.5, and 4.5 years, respectively.
36
Of note, most
of the patients from whom the Sokal score was derived
were treated with therapies currently not in use, such as
busulfan, intensive chemotherapy, and splenectomy. How-
ever, several studies performed in the 1990s have demon-
strated the applicability of this system in more modern
series. Of 625 patients with CML-CP treated with chemo-
therapy in one study, 168 (27%) were high risk and had a 2-
year actuarial survival of 70% and a subsequent probability
of death of approximately 30% per year. In contrast, the
low-risk group had a 2-year actuarial survival of 91% and a
subsequent risk of death of approximately 17% per year.
37
The Sokal system has several limitations. First, in asymp-
tomatic patients whose disease is detectable only by mo-
lecular testing, lead-time bias can be substantial relative to
those whose conditions are diagnosed after presenting with
symptoms. Second, the use of more refined and potent
therapeutic modalities in CML, such as imatinib, has been
associated with the loss of predictive value of some prog-
nostic factors.
38
Still, the Sokal score identifies patients
with significantly different probabilities of response to
imatinib, although patients with high-risk scores now have
a much better outcome.
Other scoring systems have been proposed that may be
more applicable to patients treated with interferon alfa. The
best-characterized and most widely used is the Hasford
score,
39
which considers patient age, platelet count, periph-
eral blast count, spleen size, eosinophils, and basophils to
determine risk. Some analyses have suggested that this
scoring system is less predictive among patients treated
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 976
CHRONIC MYELOID LEUKEMIA
with imatinib, but others have found it still useful. Finally,
the Gratwohl score system has been developed specifically
for patients undergoing SCT.
40
CYTOGENETICS
A conclusive diagnosis of CML relies on cytogenetic and
molecular testing to identify the t(9;22)(q34.1;q11.21) and/
or the BCR-ABL1 hybrid gene, which is pathognomonic of
this disease.
4,5
The Ph chromosome results from the bal-
anced translocation of cytogenetic material after a break on
chromosome 9 at band q34.1 that transposes the 3 segment
of the ABL1 gene to the 5 segment of the BCR gene on
chromosome 22 at band q11.21.
5
The Ph chromosome is
detected in 95% of patients with CML and in 5% of chil-
dren and 15% to 30% of adults with acute lymphoblastic
leukemia (ALL). In addition, approximately 2% of patients
with newly diagnosed AML can present with this cytoge-
netic abnormality.
41
Occasionally, translocations that in-
volve 3 or more chromosomes occur and are termed variant
Ph chromosome translocations.
42,43
Although initially
thought to confer poor prognosis,
44
recent data in patients
treated with imatinib have demonstrated that variant trans-
locations are associated with a similar prognosis compared
with patients with the classic Ph chromosome.
45
In 5% of
patients with typical CML phenotype, the Ph chromosome
is not detectable but harbors the BCR-ABL1 rearrange-
ment.
46
These patients have the same biology and outcome
as those who express the Ph chromosome. Among the 5%
to 10% of patients with CML clinical features who do not
express the Ph chromosome, 30% to 50% have molecular
evidence of the rearrangement involving BCR and ABL1.
Those who do not express this molecular abnormality (ie,
true Ph chromosome negative) are called atypical CML and
have a different prognosis and treatment.
47
The characteris-
tics and management of atypical CML are not covered any
further in this article.
Little is known about the mechanisms involved in trans-
formation to CML-BP, but the acquisition of additional
cytogenetic abnormalities, referred to as clonal evolution,
in Ph chromosomepositive cells has been thought to play
a prime role in CML progression. The most frequent sec-
ondary cytogenetic abnormalities encountered in patients
with clonal evolution are trisomy 8, isochromosome 17,
and duplicate Ph chromosome, which have been linked
to c-Myc overexpression, loss of 17p, and BCR-ABL1
overexpression, respectively.
48-50
Other cytogenetic aber-
rancies, such as trisomy 19, trisomy 21, trisomy 17, and
deletion 7, have been implicated in less than 10% of cases
of clonal evolution.
51,52
These genetic lesions are more
frequently associated with myeloid than with lymphoid
CML-BP. Nonetheless, a clear correlation between the de-
velopment of these molecular features and disease progres-
sion has not been established yet. In addition, other mo-
lecular events, such as deletion or inactivation of p53,
53
p16,
54
and the retinoblastoma gene product
55
have been
associated with disease transformation, but like overex-
pression of EVI-1,
56
these are relatively infrequent and
therefore not specific to CML-BP.
57
Recently, it was sug-
gested that progression of CML to BP probably involves
several events in primitive progenitors, including BCR-
ABL1 amplification,

acquisition of resistance to apoptosis,
genomic instability,

escape from innate and adaptive im-
mune responses, and activation

of -catenin in granulo-
cyte-macrophage progenitors, resulting

in the acquisition
of self-renewal capacity.
58
Gene methylation has also been
associated with progression in CML.
59-61
The Pa promoter
of ABL1 and the p15 promoter are hypermethylated in 81%
and 24% of patients, respectively. Interestingly, hyper-
methylation of p15, but not of the Pa promoter, is associ-
ated with disease progression. Methylation of the cadherin-
13 gene occurs at an early stage in CML and is associated
with high-risk features and lower probability of response to
interferon alfa.
62
Cytogenetic aberrancies have been reported occasion-
ally in Ph chromosomenegative cells during treatment
with interferon alfa and, more recently, with imatinib.
These aberrancies do not represent clonal evolution be-
cause they occur in cells without the Ph chromosome and
should not be considered as part of the definition of AP.
The importance of these events is currently unknown.
63
The
most common such cytogenetic abnormalities include tri-
somy 8, monosomy 5 or 7, and deletion 20q.
64-66
These
abnormalities frequently regress spontaneously, and it has
been suggested that patients with these abnormalities have
similar long-term outcome as patients without such abnor-
malities. However, occasional cases of progression to
AML or myelodysplastic syndrome have been reported.
MOLECULAR BIOLOGY OF BCR-ABL1
The ABL1 gene is the human homologue of the v-abl
oncogene harbored by the Abelson murine leukemia vi-
rus,
67
and it encodes a nonreceptor tyrosine kinase (Figure
1).
68
In turn, BCR encodes a protein with serinethreonine
kinase activity. The fusion of these 2 genes results in the
activation of the c-ABL protooncogene to its oncogenic
form.
69,70
The breakpoints within the ABL1 gene at 9q34
map to an area spanning more than 300 kilobase (kb) at its
5 end occur upstream of exon Ib, downstream of exon Ia,
or, more frequently, between the two.
71
Alternatively,
breakpoints within BCR localize to 3 main breakpoint clus-
ter regions. In most patients with CML and a third of those
with Ph chromosomepositive ALL, the break occurs
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 977
CHRONIC MYELOID LEUKEMIA
within a 5.8-kb area spanning BCR exons e12-e16 (for-
merly called b1-b5), referred to as the major breakpoint
cluster region (M-bcr). Because of alternative splicing,
a fusion transcript with either b2a2 or b3a2 junctions
can result and will give rise to a 210-kd hybrid protein
(p210
BCR-ABL
).
72,73
The clinical features and response to
treatment are apparently similar for both transcripts. In
two thirds of patients with Ph chromosomepositive ALL
and rarely in CML, the breakpoints within BCR localize
to an area of 54.4 kb between exons e2 and e2, termed the
minor breakpoint cluster region (m-bcr), giving rise to an
e1a2 messenger RNA that is translated to a 190-kd pro-
tein (p190
BCR-ABL
). Patients with CML who carry this tran-
script can present with marked monocytosis and may
have a worse prognosis than those with the typical p210.
74
A third breakpoint cluster region (-bcr) downstream of
exon 19 has been identified, giving rise to a 230-kd fusion
protein (p230
BCR-ABL
) linked to cases of chronic neutrophilic
leukemia.
75
Although all 3 transcripts can induce a CML-like
disorder in mice, p230
BCR-ABL
has reduced tyrosine kinase
activity relative to p190
BCR-ABL76
and confers only partial
growth factor independence, which parallels the milder clini-
cal course of chronic neutrophilic leukemia.
75
All the
isoforms of Bcr-Abl have an active and an inactive configu-
ration. Interestingly, using reverse transcriptasepolymerase
chain reaction (RT-PCR) with a sensitivity of approximately
10
8
, the BCR-ABL1 chimerism has been identified in up to
25% to 30% of presumably healthy adults.
77,78
One hypoth-
esis to explain this phenomenon is that BCR-ABL1 might not
be the sole genetic lesion necessary for the development of
CML. JunB is a component of the activator protein 1 family
of transcription factors
79
Modifications of JunB expression
may have a role in the pathogenesis of CML by modulating
the initiation, progression, and maintenance of the myeloid
differentiation program.
80
A recent report in which trans-
genic mice specifically lacking JunB in the myeloid lineage
(JunB
/
Ubi-JunB mice) develop a myeloproliferative disor-
der that resembles human CML, including progression to
BP, seems to support the latter hypothesis.
81
Several biological models, such as BCR-ABL1express-
ing CD34
+
cells in culture
82,83
or retrovirally transduced
BCR-ABL1
+
mouse cells,
84
have demonstrated that BCR-
ABL1 is an oncogene that promotes CML pathogenesis.
Numerous pathways are activated in BCR-ABL1express-
ing cells,
85
and phosphorylation at the Y177 site of BCR is
essential for BCR-ABL1 leukemogenesis.
86,87
This residue
constitutes a high-affinity docking site for the SH2 domain
of growth factor receptorbound protein 2 (Grb2). In turn,
Grb2 recruits SOS (a guanine-nucleotide exchanger of
RAS) and the scaffold adapter Grb2-associated binding
protein 2 (GAB2) through its SH3 domain. Phosphoryla-
tion of GAB2 leads to recruitment of phosphatidylinositol
3-kinase (PI3K) and SHP2 (also known as PTPN11),
whereas SOS activates RAS.
88
Mutation of Y177 to phenyl-
alanine (Y177F) largely abolishes Grb2 binding and abro-
gates Bcr-Ablinduced RAS activation. Therefore, despite
FIGURE 1. Activation of signal transduction pathways by BCR/ ABL. AKT =protein kinase B; ERK =
extracellular signal-regulated kinase; J AK =J anus kinase; MEK =mitogen-activated protein kinase
kinase; PI3K =phosphatidylinositol 3-kinase; SAPK =stress-activated protein kinase; STAT =
signal transducer and activator of transcription.
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 978
CHRONIC MYELOID LEUKEMIA
preservation of the kinase activity of Abl, the Y177F mutant
impedes the transformation of primary bone marrow cul-
tures.
86
In an SCT model of CML, BCR-ABL1
Y177F
has a
greatly reduced ability to induce a myeloproliferative dis-
order in mice.
89
Of note, SHP2 is required for normal
activation of the RAS extracellular signal-regulated ki-
nase pathway that most receptor tyrosine kinases signal
through.
88
Bcr-Abl also phosphorylates the Src kinases
Lyn, Hck, and Fgr. Once phosphorylated, Hck activates
signal transducer and activator of transcription (STAT)
5.
90,91
Importantly, overexpression of Hck/Lyn has been
linked with disease progression, particularly with a lym-
phoid phenotype.
92
Therefore, Bcr-Abl promotes hemato-
poietic cell transformation mainly through RAS, PI3K, and
STAT 5 activation. All these elements have been shown to
up-regulate the expression of cyclin D1, thus inducing cell-
cycle progression from G
1
to S phase.
93,94
When BCR-
ABL1positive K562 cells were induced to express domi-
nant negative forms of RAS, PI3K, or STAT 5, marked
apoptosis was observed in cells coexpressing 2 of 3 domi-
nant negative mutants in any combination.
95
Notably, Bcr-
Abl also enhances SET expression, particularly during pro-
gression to BP.
96
SET is a nucleus/cytoplasm-localized
phosphoprotein that potently inhibits the tumor suppressor
protein phosphatase 2A (PP2A). In turn, PP2A is a phos-
phatase that regulates cell proliferation, survival, and dif-
ferentiation.
97
Some of the elements involved in Bcr-Abl
signal transduction have been sought as therapeutic targets
in CML.
TREATMENT
The treatment of CML has experienced a great degree of
refinement during the past few years. Until recently, inter-
feron alfa and SCT were the only therapeutic options for
patients with CML. Currently, SCT still remains a valid
option, particularly for patients who do not respond appro-
priately to TKIs, and still remains a potentially curative
therapeutic modality. However, the explosion of the field of
molecularly targeted TKIs in recent years led by imatinib
mesylate has brought about a change in the natural history of
the disease and in the therapeutic approach to CML.
INTERFERON
Recombinant human interferon alfa has shown significant
antitumor and immunomodulatory activity in the treatment
of CML.
98,99
Major cytogenetic responses (MCGRs) (<35%
Ph chromosomepositive cells) have been reported in up to
40% of patients and complete cytogenetic responses
(CCGRs) (0% Ph chromosomepositive cells) in up to
25%.
100
The addition of cytarabine resulted in a significantly
higher probability of a CCGR in up to 35% of patients.
101-105
The achievement of a CCGR is associated with improved
survival, with 78% of patients who achieved a CCGR alive
at 10 years, establishing the achievement of a CCGR as the
therapeutic goal in the interferon alfa era.
106
Furthermore,
approximately 30% of patients in CCGR also had undetect-
able disease by PCR (ie, complete molecular remission);
none of them has relapsed after a median follow-up of 10
years and are therefore probably cured.
106,107
Interestingly,
40% to 60% of those in CCGR with the presence of mini-
mal residual disease at the molecular level have not re-
lapsed after 10 years. This has been attributed to interferon
alfainduced immune modulation, such as the presence of
cytotoxic T lymphocytes specific for PR1, a peptide de-
rived from proteinase 3, which is overexpressed in CML
cells. This phenomenon has been demonstrated in patients
in complete remission after interferon alfa therapy and SCT
but not in those who fail to achieve CCGR or those treated
with chemotherapy.
108
STEM CELL TRANSPLANTATION
Stem cell transplantation remains an important treatment
option for patients with CML, particularly younger indi-
viduals with HLA-identical siblings. Initially, SCT was
thought to render the best outcomes when performed
within 12 months from CML diagnosis. It has been re-
ported that SCT performed within the first 24 months, and
in some reports within the first 36 months, from diagnosis
may have similar outcomes.
40
Although it was initially
suggested that prior interferon alfa therapy adversely af-
fected the outcome after SCT,
109
subsequent trials demon-
strated that this adverse effect was nonexistent,
110-113
and
current evidence suggests that prior exposure to imatinib
does not affect the outcome either.
114,115
Current registry
data revealed that the post-SCT mortality within the first
year is 10% to 20% even in patients who underwent trans-
plantation in optimal conditions (ie, HLA-matched sibling,
first CP <1 year from diagnosis, age <40 years). In addi-
tion, only one third to one half of patients may have a
suitable HLA-matched sibling, and stringent inclusion cri-
teria, such as age, adequate organ function, and perfor-
mance status, may preclude the possibility of SCT in a
large proportion of patients. A strategy to overcome the
lack of suitable donors is the use of matched unrelated
donors (MUDs).
116
Unfortunately, the survival rate of pa-
tients undergoing MUD SCT is still lower when compared
with those receiving grafts from HLA-matched siblings.
The outcome of MUD SCT may be improved by careful
molecular typing, particularly for HLA-A, HLA-B, and
HLA-DRB1.
116
However, the use of stringent typing crite-
ria is inevitably linked to a lower probability of finding an
adequate donor. In addition, SCT is fraught with risks such
as graft-vs-host disease, veno-occlusive disease, life-
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 979
CHRONIC MYELOID LEUKEMIA
threatening infections, risk of secondary malignancy, and
poorer overall quality of life.
117
Also, the risk of late relapse
is being increasingly recognized. A recent series analyzed
the long-term follow-up of 89 patients who underwent
transplantation at a single institution.
118
After more than 10
years of follow-up, 28 patients (31%) were alive. The mean
time to hematologic or cytogenetic relapse was 7.7 years,
with 5 patients relapsing more than 10 years after SCT. In a
recent study by the Seattle group, 131 patients with early
CML-CP with a median age of 43 years (range, 14-66
years) received targeted busulfan and cyclophosphamide as
a conditioning regimen. At 3 years, the projected non-
relapse mortality rate was 14%, and 78% were projected to
be alive and free of disease. However, 60% of survivors
had extensive chronic graft-vs-host disease 1 year after
transplantation, although only 10% had a Karnofsky score
less than 80%. Notably, 11% of patients had minimal re-
sidual disease documented by PCR but had not relapsed at
the time of the report.
119
In contrast, patients with advanced
phase CML who undergo SCT have a 5-year survival prob-
ability of 40% to 60% in the AP and 10% to 20% in the BP.
Notably, the long-term outcome of patients with CML-BP
who achieve a second CP and who are undergoing SCT
appears to be similar to that of patients with CML-AP who
undergo transplantation.
The long-term disease-free survival rate after MUD
SCT for young patients in the early CP stage of disease is
57% compared with 67% in patients receiving grafts from
HLA-matched siblings.
120
The incorporation of molecular
matching in recent years has decreased the rate of graft-vs-
host disease and improved the probability of long-term
survival in patients undergoing MUD SCT.
Reduced-intensity conditioning regimens have been
used as a means to make SCT available to a broader num-
ber of patients who otherwise would be ineligible for stan-
dard SCT. This approach takes advantage of the ability of
the T cells harbored in the graft to eliminate the CML cells
(graft-vs-leukemia effect) as opposed to obtaining this ef-
fect by using ablative cytotoxic chemotherapy. This strat-
egy increases tolerability and decreases morbidity and
mortality significantly. Long-term results with reduced-
intensity conditioning regimens are not yet available, but
early results in small series (frequently including still
mostly young patients) report disease-free survival as high
as 85% at 70 months.
121
In contrast, a recent study from the
European Bone Marrow Transplant Registry reports a 3-
year overall survival rate of 58% and a progression-free
survival rate of 37%.
122
Patients with minimal residual
disease by PCR 12 months after SCT have a risk of relapse
of 30% to 40% compared with less than 5% for those with
negative PCR results.
123
However, relapse can be fre-
quently managed successfully with donor lymphocyte infu-
sion in up to 70% of patients who relapse in the CP,
particularly when administered at the time of molecular
relapse. Alternatively, imatinib can be used in the post-
SCT relapse setting and induces CCGR in more than 40%
of patients treated in the CP.
124
IMATINIB MESYLATE
Imatinib mesylate (Gleevec) is an orally bioavailable 2-
phenylaminopyrimidine with targeted inhibitory activity
against the constitutively active tyrosine kinase of the Bcr-
Abl chimeric fusion protein. In addition, imatinib inhibits
other kinases, such as c-Kit, platelet-derived growth factor
and , and Abl-related gene (ARG).
125,126
Imatinib has
become the standard therapy for CML because of its re-
markable activity and mild toxicity profile. In a phase 1
dose-finding study in patients who were intolerant to or in
whom interferon alfabased therapy failed, daily doses of
25 to 1000 mg were investigated. Responses among pa-
tients receiving doses lower than 250 mg/d were rare, in
contrast to 98% of patients receiving a dose of at least 300
mg/d who achieved a complete hematologic response
(CHR), including 54% who attained a cytogenetic re-
sponse. Of note, no dose-limiting toxicity was identified,
and the maximum tolerated dose was not reached.
127
A
dosage of 400 mg/d was selected for a subsequent phase 2
study, including 454 patients with late CP disease who
were intolerant to or in whom interferon alfa therapy
failed.
128
Recently updated data showed that 96% of pa-
tients had achieved a CHR and 55% a CCGR. After a
median follow-up of 5 years, the overall survival rate was
79%.
128
This rate is in keeping with results from a recent
single-institution study in which among 261 patients
treated, 73% obtained a MCGR and 63% a CCGR.
129
After
a median follow-up of 45 months, 86% of patients are
alive, of whom 80% are free of progression. More than
90% of patients who achieved a CCGR maintained a major
response. The BCR-ABL1/ABL1 ratio by nested PCR was
less than 0.05% in 31% and undetectable in 15% of pa-
tients.
129
In the prospective, multicenter phase 3 Interna-
tional Randomized Study of

Interferon and STI571, the
efficacy of imatinib was compared with the combination of
interferon alfa and low-dose cytarabine in patients with
newly diagnosed CML-CP.
130
Patients treated with
imatinib

(n=553) or interferon alfa and low-dose cytarabine
(n=553) were allowed to cross over to the alternative group
if

treatment failure or intolerance

was demonstrated.
131
Imatinib therapy was superior to interferon alfa plus low-
dose cytarabine regarding hematologic and cytogenetic re-
sponses,

tolerability, and transformation to CML-AP. After
a median follow-up of 54 months, the CHR, MCGR, and
CCGR rates were 98%, 92%, and 86%, respectively.
132
Progression-free survival is estimated to be 84%, and only
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 980
CHRONIC MYELOID LEUKEMIA
6% of patients have progressed to AP or BP. The overall
annual progression rate has declined to 1.5% in the fourth
year of therapy, compared with 4.8% and 7.5% in the
previous 2 years, suggesting that disease progression may
plateau in the ensuing years. The depth of the response after
12 months of imatinib therapy has important implications
regarding disease progression and relapse. In fact, in 97%
of patients who had a 3-log or greater reduction in BCR-
ABL transcript levels (ie, major molecular response
[MMR]) at 12 months, the

probability of remaining free of
progression to AP or BP was

97% at 54 months compared
with 89% for patients in CCGR but not in

MMR and 72%
for patients who did not achieve a CCGR at 12 months
(P=.001).
132,133
In addition, the median BCR-ABL levels
continue to decrease, with a mean log reduction at 48
months of 3.4 compared with 3.0 at 12 months.
134
Although
this study did not document a survival advantage compared
with the interferon alfa arm because of the crossover de-
sign, studies comparing survival of patients treated with
imatinib with historical cohorts treated with interferon
alfabased therapy demonstrate the anticipated survival
advantage
135,136
(Figure 2). These results suggest that a trial
of imatinib must be pursued in every patient before consid-
ering SCT.
The results from the International Randomized Study of
Interferon and STI571 have established imatinib 400 mg/d,
as the standard therapy for CML. However, because the
imatinib dosage selected after the dose-finding phase 1
study was based primarily on the excellent responses ob-
tained when doses higher than 300 mg/d were used, it has
been suggested that imatinib at higher dosages (600-800
mg/d) might result in improved outcomes (Table 2). In one
study, 36 patients with CML-CP in whom interferon alfa
therapy failed were treated with imatinib 400 mg twice
daily, resulting in a CCGR rate of 90%, compared with
48% in historical matched controls treated with standard-
dose therapy.
141
In addition, 56% of patients had an MMR,
including 41% with undetectable levels. The toxicity pro-
file was similar to that of imatinib, 400 mg, although there
was a higher rate of grade 3 or higher myelosuppression.
Several single-arm, phase 2 studies have used high-dose
imatinib as the starting dose for patients with previously
untreated CML-CP. These studies have documented a
higher rate of CCGR of up to 95% and higher rates of
molecular responses, particularly responses at the level of a
4-log or greater reduction in transcript levels. In addition,
responses occur significantly earlier with high-dose com-
pared with standard-dose therapy.
137-139
One study reported
that by multivariate analysis, only treatment with high-dose
imatinib (P=.02) was associated with achievement of an
MMR and that achieving an MMR within the first 12
months of therapy is predictive of a durable cytogenetic
remission.
142
Similarly, patients receiving the highest dose
intensity during the first year of therapy have the highest
probability of achieving a major molecular remission at 24
months (89%), whereas those receiving a median daily
FIGURE 2. Survival of patients with early chronic phase chronic myeloid leukemia treated at M. D.
Anderson Cancer Center in different eras compared with those treated with imatinib.
Year Total Dead
Imatinib 230 7
1900-2000 960 334
1982-1989 365 265
1975-1981 132 127
1965-1974 123 122
Time from referral (y)
0 2 4 6 8 10 12 14 16
1.0
0.8
0.6
0.4
0.2
0.0
P
r
o
p
o
r
t
i
o
n

s
u
r
v
i
v
i
n
g
95%
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 981
CHRONIC MYELOID LEUKEMIA
dosage less than 600 mg had a significantly lower probabil-
ity both at 12 and at 24 months.
143
Thus, these studies
suggest that high-dose imatinib may provide more and
better remissions, but the results from ongoing randomized
studies will determine whether high-dose imatinib provides
only faster responses or whether these translate into pro-
longed progression-free and/or overall survival.
Duration of Imatinib Therapy. Currently, no evidence
exists to support the belief that patients taking imatinib can
safely discontinue therapy once they achieve a complete
molecular response. Most patients who have discontinued
imatinib therapy have rapidly experienced both molecular
and cytogenetic relapse, even when some had sustained un-
detectable levels of BCR-ABL for significant periods.
144-146
Recently, Rousselot et al
147
described 8 patients who dis-
continued use of imatinib after maintaining undetectable
BCR-ABL levels for 24 months, 4 of whom still tested
PCR negative after a follow-up of 8, 9, 12, and 13 months,
respectively. All these patients had been treated with inter-
feron alfa, suggesting that interferon alfa immunomodula-
tory effects may account for sustained and prolonged mo-
lecular responses observed on therapy discontinuation.
It has been suggested that the most primitive quiescent
leukemic progenitors are insensitive to imatinib in vitro
and are responsible for CML relapse once the inhibitory
pressure of imatinib is removed.
148
Emerging data suggest
that resistance of quiescent cells to imatinib is indeed a
Bcr-Abldependent phenomenon.
149
Other potential
mechanisms implicated in quiescent CML progenitors in-
clude expression of drug transport proteins such as PgP and
Abcg2, atypical kinase domain mutations,
150
and BCR-
ABL overexpression.
151
Monitoring of Imatinib Therapy and Minimal Re-
sidual Disease. The significant MMR rate observed with
imatinib and its prognostic implications
152,153
have led to the
redefinition of the therapeutic goals in CML. Rather than
aiming solely at achieving a CCGR,
106,154
modern CML
therapy should aim for the achievement of molecular re-
sponses. The long-term importance of molecular monitor-
ing in CML has been best exemplified by patients under-
going SCT, in whom detection of BCR-ABL transcripts,
particularly at high levels and 6 months after transplanta-
tion, confer a significantly higher risk of relapse. In these
instances, prompt intervention (eg, donor lymphocyte infu-
sion) has been effective. As mentioned earlier, among pa-
tients who achieve a CCGR with interferon alfa, the risk of
relapse is minimal for those with low transcript levels.
107
Patients who achieve undetectable BCR-ABL transcripts
by nested PCR have all had sustained cytogenetic remis-
sions beyond 10 years.
106
In the imatinib setting, once a
CCGR is attained, quantitative RT-PCR in peripheral
blood samples at 3-month intervals is the mainstay of
monitoring for most patients. However, the lack of consis-
tency in reporting BCR-ABL transcript levels has been a
source of intense debate. A recent proposal aimed at the
standardization of RT-PCR techniques and result reporting
will greatly help this objective. According to this proposal,
results will be converted by comparing analysis of standard-
ized reference samples with a value of 0.1% corresponding
to MMR in all laboratories.
155
After the patient has achieved
a complete cytogenetic remission, RT-PCR should be per-
formed every 3 months and a routine cytogenetic analysis
every 12 months.
155
Fluorescence in situ hybridization in
peripheral blood specimens can be used to monitor the cyto-
genetic responses among cytogenetic analyses. In addition,
approximately 5% of patients who obtain a cytogenetic re-
sponse develop clonal karyotypic abnormalities in Ph chro-
mosomenegative cells, primarily consistent with those en-
countered in myelodysplastic syndromes, although only rare
cases of myelodysplastic syndrome or AML have been re-
ported.
63
Although the long-term implications of these ab-
normalities are unknown, the risk of transformation to AML
warrants careful follow-up of these patients.
Imatinib Resistance. A subset of patients with CML
will exhibit either primary or secondary resistance to
imatinib. Primary resistance refers to patients never re-
sponding to imatinib, whereas secondary resistance occurs
when a patient who had an initial response to imatinib
eventually loses the response. Among patients treated in
CP, the rate of resistance has been estimated to be less than
2% per year. Several mechanisms of resistance have been
reported, many of them mostly in vitro or in selected pa-
tient samples, and their individual contribution to this phe-
nomenon has not been completely defined. The most fre-
quently identified mechanism of resistance is the develop-
ment of mutations in the Abl tyrosine kinase domain. More
than 40 different BCR-ABL point mutations have been
reported thus far, and new ones continue to be described.
156
The region of the Abl kinase where mutations occur is
TABLE 2. Summary of St udi es Usi ng Hi gh-Dose Imat i ni b
as Front l i ne Therapy for Earl y Chroni c Phase
Chroni c Myel oi d Leukemi a*
Dosage No. of CGCR MMR CMR
Study (mg/d) patients (%) (%) (%)
MDACC
137
800 175 90 61 32
TIDEL
138
600800 41 90 44 25
RIGHT
139
800 114 85 64 44
GIMEMA
140
800 100 88 60 NA
*CGCR = complete cytogenetic remission; CMR = complete molecular
remission (BCR-ABL1 undetectable); GIMEMA = Gruppo Italiano
Malattie Ematologiche dellAdulto; MDACC = M. D. Anderson Cancer
Center; MMR = major molecular response (BCR-ABL1/ABL1 <0.05% or
3-log reduction); RIGHT = Rationale and Insight on Gleevec [imatinib]
High-Dose Therapy; TIDEL = Therapeutic Intensification in de novo
leukemia.
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 982
CHRONIC MYELOID LEUKEMIA
critical in their ability to confer varying degrees of insensi-
tivity to imatinib. In fact, ranges of 220 to more than 5000
nM/L in the inhibitory concentration that results in a 50%
reduction in proliferation (IC
50
) have been described in the
various Bcr-Abl mutants.
157
The most frequent mutations
are those that map to the P-loop region of the kinase do-
main.
157-159
The P-loop region is a glycine-rich structure that
serves as a docking site for phosphate groups of adenosine
triphosphate (ATP). G250E, Q252H, and the highly
imatinib-insensitive Y253F and E255K mutations locate in
this region. Despite not being directly involved in imatinib
binding, these residues have been linked to poor prognosis
in some series, with a median survival of 4 to 5 months.
160-163
Other series have refuted this notion.
164
The difference in
these series may be due to several factors, including patient
selection and the criteria used to select patients for screen-
ing for mutations. In addition, the first studies that sug-
gested a poor outcome for patients with P-loop mutations
did not address the management of patients after develop-
ing mutations, and it is now clear that several treatment
strategies can overcome resistance through mutations.
165
Also, considering that individual mutations have different
levels of resistance to imatinib and other agents, it is more
appropriate to refer to individual mutations rather than
grouping them.
155
The imatinib-insensitive T315I mutation
occurs in a highly conserved gatekeeper threonine resi-
due

near the Bcr-Abl catalytic domain, thus leading to
steric interference with imatinib binding.
161,162,166
Other mu-
tations have been described mapping to the catalytic do-
main and the activation loop of Abl, impairing the ability of
the kinase to adopt a close (inactive) spatial conformation
necessary to bind to imatinib.
167
Despite the poor prognosis conferred by the T315I mu-
tant, in vitro and clinical data suggest that other mutations
may be clinically relevant.
168
Furthermore, other mecha-
nisms of resistance, such as overexpression or amplification
of the BCR-ABL message and/or protein, may influence the
outcome of patients treated with imatinib. Recently, it was
shown that activation of Src kinases and NF-B may play a
role in imatinib resistance in some patients.
92,169
TREATMENT STRATEGIES FOR THE FUTURE
New Bcr-Abl Inhibitors. Several strategies are cur-
rently being developed to overcome imatinib resistance
(Table 3). New Abl TKIs have been developed with
enhanced potency and less stringent binding requirements
and, in some cases, with inhibitory activity against other
kinases involved in mechanisms of resistance to imatinib.
In addition, new agents that block Bcr-Abl in a nonATP-
competitive manner may represent an alternative for
patients who develop imatinib-insensitive Bcr-Abl kinase
mutations.
Nilotinib (AMN107). Nilotinib is a phenylamino-py-
rimidine derivative with a 20- to 30-fold increased potency
compared with imatinib against Bcr-Abl. Crystallographic
models suggest that this increased potency is a result of
TABLE 3. Resul t s Wi t h New Tyrosi ne Ki nase Inhi bi t ors Aft er Fai l ure of Imat i ni b
i n Al l Phases of CML*
No. of
Agent and CML phase patients Overall Complete Overall Complete
Phase 1
Nilotinib
170
CP 17 92 92 53 35
CE 10 100 100 90 20
AP 46 72 46 48 13
MyBP 24 42 8 29 4
LyBP 9 33 22 11
Dasatinib
171,172
CP 40 93 93 63 35
AP 11 81 45 36 18
MyBP 23 61 35 52 26
LyBP 10 80 70 90 30
Phase 2
Dasatinib
CP
173
186 90 90 45 33
AP
174
107 59 33 36 21
MyBP
175
74 51 24 42 27
LyBP and Ph+ ALL
176
78 40 26 54 46
*AP = accelerated phase; CE = clonal evolution; CML = chronic myeloid leukemia; CP = chronic
phase; LyBP = lymphoid blast phase; MyBP = myeloid blast phase; Ph+ ALL = Philadelphia
chromosomepositive acute lymphoblastic leukemia.
Hematologic response
(%)
Cytogenetic response
(%)
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 983
CHRONIC MYELOID LEUKEMIA
nilotinib representing a better topographical fit for the Abl
kinase pocket. However, nilotinib, like imatinib, only binds
the Abl protein in its inactive conformation. Nilotinib
inhibits the tyrosine kinase activity of most clinically
important Bcr-Abl mutants, with the exception of T315I.
157
A phase 1/2 study of nilotinib included 106 patients with
imatinib-resistant CML in all phases and 13 patients with Ph
chromosomepositive ALL.
170
Patients received nilotinib at
dosages ranging from 50 to 1200 mg/d. In CML, hematolog-
ic response rates ranged from 44% to 89%, and cytogenetic
responses ranged from 22% in lymphoid and myeloid BP to
29% in AP and 50% in CP. Among patients with Bcr-Abl
mutations before nilotinib therapy, 60% had a hematologic
response, and 41% had a cytogenetic response. Nilotinib
dosage was escalated up to 1200 mg/d with good tolerance,
and the maximum tolerated dose was estimated at 600 mg
twice daily. The most common drug-related adverse events
were gastrointestinal, rash, and hematologic. Importantly,
greater exposure to the drug and inhibition of Bcr-Abl
phosphorylation were observed with twice-daily dosing
schedules compared with equivalent total doses admin-
istered on a once-daily schedule.
170
Dasatinib (BMS-354825). Because it is possible that
inhibition of Bcr-Abl alone may not be sufficient to eradi-
cate all CML cells, particularly the leukemic imatinib-
insensitive quiescent stem cells, agents with additional in-
hibitory capacity against other kinases have been investi-
gated. Since Src kinases have been implicated in CML
pathogenesis and progression, dual Abl/Src inhibitors may
have enhanced activity compared with imatinib. Among
this new class of compounds, dasatinib has reached the
furthest level of clinical development. Dasatinib is an ATP-
competitive, dual-specific Src- and Abl-kinase inhibitor
with 100-

to 300-fold higher potency against Bcr-Abl com-
pared with imatinib and significant activity against c-kit
(IC
50
,

5 nM), platelet-derived growth factor receptor
(IC
50
,

28 nM), and Hck, Fyn, Src, and Lck kinases (IC
50
,
approximately 0.5 nM).
177
However, dasatinib does not in-
hibit the T315I mutant, even in the presence of micromolar
concentrations.
157
In phase 2 studies, dasatinib was admin-
istered at 70 mg twice daily to patients with CML in all
phases after imatinib failure. A CHR was achieved by 87%,
66%, 55%, and 46% of patients in CP, AP, myeloid BP,
and lymphoid BP/Ph chromosomepositive ALL, respec-
tively. Among patients in CP, 44% had a CCGR, whereas
46% to 66% of patients with advanced-phase CML had
MCGRs.
173-176
In a dose-escalating study, dasatinib ren-
dered 2-log or greater reductions in Bcr-Abl transcripts in
43% of patients in advanced phase, and 37% of patients in
CP achieved

2-log or greater reductions, including 4 pa-
tients in each group who had an MMR.
178
Overall, dasatinib
is well tolerated, with myelosuppression and gastrointesti-
nal toxic effects seen in some patients. Pleural effusions
have also been reported in 10% to 15% of patients, particu-
larly those in BP, and it is usually grade 1 or 2. Despite the
impressive results achieved with dasatinib, it may not
eradicate all quiescent leukemic stem cells, although it may
be more effective at this level than imatinib.
149
Another
dual Abl/Src kinase inhibitor termed SKI-606
179
is current-
ly being evaluated in phase 1 studies, and others, such as
NS-187 (INNO406),
180
AZD0530,
181
PD166326,
182
and
PD180970,
183
may soon follow.
NonATP-Competitive Bcr-Abl Inhibitors. Imatinib,
nilotinib, and dasatinib are ATP-competitive inhibitors and
therefore amenable to being affected in their ability to bind
to the Bcr-Abl kinase by the T315I mutation, which is
considered the gatekeeper of the kinase domain. Com-
pounds that target binding sites unrelated to the ATP-
kinase domain may overcome this problem. ON012380
targets the substrate-binding site of Bcr-Abl, competing
with natural substrates such as Crkl but not with ATP.
184
In
fact, ON012380 causes regression of leukemia induced by
intravenous injection of 32DcI3 cells that express the Bcr-
Abl mutant T315I.
184
BIRB796, a p38 mitogen-activated
protein kinase inhibitor, binds with excellent affinity to the
Bcr-Abl T315I mutant (Kd 40 nM), although high concen-
trations of this compound are necessary to inhibit
autophosphorylation of this mutant in Ba/F3 cells (IC
50
, 1-2
M).
159
In this regard, VX680 (MK-0457), a multikinase
inhibitor that inhibits, among others, aurora A and B ki-
nase, seems to have a more favorable profile against T315I,
although remarkably less affinity for wild-type Bcr-Abl
(IC
50
, >10 M).
159
Other Therapeutic Options. Other approaches are be-
ing developed for patients who develop resistance to
imatinib. Farnesyl transferase inhibitors, such as lonafarnib
and tipifarnib, have demonstrated activity both as single
agents and in combination with imatinib.
185-188
This ap-
proach has rendered anecdotal responses in patients har-
boring the T315I mutant. Homoharringtonine is a Cephalo-
taxus alkaloid that represented the best therapeutic option
in patients in whom interferon alfabased therapies failed
before the introduction of imatinib in clinical practice.
189
Its
potential as salvage therapy in imatinib-resistant CML has
rekindled interest in this drug. Another avenue that has
been pursued in recent years is the use of hypomethylating
agents
190,191
and histone deacetylase inhibitors.
192-194
In a
phase 2 study, decitabine administered at 15 mg/m
2
, 5 days
a week for 2 consecutive weeks, rendered a CHR in 34%
and a CCGR in 46% of patients.
190,191
These and other
agents with different targets are currently being developed
in an attempt to prevent or overcome resistance to imatinib.
Immune-mediated events play an important role in the
suppression of the CML clone, as evidenced by a subset of
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 984
CHRONIC MYELOID LEUKEMIA
patients who maintained a durable CHR after discontinua-
tion of interferon alfa therapy
195
and after allogeneic-SCT
due to a graft-vs-leukemia effect.
196
These observations
have sparked interest in developing immunologic ap-
proaches to treating CML. One such approach is the use of
vaccines to elicit specific immune responses directed to-
ward CML-restricted tumor antigens. Several vaccines are
currently being developed in clinical studies, which en-
compass the use of BCR-ABL junction-spanning se-
quences,
197-199
the nonapeptide PR1 derived from protein-
ase 3,
200
leukocyte-derived HSP70-peptide complexes,
201
and granulocyte-macrophage colony-stimulating factor se-
creting vaccines.
202
The Bcr-Abl breakpoint fusion peptide
vaccine has been already tested in phase 2 clinical stud-
ies.
198,199
Bocchia et al
199
have reported on 16 patients with
stable residual disease after more than 12 months of imatinib
therapy (n=10) or 24 months of interferon alfa therapy (n=6)
who received 6 vaccinations with a peptide vaccine derived
from the sequence p210-b3a2. Five of 9 patients taking
imatinib obtained a CCGR after vaccination, and 3 had
undetectable levels of b3a2 transcripts by RT-PCR. Among
the patients treated with interferon alfa, 2 achieved a CCGR,
and 3 had improvement in the percentage of Ph chromo-
somepositive metaphases. Overall, these data suggest that
this peptide vaccine may have a role in the management of
patients with CML and minimal residual disease.
SUMMARY
Consensus exists that imatinib therapy is the standard ap-
proach to the treatment of CML, and this modality will be
measured in the future against all new therapies. The recent
addition of the highly potent TKIs nilotinib and dasatinib
has further improved the armamentarium against CML but
also posed new challenges. First, the role of these new
compounds in modern algorithms of CML therapy is still
unknown and must be defined. Despite having shown im-
pressive response rates in patients after imatinib failure or
intolerance, the duration of these responses will be deter-
mined with longer follow-up. Their role in frontline
therapy for CML will also have to be defined. Second, as
our resources for controlling CML improve, the refinement
in molecular monitoring must improve in parallel. In this
regard, the lack of consistency among different laboratories
in BCR-ABL transcript results by current PCR technolo-
gies is still an ongoing problem. Future improvements in
molecular techniques and their standardization by using a
laboratory-specific conversion factor will be crucial to fur-
ther our understanding of the dynamics of molecular re-
sponse to TKIs, providing a uniform method and definition
of MMR and defining the concept of molecular relapse.
Other challenges, such as defining the best treatment for
patients who develop the Bcr-Abl T315I mutant, under-
standing the mechanism of resistance of patients without
detectable mutations, the emergence of new mechanisms of
resistance such as new mutations to the new TKI and other
agents, developing novel strategies for the management of
minimal residual disease, and delineating the current role
of SCT, will be the subject of arduous investigation in
future years. Despite all these uncertainties, the treatment
prospects of patients with CML have never been brighter.
REFERENCES
1. Abramson S, Miller RG, Phillips RA. The identification in adult bone
marrow of pluripotent and restricted stem cells of the myeloid and lymphoid
systems. J Exp Med. 1977;145:1567-1579.
2. Bennett J. Case of hypertrophy of the spleen and liver in which death
took place from suppuration of the blood. Edinb Med Surg J. 1845;64:413-423.
3. Dameshek W. Some speculations on the myeloproliferative syndromes.
Blood. 1951;6:372-375.
4. Nowell PC, Hungerford DA. A minute chromosome in human chronic
granulocytic leukemia. Science. 1960;132:1497.
5. Rowley JD. Letter: A new consistent chromosomal abnormality in
chronic myelogenous leukaemia identified by quinacrine fluorescence and
Giemsa staining. Nature. 1973;243:290-293.
6. Bartram CR, de Klein A, Hagemeijer A, et al. Translocation of c-ab1
oncogene correlates with the presence of a Philadelphia chromosome in
chronic myelocytic leukaemia. Nature. 1983;306:277-280.
7. Groffen J, Stephenson JR, Heisterkamp N, de Klein A, Bartram CR,
Grosveld G. Philadelphia chromosomal breakpoints are clustered within a
limited region, bcr, on chromosome 22. Cell. 1984;36:93-99.
8. Jemal A, Tiwari RC, Murray T, et al. Cancer statistics, 2004. CA Cancer
J Clin. 2004;54:8-29.
9. Deininger MW, Goldman JM, Melo JV. The molecular biology of
chronic myeloid leukemia. Blood. 2000;96:3343-3356.
10. Bizzozero OJ Jr, Johnson KG, Ciocco A, Kawasaki S, Toyoda S.
Radiation-related leukemia in Hiroshima and Nagasaki 1946-1964: II. Ann
Intern Med. 1967;66:522-530.
11. Kantarjian HM, Keating MJ, Smith TL, Talpaz M, McCredie KB.
Proposal for a simple synthesis prognostic staging system in chronic myelog-
enous leukemia. Am J Med. 1990;88:1-8.
12. Speck B, Bortin MM, Champlin R, et al. Allogeneic bone-marrow
transplantation for chronic myelogenous leukaemia. Lancet. 1984;1:665-668.
13. Vardiman JW, Harris NL, Brunning RD. The World Health Organization
(WHO) classification of the myeloid neoplasms. Blood. 2002;100:2292-2302.
14. Kantarjian HM, Smith TL, McCredie KB, et al. Chronic myelogenous
leukemia: a multivariate analysis of the associations of patient characteristics
and therapy with survival. Blood. 1985;66:1326-1335.
15. Sokal JE, Baccarani M, Russo D, Tura S. Staging and prognosis in
chronic myelogenous leukemia. Semin Hematol. 1988;25:49-61.
16. Sokal JE, Baccarani M, Tura S, et al. Prognostic discrimination among
younger patients with chronic granulocytic leukemia: relevance to bone mar-
row transplantation. Blood. 1985;66:1352-1357.
17. Kantarjian HM, Dixon D, Keating MJ, et al. Characteristics of acceler-
ated disease in chronic myelogenous leukemia. Cancer. 1988;61:1441-1446.
18. Cortes JE, Talpaz M, OBrien S, et al. Staging of chronic myeloid
leukemia in the imatinib era: an evaluation of the World Health Organization
proposal. Cancer. 2006;106:1306-1315.
19. Cortes J, Kantarjian H. Advanced-phase chronic myeloid leukemia.
Semin Hematol. 2003;40:79-86.
20. Karanas A, Silver RT. Characteristics of the terminal phase of chronic
granulocytic leukemia. Blood. 1968;32:445-459.
21. Griesshammer M, Heinze B, Hellmann A, et al. Chronic myelogenous
leukemia in blast crisis: retrospective analysis of prognostic factors in 90
patients. Ann Hematol. 1996;73:225-230.
22. Kantarjian HM, Keating MJ, Talpaz M, et al. Chronic myelogenous
leukemia in blast crisis: analysis of 242 patients. Am J Med. 1987;83:445-
454.
23. Cortes JE, Talpaz M, Kantarjian H. Chronic myelogenous leukemia: a
review. Am J Med. 1996;100:555-570.
24. Cramer E, Auclair C, Hakim J, et al. Metabolic activity of phagocy-
tosing granulocytes in chronic granulocytic leukemia: ultrastructural observa-
tion of a degranulation defect. Blood. 1977;50:93-106.
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 985
CHRONIC MYELOID LEUKEMIA
25. Denburg JA, Wilson WE, Bienenstock J. Basophil production in myelo-
proliferative disorders: increases during acute blastic transformation of chronic
myeloid leukemia. Blood. 1982;60:113-120.
26. Silver RT. Morphology Of The Blood And Marrow In Clinical Practice.
New York, NY: Grune & Stratton; 1970:84.
27. Strife A, Clarkson B. Biology of chronic myelogenous leukemia: is
discordant maturation the primary defect? Semin Hematol. 1988;25:1-19.
28. Gordon MY, Dowding CR, Riley GP, Goldman JM, Greaves MF.
Altered adhesive interactions with marrow stroma of haematopoietic progeni-
tor cells in chronic myeloid leukaemia. Nature. 1987;328:342-344.
29. Dosik H, Rosner F, Sawitsky A. Acquired lipidosis: Gaucher-like cells
and blue cells in chronic granulocytic leukemia. Semin Hematol. 1972;9:
309-316.
30. Castro-Malaspina H, Moore MA. Pathophysiological mechanisms op-
erating in the development of myelofibrosis: role of megakaryocytes. Nouv Rev
Fr Hematol. 1982;24:221-226.
31. Aguayo A, Kantarjian H, Manshouri T, et al. Angiogenesis in acute and
chronic leukemias and myelodysplastic syndromes. Blood. 2000;96:2240-
2245.
32. Derderian PM, Kantarjian HM, Talpaz M, et al. Chronic myelogenous
leukemia in the lymphoid blastic phase: characteristics, treatment response,
and prognosis. Am J Med. 1993;94:69-74.
33. Bain B, Catovsky D, OBrien M, Spiers AS, Richards GH.
Megakaryoblastic transformation of chronic granulocytic leukaemia. An elec-
tron microscopy and cytochemial study. J Clin Pathol. 1977;30:235-242.
34. Rosenthal S, Cancellos GP, Gralnick HP. Erythroblastic transformation
of chronic granulocytic leukemia. Am J Med. 1977;63:116-124.
35. Soler J, OBrien M, de Castro JT, et al. Blast crisis of chronic granulo-
cytic leukemia with mast cell and basophilic precursors. Am J Clin Pathol.
1985;83:254-259.
36. Sokal JE, Cox EB, Baccarani M, et al. Prognostic discrimination in
good-risk chronic granulocytic leukemia. Blood. 1984;63:789-799.
37. Silver RT, Peterson BL, Szatrowski TP, et al. Treatment of the chronic
phase of chronic myeloid leukemia with an intermittent schedule of recombi-
nant interferon alfa-2b and cytarabine: results from CALGB study 9013. Leuk
Lymphoma. 2003;44:39-48.
38. Quintas-Cardama A, Kantarjian H, Talpaz M, et al. Imatinib mesylate
therapy may overcome the poor prognostic significance of deletions of deriva-
tive chromosome 9 in patients with chronic myelogenous leukemia. Blood.
2005;105:2281-2286.
39. Hasford J, Pfirrmann M, Hehlmann R, et al, Writing Committee for the
Collaborative CML Prognostic Factors Project Group. A new prognostic score
for survival of patients with chronic myeloid leukemia treated with interferon
alfa. J Natl Cancer Inst. 1998;90:850-858.
40. Gratwohl A, Hermans J, Niederwieser D, et al, Chronic Leukemia
Working Party of the European Group for Bone Marrow Transplantation. Bone
marrow transplantation for chronic myeloid leukemia: long-term results. Bone
Marrow Transplant. 1993;12:509-516.
41. Kurzrock R, Gutterman JU, Talpaz M. The molecular genetics of Phila-
delphia chromosome-positive leukemias. N Engl J Med. 1988;319:990-998.
42. Huret JL. Complex translocations, simple variant translocations and Ph-
negative cases in chronic myelogenous leukaemia. Hum Genet. 1990;85:565-
568.
43. De Braekeleer M. Variant Philadelphia translocations in chronic my-
eloid leukemia. Cytogenet Cell Genet. 1987;44:215-222.
44. Albano F, Specchia G, Pannunzio A, et al. Variant t(9;22) translocation
may identify a subgroup of chronic myeloid leukemia patients with poor
prognosis [abstract]. Blood. 2001;98:321a.
45. El-Zimaity MM, Kantarjian H, Talpaz M, et al. Results of imatinib
mesylate therapy in chronic myelogenous leukaemia with variant Philadelphia
chromosome. Br J Haematol. 2004;125:187-195.
46. Shtalrid M, Talpaz M, Blick M, et al. Philadelphia-negative chronic
myelogenous leukemia with breakpoint cluster region rearrangement: molecu-
lar analysis, clinical characteristics, and response to therapy. J Clin Oncol.
1988;6:1569-1575.
47. Kurzrock R, Kantarjian HM, Shtalrid M, Gutterman JU, Talpaz M.
Philadelphia chromosome-negative chronic myelogenous leukemia without
breakpoint cluster region rearrangement: a chronic myeloid leukemia with a
distinct clinical course. Blood. 1990;75:445-452.
48. Jennings BA, Mills KI. c-myc locus amplification and the acquisition of
trisomy 8 in the evolution of chronic myeloid leukaemia. Leuk Res. 1998;
22:899-903.
49. Calabretta B, Perrotti D. The biology of CML blast crisis. Blood. 2004;
103:4010-4022.
50. Gaiger A, Henn T, Horth E, et al. Increase of bcr-abl chimeric mRNA
expression in tumor cells of patients with chronic myeloid leukemia precedes
disease progression. Blood. 1995;86:2371-2378.
51. Mitelman F, Levan G, Nilsson PG, Brandt L. Non-random karyotypic
evolution in chronic myeloid leukemia. Int J Cancer. 1976;18:24-30.
52. Lawler SD, OMalley F, Lobb DS. Chromosome banding studies in
Philadelphia chromosome positive myeloid leukaemia. Scand J Haematol.
1976;17:17-28.
53. Feinstein E, Cimino G, Gale RP, et al. p53 in chronic myelogenous
leukemia in acute phase. Proc Natl Acad Sci U S A. 1991;88:6293-6297.
54. Sill H, Goldman JM, Cross NC. Homozygous deletions of the p16
tumor-suppressor gene are associated with lymphoid transformation of chronic
myeloid leukemia. Blood. 1995;85:2013-2016.
55. Towatari M, Adachi K, Kato H, Saito H. Absence of the human retino-
blastoma gene product in the megakaryoblastic crisis of chronic myelogenous
leukemia. Blood. 1991;78:2178-2181.
56. Ogawa S, Mitani K, Kurokawa M, et al. Abnormal expression of Evi-1
gene in human leukemias. Hum Cell. 1996;9:323-332.
57. Ahuja H, Bar-Eli M, Arlin Z, et al. The spectrum of molecular alter-
ations in the evolution of chronic myelocytic leukemia. J Clin Invest. 1991;
87:2042-2047.
58. Jamieson CH, Ailles LE, Dylla SJ, et al. Granulocyte-macrophage
progenitors as candidate leukemic stem cells in blast-crisis CML. N Engl J
Med. 2004;351:657-667.
59. Zion M, Ben-Yehuda D, Avraham A, et al. Progressive de novo DNA
methylation at the bcr-abl locus in the course of chronic myelogenous leuke-
mia. Proc Natl Acad Sci U S A. 1994;91:10722-10726.
60. Issa JP, Kantarjian H, Mohan A, et al. Methylation of the ABL1 pro-
moter in chronic myelogenous leukemia: lack of prognostic significance.
Blood. 1999;93:2075-2080.
61. Asimakopoulos FA, Shteper PJ, Krichevsky S, et al. ABL1 methylation
is a distinct molecular event associated with clonal evolution of chronic my-
eloid leukemia. Blood. 1999;94:2452-2460.
62. Roman-Gomez J, Castillejo JA, Jimenez A, et al. Cadherin-13, a media-
tor of calcium-dependent cell-cell adhesion, is silenced by methylation in
chronic myeloid leukemia and correlates with pretreatment risk profile and
cytogenetic response to interferon alfa. J Clin Oncol. 2003;21:1472-1479.
63. Loriaux M, Deininger M. Clonal cytogenetic abnormalities in Philadel-
phia chromosome negative cells in chronic myeloid leukemia patients treated
with imatinib. Leuk Lymphoma. 2004;45:2197-2203.
64. Fayad L, Kantarjian H, OBrien S, et al. Emergence of new clonal
abnormalities following interferon-alpha induced complete cytogenetic re-
sponse in patients with chronic myeloid leukemia: report of three cases. Leuke-
mia. 1997;11:767-771.
65. Medina J, Kantarjian H, Talpaz M, et al. Chromosomal abnormalities in
Philadelphia chromosome-negative metaphases appearing during imatinib
mesylate therapy in patients with Philadelphia chromosome-positive chronic
myelogenous leukemia in chronic phase. Cancer. 2003;98:1905-1911.
66. Bumm T, Muller C, Al-Ali HK, et al. Emergence of clonal cytogenetic
abnormalities in Ph-cells in some CML patients in cytogenetic remission to
imatinib but restoration of polyclonal hematopoiesis in the majority. Blood.
2003;101:1941-1949.
67. Abelson HT, Rabstein LS. Lymphosarcoma: virus-induced thymic-
independent disease in mice. Cancer Res. 1970;30:2213-2222.
68. Laneuville P. Abl tyrosine protein kinase. Semin Immunol. 1995;7:255-
266.
69. Bernards A, Rubin CM, Westbrook CA, Paskind M, Baltimore D. The
first intron in the human c-abl gene is at least 200 kilobases long and is a target
for translocations in chronic myelogenous leukemia. Mol Cell Biol.
1987;7:3231-3236.
70. Heisterkamp N, Jenster G, ten Hoeve J, Zovich D, Pattengale PK,
Groffen J. Acute leukaemia in bcr/abl transgenic mice. Nature. 1990;344:251-
253.
71. Melo JV. The diversity of BCR-ABL fusion proteins and their relation-
ship to leukemia phenotype. Blood. 1996;88:2375-2384.
72. Faderl S, Talpaz M, Estrov Z, OBrien S, Kurzrock R, Kantarjian HM.
The biology of chronic myeloid leukemia. N Engl J Med. 1999;341:164-172.
73. Sawyers CL. Chronic myeloid leukemia. N Engl J Med. 1999;340:
1330-1340.
74. Ravandi F, Cortes J, Albitar M, et al. Chronic myelogenous leukaemia
with p185(BCR/ABL) expression: characteristics and clinical significance. Br
J Haematol. 1999;107:581-586.
75. Pane F, Frigeri F, Sindona M, et al. Neutrophilic-chronic myeloid leu-
kemia: a distinct disease with a specific molecular marker (BCR/ABL with C3/
A2 junction). Blood. 1996;88:2410-2414.
76. Lugo TG, Pendergast AM, Muller AJ, Witte ON. Tyrosine kinase
activity and transformation potency of bcr-abl oncogene products. Science.
1990;247:1079-1082.
77. Bose S, Deininger M, Gora-Tybor J, Goldman JM, Melo JV. The
presence of typical and atypical BCR-ABL fusion genes in leukocytes of
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 986
CHRONIC MYELOID LEUKEMIA
normal individuals: biologic significance and implications for the assessment
of minimal residual disease. Blood. 1998;92:3362-3367.
78. Biernaux C, Sels A, Huez G, Stryckmans P. Very low level of major
BCR-ABL expression in blood of some healthy individuals. Bone Marrow
Transplant. 1996;17(suppl 3):S45-S47.
79. Angel P, Karin M. The role of Jun, Fos and the AP-1 complex in cell-
proliferation and transformation. Biochim Biophys Acta. 1991;1072:129-157.
80. Liebermann DA, Gregory B, Hoffman B. AP-1 (Fos/Jun) transcription
factors in hematopoietic differentiation and apoptosis. Int J Oncol. 1998;12:
685-700.
81. Passegue E, Jochum W, Schorpp-Kistner M, Mohle-Steinlein U,
Wagner EF. Chronic myeloid leukemia with increased granulocyte progenitors
in mice lacking junB expression in the myeloid lineage. Cell. 2001;104:21-32.
82. Ramaraj P, Singh H, Niu N, et al. Effect of mutational inactivation of
tyrosine kinase activity on BCR/ABL-induced abnormalities in cell growth and
adhesion in human hematopoietic progenitors. Cancer Res. 2004;64:5322-
5331.
83. Zhao RC, Jiang Y, Verfaillie CM. A model of human p210(bcr/ABL)-
mediated chronic myelogenous leukemia by transduction of primary normal
human CD34(+) cells with a BCR/ABL-containing retroviral vector. Blood.
2001;97:2406-2412.
84. Daley GQ, Van Etten RA, Baltimore D. Induction of chronic myelog-
enous leukemia in mice by the P210bcr/abl gene of the Philadelphia chromo-
some. Science. 1990;247:824-830.
85. Kelliher MA, McLaughlin J, Witte ON, Rosenberg N. Induction of a
chronic myelogenous leukemia-like syndrome in mice with v-abl and BCR/
ABL. Proc Natl Acad Sci U S A. 1990;87:6649-6653.
86. Pendergast AM, Gishizky ML, Havlik MH, Witte ON. SH1 domain
autophosphorylation of P210 BCR/ABL is required for transformation but not
growth factor independence. Mol Cell Biol. 1993;13:1728-1736.
87. Zhang X, Subrahmanyam R, Wong R, Gross AW, Ren R. The NH(2)-
terminal coiled-coil domain and tyrosine 177 play important roles in induction
of a myeloproliferative disease in mice by Bcr-Abl. Mol Cell Biol. 2001;21:
840-853.
88. Ren R. Mechanisms of BCR-ABL in the pathogenesis of chronic mye-
logenous leukaemia. Nat Rev Cancer. 2005;5:172-183.
89. Pendergast AM, Quilliam LA, Cripe LD, et al. BCR-ABL-induced
oncogenesis is mediated by direct interaction with the SH2 domain of the
GRB-2 adaptor protein. Cell. 1993;75:175-185.
90. Hu Y, Liu Y, Pelletier S, et al. Requirement of Src kinases Lyn, Hck and
Fgr for BCR-ABL1-induced B-lymphoblastic leukemia but not chronic my-
eloid leukemia. Nat Genet. 2004;36:453-461.
91. Bhatia R, Holtz M, Niu N, et al. Persistence of malignant hematopoietic
progenitors in chronic myelogenous leukemia patients in complete cytogenetic
remission following imatinib mesylate treatment. Blood. 2003;101:4701-
4707.
92. Donato NJ, Wu JY, Stapley J, et al. BCR-ABL independence and LYN
kinase overexpression in chronic myelogenous leukemia cells selected for
resistance to STI571. Blood. 2003;101:690-698.
93. Kerkhoff E, Rapp UR. Cell cycle targets of Ras/Raf signalling.
Oncogene. 1998;17:1457-1462.
94. Nosaka T, Kawashima T, Misawa K, Ikuta K, Mui AL, Kitamura T.
STAT5 as a molecular regulator of proliferation, differentiation and apoptosis
in hematopoietic cells. Embo J. 1999;18:4754-4765.
95. Sonoyama J, Matsumura I, Ezoe S, et al. Functional cooperation among
Ras, STAT5, and phosphatidylinositol 3-kinase is required for full oncogenic
activities of BCR/ABL in K562 cells. J Biol Chem. 2002;277:8076-8082.
96. Neviani P, Santhanam R, Trotta R, et al. The tumor suppressor PP2A is
functionally inactivated in blast crisis CML through the inhibitory activity of
the BCR/ABL-regulated SET protein. Cancer Cell. 2005;8:355-368.
97. Janssens V, Goris J. Protein phosphatase 2A: a highly regulated family
of serine/threonine phosphatases implicated in cell growth and signalling.
Biochem J. 2001;353(pt 3):417-439.
98. Talpaz M, Mavligit G, Keating M, Walters RS, Gutterman JU. Human
leukocyte interferon to control thrombocytosis in chronic myelogenous leuke-
mia. Ann Intern Med. 1983;99:789-792.
99. Talpaz M, Kantarjian HM, McCredie K, Trujillo JM, Keating MJ,
Gutterman JU. Hematologic remission and cytogenetic improvement induced
by recombinant human interferon alpha A in chronic myelogenous leukemia. N
Engl J Med. 1986;314:1065-1069.
100. Kantarjian HM, Smith TL, OBrien S, Beran M, Pierce S, Talpaz M,
The Leukemia Service. Prolonged survival in chronic myelogenous leukemia
after cytogenetic response to interferon-alpha therapy. Ann Intern Med. 1995;
122:254-261.
101. OBrien S, Kantarjian H, Koller C, et al. Sequential homoharringtonine
and interferon-alpha in the treatment of early chronic phase chronic myelog-
enous leukemia. Blood. 1999;93:4149-4153.
102. OBrien S, Talpaz M, Cortes J, et al. Simultaneous homoharringtonine
and interferon-alpha in the treatment of patients with chronic-phase chronic
myelogenous leukemia. Cancer. 2002;94:2024-2032.
103. Baccarani M, Rosti G, de Vivo A, et al. A randomized study of inter-
feron-alpha versus interferon-alpha and low-dose arabinosyl cytosine in
chronic myeloid leukemia. Blood. 2002;99:1527-1535.
104. Kantarjian HM, OBrien S, Smith TL, et al. Treatment of Philadelphia
chromosome-positive early chronic phase chronic myelogenous leukemia with
daily doses of interferon alpha and low-dose cytarabine. J Clin Oncol. 1999;17:
284-292.
105. Guilhot F, Chastang C, Michallet M, et al, French Chronic Myeloid
Leukemia Study Group. Interferon alfa-2b combined with cytarabine versus
interferon alone in chronic myelogenous leukemia. N Engl J Med. 1997;337:
223-229.
106. Kantarjian HM, OBrien S, Cortes JE, et al. Complete cytogenetic and
molecular responses to interferon-alpha-based therapy for chronic myelog-
enous leukemia are associated with excellent long-term prognosis. Cancer.
2003;97:1033-1041.
107. Hochhaus A, Reiter A, Saussele S, et al, German CML Study Group and
the UK MRC CML Study Group. Molecular heterogeneity in complete cytoge-
netic responders after interferon-alpha therapy for chronic myelogenous leuke-
mia: low levels of minimal residual disease are associated with continuing
remission. Blood. 2000;95:62-66.
108. Molldrem JJ, Lee PP, Wang C, et al. Evidence that specific T lympho-
cytes may participate in the elimination of chronic myelogenous leukemia. Nat
Med. 2000;6:1018-1023.
109. Morton AJ, Gooley T, Hansen JA, et al. Association between
pretransplant interferon-alpha and outcome after unrelated donor marrow
transplantation for chronic myelogenous leukemia in chronic phase. Blood.
1998;92:394-401.
110. Giralt SA, Kantarjian HM, Talpaz M, et al. Effect of prior interferon
alfa therapy on the outcome of allogeneic bone marrow transplantation for
chronic myelogenous leukemia. J Clin Oncol. 1993;11:1055-1061.
111. Giralt S, Szydlo R, Goldman JM, et al. Effect of short-term interferon
therapy on the outcome of subsequent HLA-identical sibling bone marrow
transplantation for chronic myelogenous leukemia: an analysis from the inter-
national bone marrow transplant registry. Blood. 2000;95:410-415.
112. Hehlmann R, Hochhaus A, Kolb HJ, et al. Interferon-alpha before
allogeneic bone marrow transplantation in chronic myelogenous leukemia
does not affect outcome adversely, provided it is discontinued at least 90 days
before the procedure. Blood. 1999;94:3668-3677.
113. Lee SJ, Klein JP, Anasetti C, et al. The effect of pretransplant interferon
therapy on the outcome of unrelated donor hematopoietic stem cell transplanta-
tion for patients with chronic myelogenous leukemia in first chronic phase.
Blood. 2001;98:3205-3211.
114. Holowiecki J, Kruzel T, Wojnar J, et al. Allogeneic stem cell transplan-
tation from unrelated and related donors is effective in CML Ph+ and ALL Ph+
patients pretreated with imatinib [abstract]. Blood. 2003;102(pt 2):436b.
115. Prejzner W, Zaucha JM, Szatkowski D, et al. Imatinib therapy prior to
myeloablative allogeneic stem cell transplantation does not increase acute
transplant related toxicity [abstract]. Blood. 2003;102(pt 1):911a.
116. Hansen JA, Gooley TA, Martin PJ, et al. Bone marrow transplants from
unrelated donors for patients with chronic myeloid leukemia. N Engl J Med.
1998;338:962-968.
117. Baker KS, Gurney JG, Ness KK, et al. Late effects in survivors of
chronic myeloid leukemia treated with hematopoietic cell transplantation:
results from the Bone Marrow Transplant Survivor Study. Blood. 2004;104:
1898-1906.
118. Kiss TL, Abdolell M, Jamal N, Minden MD, Lipton JH, Messner HA.
Long-term medical outcomes and quality-of-life assessment of patients with
chronic myeloid leukemia followed at least 10 years after allogeneic bone
marrow transplantation. J Clin Oncol. 2002;20:2334-2343.
119. Radich JP, Gooley T, Bensinger W, et al. HLA-matched related he-
matopoietic cell transplantation for chronic-phase CML using a targeted
busulfan and cyclophosphamide preparative regimen. Blood. 2003;102:31-
35.
120. Weisdorf DJ, Anasetti C, Antin JH, et al. Allogeneic bone marrow
transplantation for chronic myelogenous leukemia: comparative analysis of
unrelated versus matched sibling donor transplantation. Blood. 2002;99:1971-
1977.
121. Or R, Shapira MY, Resnick I, et al. Nonmyeloablative allogeneic stem
cell transplantation for the treatment of chronic myeloid leukemia in first
chronic phase. Blood. 2003;101:441-445.
122. Crawley C, Szydlo R, Lalancette M, et al. Outcomes of reduced-inten-
sity transplantation for chronic myeloid leukemia: an analysis of prognostic
factors from the Chronic Leukemia Working Party of the EBMT. Blood.
2005;106:2969-2976. Epub 2005 Jul 5.
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 987
CHRONIC MYELOID LEUKEMIA
123. Radich JP, Gehly G, Gooley T, et al. Polymerase chain reaction detec-
tion of the BCR-ABL fusion transcript after allogeneic marrow transplantation
for chronic myeloid leukemia: results and implications in 346 patients. Blood.
1995;85:2632-2638.
124. Olavarria E, Craddock C, Dazzi F, et al. Imatinib mesylate (STI571) in
the treatment of relapse of chronic myeloid leukemia after allogeneic stem cell
transplantation. Blood. 2002;99:3861-3862.
125. Druker BJ, Tamura S, Buchdunger E, et al. Effects of a selective
inhibitor of the Abl tyrosine kinase on the growth of Bcr-Abl positive cells. Nat
Med. 1996;2:561-566.
126. Beran M, Cao X, Estrov Z, et al. Selective inhibition of cell proliferation
and BCR-ABL phosphorylation in acute lymphoblastic leukemia cells express-
ing Mr 190,000 BCR-ABL protein by a tyrosine kinase inhibitor (CGP-57148).
Clin Cancer Res. 1998;4:1661-1672.
127. Druker BJ, Talpaz M, Resta DJ, et al. Efficacy and safety of a specific
inhibitor of the BCR-ABL tyrosine kinase in chronic myeloid leukemia. N Engl
J Med. 2001;344:1031-1037.
128. Kantarjian H, Sawyers C, Hochhaus A, et al. Hematologic and cytoge-
netic responses to imatinib mesylate in chronic myelogenous leukemia. N Engl
J Med. 2002;346:645-652.
129. Kantarjian HM, Cortes JE, OBrien S, et al. Long-term survival benefit
and improved complete cytogenetic and molecular response rates with imatinib
mesylate in Philadelphia chromosome-positive chronic-phase chronic myeloid
leukemia after failure of interferon-alpha. Blood. 2004;104:1979-1988.
130. OBrien SG, Guilhot F, Larson RA, et al. Imatinib compared with
interferon and low-dose cytarabine for newly diagnosed chronic-phase chronic
myeloid leukemia. N Engl J Med. 2003;348:994-1004.
131. Hahn EA, Glendenning GA, Sorensen MV, et al. Quality of life in
patients with newly diagnosed chronic phase chronic myeloid leukemia on
imatinib versus interferon alfa plus low-dose cytarabine: results from the IRIS
Study. J Clin Oncol. 2003;21:2138-2146.
132. Simonsson B, the IRIS (International Randomized IFN vs ST1571)
Study Group. Beneficial effects of cytogenetic and molecular response on
long-term outcome in patients with newly diagnosed chronic myeloid leukemia
in chronic phase (CML-CP) treated with imatinib (IM): update from the IRIS
study [abstract]. Blood. 2005;106:52a.
133. Colombat M, Fort MP, Chollet C, et al. Molecular remission in chronic
myeloid leukemia patients with sustained complete cytogenetic remission after
imatinib mesylate treatment. Haematologica. 2006;91:162-168.
134. Goldman JM, Hughes T, Radich J, et al. Continuing reduction in level
of residual disease after 4 years in patients with CML in chronic phase
responding to first-line imatinib (IM) in the IRIS study [abstract]. Blood. 2005;
106:51a.
135. Kantarjian H, OBrien S, Cortes J, et al. Analysis of the impact of
imatinib mesylate therapy on the prognosis of patients with Philadelphia
chromosome-positive chronic myelogenous leukemia treated with interferon-
alpha regimens for early chronic phase. Cancer. 2003;98:1430-1437.
136. Roy L, Guilhot J, Krahnke T, et al. Survival advantage from imatinib
compared to the combination interferon-{alpha} plus cytarabine in chronic
phase CML: historical comparison between two phase III trials. Blood. Epub
2006 Apr 20.
137. Kantarjian H, Talpaz M, OBrien S, et al. High-dose imatinib mesylate
therapy in newly diagnosed Philadelphia chromosome-positive chronic phase
chronic myeloid leukemia. Blood. 2004;103:2873-2878.
138. Hughes TP, Branford S, Matthews J, et al. Trial of higher dose imatinib
with selective intensification in newly diagnosed CML patients in the chronic
phase [abstract]. Blood. 2003;102(Pt 1):31a.
139. Cortes J, Giles F, Salvado A, et al. High-dose (HD) imatinib in patients
(pts) with previously untreated chronic myeloid leukemia (CML) in early
chronic phase (CP): preliminary results of a multicenter community based trial
[abstract]. J Clin Oncol. 2005;23(suppl):564s.
140. Rosti G, Martinelli G, Castagnetti F, et al. Imatinib 800 mg: prelimi-
nary results of a phase II trial of the GIMEMA CML working party in
intermediate sokal risk patients and status-of-the-art of an ongoing multina-
tional, prospective randomized trial of imatinib standard dose (400 mg daily)
vs high dose (800 mg daily) in high sokal risk patients [abstract]. Blood.
2005;106:320a.
141. Cortes J, Giles F, OBrien S, et al. Result of high-dose imatinib
mesylate in patients with Philadelphia chromosome-positive chronic myeloid
leukemia after failure of interferon-alpha. Blood. 2003;102:83-86.
142. Cortes J, Talpaz M, OBrien S, et al. Molecular responses in patients
with chronic myelogenous leukemia in chronic phase treated with imatinib
mesylate. Clin Cancer Res. 2005;11:3425-3432.
143. Hughes TP, Branford S, Reynolds J, et al. Maintenance of imatinib dose
intensity in the first six months of therapy for newly diagnosed patients with
CML is predictive of molecular response, independent of the ability to increase
dose at a later point [abstract]. Blood. 2005;106:51a.
144. Cortes J, OBrien S, Kantarjian H. Discontinuation of imatinib therapy
after achieving a molecular response. Blood. 2004;104:2204-2205.
145. Mauro MJ, Druker BJ, Maziarz RT. Divergent clinical outcome in two
CML patients who discontinued imatinib therapy after achieving a molecular
remission. Leuk Res. 2004;28(suppl 1):S71-S73.
146. Ghanima W, Kahrs J, Dahl TG, 3rd, Tjonnfjord GE. Sustained cytoge-
netic response after discontinuation of imatinib mesylate in a patient with
chronic myeloid leukaemia. Eur J Haematol. 2004;72:441-443.
147. Rousselot P, Huguet F, Cayuela JM, et al. Imatinib mesylate discontinu-
ation in patients with chronic myelogenous leukaemia in complete molecular
remission for more than two years [abstract]. Blood. 2005;106:321a.
148. Graham SM, Jorgensen HG, Allan E, et al. Primitive, quiescent, Phila-
delphia-positive stem cells from patients with chronic myeloid leukemia are
insensitive to STI571 in vitro. Blood. 2002;99:319-325.
149. Copland M. Hamilton A, Baird JW, et al. Dasatinib (BMS-354825) has
increased activity against Bcr-Abl compared to imatinib in primary CML cells
in vitro, but does not eradicate quiescent CML stem cells [abstract]. Blood.
2005;106:205a.
150. Chu S, Xu H, Shah NP, et al. Detection of BCR-ABL kinase mutations
in CD34+ cells from chronic myelogenous leukemia patients in complete
cytogenetic remission on imatinib mesylate treatment. Blood. 2005;105:
2093-2098.
151. Jiang X, Zhao Y, Chan P, et al. Quantitative real-time RT-PCR analysis
shows BCR-ABL expression progressively decreases as primitive leukemic
cells from patients with chronic myeloid leukemia (CML) differentiate in vivo
[abstract]. Exp Hematol. 2003;31:228-229.
152. Hughes T, Kaeda J, Branford S, et al. Molecular responses to imatinib
(STI571) or interferon + Ara-C as initial therapy for CML: results in the IRIS
study [abstract]. Blood. 2002;100:93a.
153. Cortes J, OBrien S, Talpaz M, et al. Clinical significance of molecular
response in chronic myeloid leukemia (CML) after imatinib mesylate
(Gleevec) therapy: low levels of residual disease predict for response duration
[abstract]. Blood. 2003;102(Pt 1):416a.
154. Talpaz M, Kantarjian H, Kurzrock R, Trujillo JM, Gutterman JU.
Interferon-alpha produces sustained cytogenetic responses in chronic myelog-
enous leukemia. Philadelphia chromosome-positive patients. Ann Intern Med.
1991;114:532-538.
155. Hughes TP, Deininger MW, Hochhaus A, et al. Monitoring CML
patients responding to treatment with tyrosine kinase inhibitors: review and
recommendations for harmonizing current methodology for detecting BCR-
ABL transcripts and kinase domain mutations and for expressing results.
Blood. Epub 2006 Mar 7.
156. Hochhaus A, La Rosee P. Imatinib therapy in chronic myelogenous
leukemia: strategies to avoid and overcome resistance. Leukemia. 2004;18:
1321-1331.
157. OHare T, Walters DK, Stoffregen EP, et al. In vitro activity of Bcr-Abl
inhibitors AMN107 and BMS-354825 against clinically relevant imatinib-
resistant Abl kinase domain mutants. Cancer Res. 2005;65:4500-4505.
158. Deininger M, Buchdunger E, Druker BJ. The development of imatinib
as a therapeutic agent for chronic myeloid leukemia. Blood. 2005;105:2640-
2653.
159. Carter TA, Wodicka LM, Shah NP, et al. Inhibition of drug-resistant
mutants of ABL, KIT, and EGF receptor kinases. Proc Natl Acad Sci U S A.
2005;102:11011-11016.
160. Cowan-Jacob SW, Guez V, Fendrich G, et al. Imatinib (STI571) resis-
tance in chronic myelogenous leukemia: molecular basis of the underlying
mechanisms and potential strategies for treatment. Mini Rev Med Chem.
2004;4:285-299.
161. Branford S, Rudzki Z, Walsh S, et al. Detection of BCR-ABL mutations
in patients with CML treated with imatinib is virtually always accompanied by
clinical resistance, and mutations in the ATP phosphate-binding loop (P-loop)
are associated with a poor prognosis. Blood. 2003;102:276-283.
162. Hochhaus A, Kreil S, Corbin A, et al. Roots of clinical resistance to STI-
571 cancer therapy. Science. 2001;293:2163.
163. Soverini S, Martinelli G, Rosti G, et al. ABL mutations in late chronic
phase chronic myeloid leukemia patients with up-front cytogenetic resistance
to imatinib are associated with a greater likelihood of progression to blast crisis
and shorter survival: a study by the GIMEMA Working Party on Chronic
Myeloid Leukemia. J Clin Oncol. 2005;23:4100-4109.
164. Jabbour E, Kantarjian H, Jones D, et al. Long-term incidence and
outcome of BCR-ABL mutations in patients (pts) with chronic myeloid leuke-
mia (CML) treated with imatinib mesylate: P-loop mutations are not associated
with worse outcome [abstract]. Blood. 2004;104:288a.
165. Jabbour E, Kantarjian H, Talpaz M, et al. Outcome of salvage therapy
in patients (pts) with chronic myeloid leukemia (CML) who failed imatinib
after developing BCR-ABL kinase mutation [abstract]. Blood. 2005;106:
318a.
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.
Mayo Clin Proc. July 2006;81(7):973-988 www.mayoclinicproceedings.com 988
CHRONIC MYELOID LEUKEMIA
166. Pricl S, Fermeglia M, Ferrone M, Tamborini E. T315I-mutated Bcr-Abl
in chronic myeloid leukemia and imatinib: insights from a computational
study. Mol Cancer Ther. 2005;4:1167-1174.
167. Tanis KQ, Veach D, Duewel HS, Bornmann WG, Koleske AJ. Two
distinct phosphorylation pathways have additive effects on Abl family kinase
activation. Mol Cell Biol. 2003;23:3884-3896.
168. Leguay T, Desplat V, Marit G, Mahon FX. D276G mutation is associ-
ated with a poor prognosis in imatinib mesylate-resistant chronic myeloid
leukemia patients [letter]. Leukemia. 2005;19:2332-2333; author reply 2333-
2334.
169. Donato NJ, Wu JY, Stapley J, et al. Imatinib mesylate resistance
through BCR-ABL independence in chronic myelogenous leukemia. Cancer
Res. 2004;64:672-677.
170. Kantarjian H, Ottmann O, Cortes J, et al. AMN107, a novel amino-
pyrimidine inhibitor of Bcr-Abl, has significant activity in imatinib-resistant
bcr-abl positive chronic myeloid leukemia (CML) [abstract]. J Clin Oncol.
2005;23(suppl):195s.
171. Sawyers CL, Shah NP, Kantarjian HM, et al. A phase I study of BMS-
354825 in patients with imatinib-resistant and intolerant accelerated and blast
phase chronic myeloid leukemia (CML): results from CA180002 [abstract]. J
Clin Oncol. 2005;23(suppl):565s.
172. Talpaz M, Kantarjian H, Paquette R, et al. A phase I study of BMS-
354825 in patients with imatinib-resistant and intolerant chronic phase chronic
myeloid leukemia (CML): results from CA180002 [abstract]. J Clin Oncol.
2005;23(suppl):564s.
173. Hochhaus A, Baccarani M, Sawhers C, et al. Efficacy of dasatinib in
patients with chronic phase Philadelphia chromosome-positive CML resistant
or intolerant to imatinib: first results of the CA180013 START-C phase II
study [abstract]. Blood. 2005;106:17a.
174. Guilhot F, Apperley JF, Shah N, et al. A phase II study of dasatinib in
patients with accelerated phase chronic myeloid leukemia (CML) who are
resistant or intolerant to imatinib: first results of the CA180005 START-A
study [abstract]. Blood. 2005;106:16a.
175. Talpaz M, Rousselot P, Kim DW, et al. A phase II study of dasatinib in
patients with chronic myeloid leukemia (CML) in myeloid blast crisis who are
resistant or intolerant to imatinib: first results of the CA180006 START-B
study [abstract]. Blood. 2005;106:16a.
176. Ottmann OG, Martinelli G, Dombret H, et al. A phase II study of
dasatinib in patients with chronic myeloid leukemia (CML) in lymphoid blast
crisis or Philadelphia-chromosome positive acute lymphoblastic leukemia
(Ph+ ALL) who are resistant or intolerant to imatinib: the START-L
CA180015 study [abstract]. Blood. 2005;106:17a.
177. Lombardo LJ, Lee FY, Chen P, et al. Discovery of N-(2-chloro-6-
methyl-phenyl)-2-(6-(4-(2-hydroxyethyl)-piperazin-1-yl)-2-methylpyrimidin-
4- ylamino)thiazole-5-carboxamide (BMS-354825), a dual Src/Abl kinase in-
hibitor with potent antitumor activity in preclinical assays. J Med Chem.
2004;47:6658-6661.
178. Shah NP, Nicoll JM, Branford S, et al. Molecular analysis of dasatinib
resistance mechanisms in CML patients identified novel BCR-ABL mutations
predicted to retain sensitivity to imatinib: rationale for combination tyrosine
kinase inhibitor therapy [abstract]. Blood. 2005;106:318a.
179. Golas JM, Arndt K, Etienne C, et al. SKI-606, a 4-anilino-3-
quinolinecarbonitrile dual inhibitor of Src and Abl kinases, is a potent
antiproliferative agent against chronic myelogenous leukemia cells in culture
and causes regression of K562 xenografts in nude mice. Cancer Res. 2003;
63:375-381.
180. Kimura S, Naito H, Segawa H, et al. NS-187, a potent and selective dual
Bcr-Abl/Lyn tyrosine kinase inhibitor, is a novel agent for imatinib-resistant
leukemia. Blood. 2005;106:3948-3954.
181. Hennequin LF, Allen J, Costello GF, et al. The discovery of AZD0530:
a novel, oral, highly selective and dual-specific inhibitor of the Src and Abl
family kinases [abstract]. Proc Amer Assoc Cancer Res. 2005;46:2537.
182. Wolff NC, Veach DR, Tong WP, Bornmann WG, Clarkson B, Ilaria RL
Jr. PD166326, a novel tyrosine kinase inhibitor, has greater antileukemic
activity than imatinib mesylate in a murine model of chronic myeloid leuke-
mia. Blood. 2005;105:3995-4003.
183. Dorsey JF, Jove R, Kraker AJ, Wu J. The pyrido[2,3-d]pyrimidine
derivative PD180970 inhibits p210Bcr-Abl tyrosine kinase and induces
apoptosis of K562 leukemic cells. Cancer Res. 2000;60:3127-3131.
184. Gumireddy K, Baker SJ, Cosenza SC, et al. A non-ATP-competitive
inhibitor of BCR-ABL overrides imatinib resistance. Proc Natl Acad Sci U S A.
2005;102:1992-1997.
185. Cortes J, Albitar M, Thomas D, et al. Efficacy of the farnesyl transferase
inhibitor R115777 in chronic myeloid leukemia and other hematologic malig-
nancies. Blood. 2003;101:1692-1697.
186. Peters DG, Hoover RR, Gerlach MJ, et al. Activity of the farnesyl
protein transferase inhibitor SCH66336 against BCR/ABL-induced murine
leukemia and primary cells from patients with chronic myeloid leukemia.
Blood. 2001;97:1404-1412.
187. Hoover RR, Mahon FX, Melo JV, Daley GQ. Overcoming STI571
resistance with the farnesyl transferase inhibitor SCH66336. Blood. 2002;
100:1068-1071.
188. Reichert A, Heisterkamp N, Daley GQ, Groffen J. Treatment of Bcr/
Abl-positive acute lymphoblastic leukemia in P190 transgenic mice with the
farnesyl transferase inhibitor SCH66336. Blood. 2001;97:1399-1403.
189. Marin D, Kaeda JS, Andreasson C, et al. Phase I/II trial of adding
semisynthetic homoharringtonine in chronic myeloid leukemia patients who
have achieved partial or complete cytogenetic response on imatinib. Cancer.
2005;103:1850-1855.
190. Issa JP, Gharibyan V, Cortes J, et al. Phase II study of low-dose
decitabine in patients with chronic myelogenous leukemia resistant to imatinib
mesylate. J Clin Oncol. 2005;23:3948-3956.
191. Kantarjian HM, OBrien S, Cortes J, et al. Results of decitabine (5-aza-
2deoxycytidine) therapy in 130 patients with chronic myelogenous leukemia.
Cancer. 2003;98:522-528.
192. Yu C, Rahmani M, Almenara J, et al. Histone deacetylase inhibitors
promote STI571-mediated apoptosis in STI571-sensitive and -resistant Bcr/
Abl+ human myeloid leukemia cells. Cancer Res. 2003;63:2118-2126.
193. Nimmanapalli R, Fuino L, Stobaugh C, Richon V, Bhalla K.
Cotreatment with the histone deacetylase inhibitor suberoylanilide hydroxamic
acid (SAHA) enhances imatinib-induced apoptosis of Bcr-Abl-positive human
acute leukemia cells. Blood. 2003;101:3236-3239.
194. Nimmanapalli R, Fuino L, Bali P, et al. Histone deacetylase inhibitor
LAQ824 both lowers expression and promotes proteasomal degradation of Bcr-
Abl and induces apoptosis of imatinib mesylate-sensitive or -refractory chronic
myelogenous leukemia-blast crisis cells. Cancer Res. 2003;63:5126-5135.
195. Guilhot F, Lacotte-Thierry L. Interferon-alpha: mechanisms of action in
chronic myelogenous leukemia in chronic phase. Hematol Cell Ther. 1998;40:
237-239.
196. Barrett J. Allogeneic stem cell transplantation for chronic myeloid
leukemia. Semin Hematol. 2003;40:59-71.
197. Pinilla-Ibarz J, Cathcart K, Korontsvit T, et al. Vaccination of patients
with chronic myelogenous leukemia with bcr-abl oncogene breakpoint fusion
peptides generates specific immune responses. Blood. 2000;95:1781-1787.
198. Cathcart K, Pinilla-Ibarz J, Korontsvit T, et al. A multivalent bcr-abl
fusion peptide vaccination trial in patients with chronic myeloid leukemia.
Blood. 2004;103:1037-1042.
199. Bocchia M, Gentili S, Abruzzese E, et al. Effect of a p210 multipeptide
vaccine associated with imatinib or interferon in patients with chronic myeloid
leukaemia and persistent residual disease: a multicentre observational trial.
Lancet. 2005;365:657-662.
200. Qazilbash MH, Wieder E, Rios R, et al. Vaccination with the PR1
leukemia-associated antigen can induce complete remission in patients with
myeloid leukemia [abstract]. Blood. 2004;104:77a.
201. Li Z, Qiao Y, Liu B, et al. Combination of imatinib mesylate with
autologous leukocyte-derived heat shock protein and chronic myelogenous
leukemia. Clin Cancer Res. 2005;11:4460-4468.
202. Borrello I, Levitsky H, Damon L, et al. Vaccine-associated immune and
WT-1 responses are associated with better relapse-free survival in patients with
AML in remission treated with a GM-CSF secreting leukemia vaccine and
autologous stem cell transplant (ASCT) [abstract]. J Clin Oncol. 2005;23
(suppl):569s.
The Symposium on Oncology Practice: Hematological Malignancies
will continue in the August issue.
For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings. For personal use. Mass reproduce only with permission fromMayo Clinic Proceedings.

Vous aimerez peut-être aussi