Genetic Aberrations and Response to Fludarabine as First Line Treatment in a Serie of B-CLL Patients.

Blood ◽  
2005 ◽  
Vol 106 (11) ◽  
pp. 2131-2131
Author(s):  
Pilar Giraldo ◽  
Araceli Rubio-Martinez ◽  
Valle Recasens ◽  
Mikel Valganon ◽  
Paz Latre ◽  
...  

Abstract Background: Genetic factors have been established in the recent years as the most important independent predictors of disease progression and survival in patients with B-chronic lymphocytic leukaemia (B-CLL). Fludarabine is an approved first-line treatment for B-CLL, which achieves superior remission rates than traditional therapies. The drug is equally effective in early and advanced disease and in younger and elderly patients with B-CLL but different results have been found in patients with cytogenetic aberrations as methylation in the TP53 promoter Aims: To evaluate the response to Fludarabine as first-line therapy in B-CLL patients with advanced stages. Patients and methods: In the study 55 B-CLL patients were enrolled. The diagnosis was established based on standard morphologic and immunophenotypic criteria, and classified on high clinical risk category Rai stage III/IV treated with Fludarabine (25 mg/m2 daily for 5 days IV in a 28-day cycle) as first-line therapy from January 00 to December 03. Response was evaluated after 4 cycles of therapy and in those patients which had not achieved complete remission (CR) two additional cycles were administrated. Criteria for response were established using the revised 1996 NCIWG. FISH studies were performed using the probes LSI p53 (17p13), LSI D13S25 (13q14), CEP 12 and ATM gene (11q22). The sequences of the IGVH genes were determined in cDNA of the patients. Responses to therapy, time to progression disease and overall survival were available to analyse in 41 patients. Association was studied using contingence tables followed by Fisher exact test and to evaluate survival by acturial method of Kaplan Meier and Cox regression. Results: Mean age 64 (36.9–83.9), female 53.7%. 75% of patients presented genetic aberrations associated with worse prognosis: del (17p) and/or del (11q) (45%). No-mutated IGVH genes (57%). Response: 83.5% of patients achieved response (CR 47.8%). Related to genetic aberrations the responses were: 55.5% in patients with del(17p), 75.0% del(11q), 100% +12, 88.8% del(13q) and 100% of patients without genetic alterations showed response to Fludarabine. The 48.8 % of patients developed progression of disease with mean duration of the response of 15.1 months. Related to genetic aberrations the progression was over 81% of patients with genetic aberrations versus 55.5% without genetic alterations (p=0.044). The overall survival was 68.0 months. 34.1% of patients are alive in CR, 36.6% are alive with stable disease 7.3 % are alive with disease in progression and 22.0 % have died. Conclusions: A high response to Fludarabine in first line therapy was observed, however only 55.5% of patients with del(17p) showed response. Patients with genetic aberrations had higher significant rate of progression in comparison with those containing normal genotypes.

2020 ◽  
Vol 22 (3) ◽  
pp. 142-148
Author(s):  
L. V. Bolotina

Throughout the last 10 years, liver cancer mortality rate in the Russian Federation consistently exceeded the morbidity rate, which is related to the complexity of early diagnostics, absence of effective screening and oncological alertness of allied-profession doctors. In the situation when late disease intelligence does not frequently allow radical treatment, palliative methods remain the only option of survivability enhancement and improving the patients quality of life. Lenvatinib was approved as the first-line drug in the treatment of unresectable hepatocellular carcinoma based on the data of the REFLECT trial, in which the drug demonstrated achieving the patients overall survival (OS) comparable to the activity of sorafenib (13.6 months for lenvatinib vs 12.3 months for sorafenib; hazard ratio HR 0.92; 95% confidence interval CI 0.791.06). At the same time, significant inferiority of lenvatinib was observed for secondary endpoints: progression-free survival PFS (7.4 months for lenvatinib vs 3.7 months for sorafenib; HR 0.66; 95% CI 0.570.77;р0.0001), time to progression (8.9 months for lenvatinib vs 3.7 months for sorafenib; HR 0.63; 95% CI 0.530.73;р0.0001) and objective response rate ORR (24.1% for lenvatinib vs 9.2% for sorafenib). The further analysis of the results of the REFLECT study revealed the additional factors impacting patients survival, such as the level of a-fetoprotein (AFP) before treatment, treatment ORR, performance of subsequent antitumor therapy and procedures after completion of the target first-line therapy. In patients responding to lenvatinib in the first line and further receiving any second-line therapy, the mOS was 25.7 months as compared with the median overall survival (mOS) of 22.3 months in patients responding to sorafenib and receiving further second-line therapy. Additionally, in responders switching from lenvatinib to sorafenib, the mOS was 26.2 months. In the recently published comparative study of lenvatinib and transarterial chemoembolization on the BCLC B stage, inferiority of lenvatinib was demonstrated in terms of OS, PFS and ORR in certain patient categories. Considering the data obtained in the REFLECT population, where in patients achieving the RR to the first-line treatment with lenvatinib and further receiving the local antitumor procedures the mOS increased to 27.2 months (95% CI 20.729.8), prescribing target and locoregional therapy in certain cases in this very sequence is possible. The recently published data about administration of lenvatinib outside of the inclusion criteria for the REFLECT trial, have proved the efficacy and safety of this drug administration in real clinical practice, thus significantly expanding our understanding of the key role of lenvatinib in the first-line treatment of unresectable hepatocellular carcinoma.


2017 ◽  
Vol 35 (4_suppl) ◽  
pp. 468-468
Author(s):  
Hui-Li Wong ◽  
Ying Wang ◽  
Yaling Yin ◽  
Hagen F. Kennecke ◽  
Winson Y. Cheung ◽  
...  

468 Background: Chemotherapy options currently available for the first-line treatment of advanced PDAC include FOLFIRINOX (FX), gemcitabine with nab-paclitaxel (GP) and single agent gemcitabine (Gem). GP was introduced most recently and funded for clinical use in British Columbia (BC) in September 2014. In this retrospective analysis, we explore the impact of GP availability on first-line treatment selection and overall survival (OS) in advanced PDAC. Methods: The BC Cancer Agency provincial pharmacy database was used to identify patients (pts) who started FX, GP or Gem between January and August 2014 (pre-GP) or January and August 2015 (post-GP). Pts were eligible for inclusion if they received at least one cycle of first-line therapy for locally advanced or metastatic PDAC. Clinical data were extracted from electronic medical records. OS was defined as time from diagnosis of advanced PDAC to death and compared by treatment era, adjusting for age, ECOG, comorbidities, disease extent and baseline CA19-9. Results: 286 pts fulfilled eligibility criteria: 88 (31%) with locally advanced and 198 (69%) with metastatic disease. 131 and 155 pts were treated in the pre- and post-GP eras respectively. Prior to GP approval, 44% and 49% of pts received Gem and FX; this decreased to 21% and 33% after GP funding, with 46% of pts receiving GP in the latter period. Nine (7%) pts received GP in the pre-GP era, either through self-pay or addition of nab-paclitaxel after approval. There were no significant differences in pt characteristics across both eras. 46% of pts who received GP post approval had ECOG ≥ 2. The proportion of pts receiving second-line therapy was lower in the post-GP era (22% vs. 38%). Median OS in the post-GP era was 8.1 vs. 10.1 months in the pre-GP era; adjusted HR 1.28 (95% CI 0.96–1.71). Pts with ECOG ≥ 2 who received GP had a median OS of 6.5 months. Conclusions: After GP was funded, it became the preferred first-line regimen for advanced PDAC. Its more frequent use instead of FX did not appear to compromise overall survival even though a substantial proportion of pts were ECOG ≥ 2 and few pts received second-line therapy.


2019 ◽  
Vol 37 (15_suppl) ◽  
pp. e12590-e12590
Author(s):  
Hongnan Mo ◽  
Binghe Xu ◽  
Fei Ma ◽  
Qing Li ◽  
Pin Zhang ◽  
...  

e12590 Background: Use of progression-free survival (PFS) as a clinical trial endpoint in first-line treatment of patients with advanced breast cancer is attractive, but would be enhanced by establishing a correlation between PFS and overall survival (OS). Methods: From January 2003 to December 2012, 1851 patients with advanced breast cancer at start of first-line therapy were enrolled in this real-world study. An independent cohort of patients hospitalized in 2013 was used for external validation. All data were collected from the Database of China National Cancer Cancer. Results: The correlation coefficient (Pearson’s r) between PFS and OS was 0.807 in patients only receiving endocrine therapy as first-line treatment, 0.643 in those treated with chemotherapy, and 0.642 in the whole cohort. Receiver operating characteristic curve indicated that PFS = 12 months was the optimal cutoff value for predicting patient’s survival. The median OS was 30.0 months (95% CI 27.8-32.2) in the PFS < 12 months group, and 69.0 months (95% CI 60.8-77.2) in the other group (P < 0.0001). Multivariate analysis revealed that compared with patients who did not progress at 12 months, the adjusted hazard ratio (HR) for death was 2.681 (95% CI, 2.301-3.124; P < 0.0001) for patients with PFS < 12 months. Subgroup analysis based on patient’s age, molecular subtype, visceral metastasis and types of first-line treatment further confirmed that PFS < 12 months was associated with significant poor prognosis in all these subgroups. In patients with luminal type of breast cancer, the HR for death was 2.567 (95%CI 2.147-3.069; P < 0.0001) for patients with PFS < 12 months. Notably, for these patients with luminal type breast cancer who had progressed within 12 months after first-line treatment, addition of chemotherapy in the second-line therapy would surprisingly have adverse effects on patients’ survival when compared with endocrine therapy alone (HR = 1.627, 95%CI 1.016-2.604, P = 0.043). The findings were externally validated in the independent cohort. Conclusions: To our knowledge, this is the first real-world study revealed that PFS at 12 months in first-line therapy predict OS of patients with advanced breast cancer.


2019 ◽  
Vol 37 (15_suppl) ◽  
pp. e12565-e12565
Author(s):  
Nan Wang ◽  
Kun Li ◽  
Wei-Yao Kong ◽  
Xiao-Ran Liu ◽  
Meng-Yao Tan ◽  
...  

e12565 Background: Platinum-based therapy remains an effective treatment for triple-negative breast cancer (TNBC), however, the usage is largely limited due to its side effect and rapidly developed drug resistance. High prevalence of BRCA1/2 mutations are reported in TNBC. Here we explored efficacy of platinum-based regimens as the first-line treatment for Chinese patients (pts) with advanced TNBC, and analyzed its association with mutations of germline BRCA1/2 (gBRCA). Methods: We retrospectively analyze 220 patients diagnosed as advanced TNBC who were treated at the Dept. of Breast Oncology, Peking University Cancer Hospital in 2013-2018, and routinely evaluated by RECIST 1.1. Statistical analysis is performed in R 3.5.1. Cox proportional-hazard models are used for survival analysis. Results: 129 pts received non-platinum chemotherapy (NPCT) as the first-line therapy, and 91 pts received platinum-based chemotherapy (PBCT). The clinical benefit rate (CBR) and median PFS were not statistically different between NPCT and PBCT groups The median OS was 30.0 and 22.5 months for PBCT and NPCT group respectively (P = 0.09, HR = 0.70). Among them, 114 pts had BRCA gene tested, of which 14 had deleterious gBRCA mutations, 7 in each group. In PBCT group, the CBR was 85.7% and 35.1% for pts with and without deleterious gBRCA mutations respectively (P = 0.04). The median PFS, OS were 14.9 months and 5.3months (P = 0.001), 26.5 and 15.5 months (P = 0.16) for pts with and without the gBRCA mutations. No significant difference was observed between NPCT pts with and without gBRCA mutations as regards the CBR, PFS and OS. PBCT Pts had significantly more grade 3-4 anaemia (5.5% vs 0%) and thrombocytopenia (8.8% vs 0%) while higher percentage of palmar-plantar erythrodysesthesia (PPE) (12.4% vs 0%) and peripheral neuropathy (8.6% vs 1.1%) were observed in NPCT pts. Conclusions: The efficacy of platinum-based regimens as the first-line treatment of advanced TNBC insignificantly differs from that of non-platinum therapy. However, they are more effective for patients with deleterious gBRCA mutations, suggesting a BRCA1/2 genetic testing may be warranted for such patients.


2019 ◽  
Vol 37 (15_suppl) ◽  
pp. e15004-e15004
Author(s):  
Tao Jiang ◽  
Xiaoyan Lin ◽  
Hao Chen ◽  
Jianwei Zhen ◽  
Bin Du ◽  
...  

e15004 Background: Cetuximab plus FOLFIRI (leucovorin, fluorouracil and irinotecan) is preferred first line therapy for RAS and BRAF wild-type (RBWT) metastatic colorectal cancer (mCRC) patients. However, the side effects often require dose-reductions or discontinuation calling for maintenance strategies to reduce toxicity and preserve efficacy. Therefore, we evaluated efficacy and safety of cetuximab maintenance therapy after first-line treatment in unresectable mCRC patients with RBWT. Methods: This single center, retrospective study enrolled patients (aged 18-71 years) with untreated mCRC with RBWT, Eastern Cooperative Oncology Group (ECOG) performance status ≤ 2, with at least one measurable lesion according to response evaluation criteria in solid tumor (RECIST). All patients received first line cetuximab plus FOLFIRI every two weeks for 9-12 cycles, after which, patients with treatment response chose to either enter observation (stop treatment) or maintenance 1 (cetuximab plus irinotecan) group. After 6-12 cycles of maintenance 1, patients with treatment response entered maintenance 2 (cetuximab only). If patients progressed on maintenance 2, reintroduction of cetuximab plus FOLFIRI was suggested. Primary endpoint was failure-free survival (FFS); while secondary endpoints included disease control rate (DCR), objective remission rate (ORR) and progression free survival (PFS). Safety events were also evaluated. Results: Among 62 enrolled patients, 56 patients completed first-line treatment (DCR: 90.3%, ORR: 62.9%) and 30 patients entered maintenance 1 [median PFS1 (mPFS) =6.1 months, 95%CI: 6.0-6.2, DCR=60%, ORR=30%]. Of these, 18 entered maintenance 2 [mPFS2=6.6 months, 95%CI: 5.1-8.1, DCR=27.8%, ORR=5.6%]. The mFFS was significantly longer in maintenance 1 (12.8 months, 95%CI: 10.8-14.8) compared to observation group (3.0 months; hazard ratio [HR]: 0.21, 95%CI: 2.6-3.4, P<0.001). Overall mFFS was 19.0 (95%CI: 11.0-27.0) in maintenance and 9.5 months in observation group (HR: 0.195, 95%CI: 7.4-11.6; P<0.001). Rash acneiform, mucositis and asthenia were the common adverse events observed during maintenance treatment. Conclusions: Maintenance treatment with cetuximab after first line therapy significantly improved FFS, with acceptable safety profile in untreated mCRC RBWT patients.


Blood ◽  
2016 ◽  
Vol 128 (22) ◽  
pp. 1788-1788
Author(s):  
Francisco Javier Peñalver Párraga ◽  
José Antonio Márquez Navarro ◽  
Soledad Durán Nieto ◽  
Pilar Giraldo Castellano ◽  
Carlos Montalbán Sanz ◽  
...  

Abstract Objectives: To evaluate the efficacy and toxicity of a response-adapted therapy with rituximab/bendamustine/mitoxantrone/dexamethasone (RBMD), followed by rituximab (R) maintenance therapy in patients with relapsed or refractory follicular lymphoma (R/R FL) to first-line treatment with R-chemotherapy (R-ChemoT). Material and methods: Multicenter phase II trial including 60 patients with R/R FL, after a first R-ChemoT line. Induction therapy: R 375 mg/m2 IV, day 1; bendamustine 90 mg/m2 IV, days 1 and 2; mitoxantrone 6 mg/m2 IV, day 1; dexamethasone 20 mg/day, PO, days 1 to 5. Cycles of 28 days. Evaluation of response after third cycle. If stable (SD) or progression disease (PD): patient withdrawn from the study. If complete response (CR) or unconfirmed complete response (CRu): administration of fourth cycle. If partial response (PR): administration up to 6 cycles. If CR, CRu, or PR after induction: patient received maintenance therapy with R (375 mg/m2/day every 12 weeks for 2 years). Results: N=60 patients: 50% female, age 63 (32-76) years. Ann Arbor stage III-IV 70% (42/60). FLIPI intermediate-high risk: 50% (30/60). Refractory to R-ChemoT: 18% (11/60). Received RCHOP as first line therapy: 77% (43/60). R maintenance after first line therapy: 43% (26/60). Number of administered RBMD cycles: 4 (1-6). Efectiveness: CR/CRu after 3 cycles (CR-3), 27 patients. Response after induction therapy (4-6 cycles): Overall response (OR), 88.5% (53/60); CR, 58.5% (35/60); CRu, 12% (7/60); PR, 18% (11/60); SD, 1.5% (1/60); PD, 10% (6/60). Median follow-up: 24.41 mo. (15.58-30.15). Progression free survival, median not reached (NR) (28.28-NR). Overall survival, median NR (NR-NR). OR after RBMD in patients who received R maintenance after first line therapy, 96% (25/26; CR + CRu, 65%). OR in patients who did not receive R maintenance after first line therapy, 94% (32/34; CR + CRu, 82%). Safety: Grade 3/4 hematologic toxicity: neutropenia, 60% (n=36; 34 patients received G-CSF); anemia, 2% (n=1); thrombopenia, 4% (n=2). Grade 3/4 infections: 8% (n=4); febrile neutropenia 8%, (n=4). Exitus, 2/60 (PD, cycle 1; influenza A, cycle 5). Grade 3/4 non-hematologic toxicity: infusion reaction, 4% (n=2). No skin reactions. Patient withdrawals before R maintenance therapy: no inclusion criteria (NIC), 2/60; protocol delay of over 4 weeks, 7/60. Forty-four patients began R maintenance therapy (CR-3, 24/44). During R maintenance therapy: PD, 8/44 (18%; 5/8 in CR-3 patients); SD, 1/44 (2.5%); NIC, 3/44 (7%, all CR); protocol delay, 1/44 (2.5%); toxicity, 1/44 (2.5%, myelodysplastic syndrome). Response to R maintenance therapy: OR, 68% (30/44); CR, 61% (27/44; in CR-3 patients, 65%, 17/26); CRu, 4.5% (2/44); PR, 2.5% (1/44). Conclusions: RBMD is an effective and safe alternative for patients with R/R FL who have received first line treatment with R-ChemoT and R maintenance therapy. This response-adapted treatment strategy achieved results similar to the scheme with 6 cycles and may allow reduction of the intensity and duration of induction therapy and minimize toxicity. No skin reactions were reported, possibly due to the inclusion of dexamethasone in the treatment scheme. Disclosures No relevant conflicts of interest to declare.


Cancers ◽  
2018 ◽  
Vol 10 (11) ◽  
pp. 421 ◽  
Author(s):  
Teresa Steinbichler ◽  
Madeleine Lichtenecker ◽  
Maria Anegg ◽  
Daniel Dejaco ◽  
Barbara Kofler ◽  
...  

Background: Following first-line treatment of head and neck cancer (HNC), persistent disease may require second-line treatment. Methods: All patients with HNC treated between 2008 and 2016 were included. Second-line treatment modalities and survival of patients were analyzed. Results: After first-line therapy, 175/741 patients had persistent disease. Of these, 112 were considered eligible for second-line treatment. Second-line treatment resulted in 50% complete response. Median overall survival of patients receiving second-line therapy was 24 (95% CI: 19 to 29) months; otherwise survival was 10 (9 to 11; p < 0.0001) months. Patients receiving second-line surgery had a median overall survival of 45 (28 to 62) months, patients receiving second-line radiotherapy had a median overall survival of 37 (0 to 79; p = 0.17) months, and patients receiving systemic therapy had a median overall survival of 13 (10 to 16; p < 0.001) months. Patients with persistent HNC in the neck had a better median survival (45 months; 16 to 74 months; p = 0.001) than patients with persistence at other sites. Conclusion: Early treatment response evaluation allows early initiation of second-line treatment and offers selected patients with persistent disease a realistic chance to achieve complete response after all. If possible, surgery or radiotherapy are preferable.


2020 ◽  
Vol 38 (36) ◽  
pp. 4317-4345 ◽  
Author(s):  
John D. Gordan ◽  
Erin B. Kennedy ◽  
Ghassan K. Abou-Alfa ◽  
Muhammad Shaalan Beg ◽  
Steven T. Brower ◽  
...  

PURPOSE To develop an evidence-based clinical practice guideline to assist in clinical decision making for patients with advanced hepatocellular carcinoma (HCC). METHODS ASCO convened an Expert Panel to conduct a systematic review of published phase III randomized controlled trials (2007-2020) on systemic therapy for advanced HCC and provide recommended care options for this patient population. RESULTS Nine phase III randomized controlled trials met the inclusion criteria. RECOMMENDATIONS Atezolizumab + bevacizumab (atezo + bev) may be offered as first-line treatment of most patients with advanced HCC, Child-Pugh class A liver disease, Eastern Cooperative Oncology Group Performance Status (ECOG PS) 0-1, and following management of esophageal varices, when present, according to institutional guidelines. Where there are contraindications to atezolizumab and/or bevacizumab, tyrosine kinase inhibitors sorafenib or lenvatinib may be offered as first-line treatment of patients with advanced HCC, Child-Pugh class A liver disease, and ECOG PS 0-1. Following first-line treatment with atezo + bev, and until better data are available, second-line therapy with a tyrosine kinase inhibitor may be recommended for appropriate candidates. Following first-line therapy with sorafenib or lenvatinib, second-line therapy options for appropriate candidates include cabozantinib, regorafenib for patients who previously tolerated sorafenib, or ramucirumab (for patients with α-fetoprotein ≥ 400 ng/mL), or atezo + bev where patients did not have access to this option as first-line therapy. Pembrolizumab or nivolumab are also reasonable options for appropriate patients following sorafenib or lenvatinib. Consideration of nivolumab + ipilimumab as an option for second-line therapy and third-line therapy is discussed. Further guidance on choosing between therapy options is included within the guideline. Additional information is available at www.asco.org/gastrointestinal-cancer-guidelines .


Blood ◽  
2021 ◽  
Vol 138 (Supplement 1) ◽  
pp. 3605-3605
Author(s):  
Hiroyuki Shimada ◽  
Akihiro Watanabe ◽  
Masaki Ito ◽  
Chikako Tono ◽  
Haruko Shima ◽  
...  

Abstract Background: Tyrosine kinase inhibitor (TKI) has been used in pediatric chronic myeloid leukemia (CML) for more than 10 years, but only a few prospective clinical studies have been conducted in pediatric patients with CML due to their rarity. We conducted the JPLSG CML-08 study to determine the efficacy and tolerability of TKIs in children and adolescents with newly diagnosed CML in chronic phase (CML-CP). Methods: The JPLSG CML-08 study was a prospective multicenter observational study (UMIN000002581). Patients under 18 years of age with untreated BCR-ABL1-positive CML-CP were eligible and treated according to the modified ELN-2009 recommendation, and the efficacy and safety of TKIs were evaluated. Results: From October 2009 until September 2014, 79 patients were enrolled in 46 hospitals in Japan. A total of 78 patients (49 males and 29 females) were eligible for inclusion. Median age at diagnosis was 11 years (range, 1-17). Median observational period for survivors was 82 months (range, 48-118). Median WBC, Hb and platelet counts were 275x10 9/L (range, 8-765), 9.6g/dL (range, 5.8-14.6) and 560x10 9/L (range, 110-2875), respectively. Splenomegaly was found in 76%. High risk of Sokal, Hasford, EUTOS, and ELTS scores were observed in 21, 13, 27, and 9%, respectively. Clonal chromosome abnormalities in Ph-positive cells occurred in 1 patient at diagnosis. Imatinib, dasatinib, and, nilotinib were used as a first-line treatment in 69 (88%), 7 (9%), and 2 (3%) patients, respectively. The median initial dose of imatinib, dasatinib, and nilotinib was 276, 63, and 262mg/m2, respectively. 5y-PFS and OS was 96.2% (95%CI, 88.6 to 98.7%) and 97.4% (95%CI, 90.1 to 99.4%), respectively. Deaths were observed in 2 patients due to transplant complications. Hematopoietic cell transplantation was conducted in 14 patients (18%). Nine patients (12%) discontinued TKI with the aim of treatment-free remission (TFR), and five of them achieved TFR. In 69 patients with first-line imatinib, complete hematologic response was achieved in 95.7% at 3 months, complete cytogenetic response in 75.4% at 12 months, major molecular response (MMR) in 40.1% at 18 months, and MR4.0 in 52.8% at 60 months; If a transplant was performed, the follow-up period was censored at the date of transplant. Of the 69 patients, 52% changed treatment from imatinib to another TKI or transplant due to poor response, and 20% did due to intolerance. The most common cause of intolerance to imatinib was musculoskeletal events. BCR-ABL1 (IS) &lt;10% at 3 months strongly correlated with higher achievement of MMR, MR4.0, and MR4.5. The EUTOS score was significantly associated with achievement of IS &lt;10% at 3 months. Patients with a first-line second-generation TKI had a higher cumulative incidence of MR4.5 (P = 0.0191) than patients with a first-line imatinib. Second-generation TKI was used as first-line therapy only in patients older than 9 years, but other clinical characteristics, including risk scores, did not differ significantly between the two groups. The incidence of grade 3/4 adverse events (≥ 10%) included neutropenia (47%), anemia (39%), leukopenia (13%), arthralgia (13%), and myalgia (11%) for imatinib, neutropenia (21%), anemia (13%), and thrombocytopenia (11%) for dasatinib, and neutropenia (14%), elevated ALT (14%), hyperbilirubinemia (14%), skin rash (14%), and high CPK (14%) for nilotinib. Gastrointestinal bleeding was an adverse event specific to dasatinib (11% in all grades). Conclusion: This clinical study extends and confirms previous data showing that first-line treatment with imatinib is effective in children and adolescents, with response rates similar to those seen in adults. Although longer follow-up is needed to fully assess the long-term toxic effects, adverse events with imatinib, dasatinib, and nilotinib have been acceptable. As reported in adults, there was an advantage in selecting second-generation TKI over imatinib as first-line therapy to achieve deep molecular remission (DMR). Since discontinuation of TKI after achieving DMR is the preferred strategy, second-generation TKI is expected to become the standard therapy for children and adolescents. Disclosures No relevant conflicts of interest to declare.


Sign in / Sign up

Export Citation Format

Share Document