Erythropoietin Stimulates Thrombopoiesis in the Absence of C-Mpl Signaling

Blood ◽  
2008 ◽  
Vol 112 (11) ◽  
pp. 2451-2451
Author(s):  
Norma E Fox ◽  
Rose Chen ◽  
Ian Hitchcock ◽  
Kenneth Kaushansky ◽  
Amy Geddis

Abstract Thrombopoietin (TPO) is essential for normal megakaryopoiesis, and mice and humans lacking the TPO receptor c-Mpl have significantly impaired platelet production. However, in the c-Mpl-null mouse model platelet counts, while reduced to ~10% of normal, are not zero, suggesting that another cytokine is able to support some degree of residual thrombopoiesis. We and others have reported that elimination or severe reduction of stem cell factor, G-CSF, IL-3, IL-6 or IL-11 does not eliminate residual thrombopoiesis. Because megakaryocytes (MKs) and erythrocytes are derived from a common progenitor, we asked if erythropoietin (EPO) can stimulate thrombopoiesis in c-Mpl-null mice. We administered 90 u recombinant EPO or vehicle by subcutaneous injection every 3 days to c-Mpl-null or WT control mice and measured baseline and weekly platelet counts. In three independent experiments, at 2 weeks platelet counts in c-Mpl-null mice receiving EPO were significantly higher that at baseline (5–7 mice per group, average of mean platelet counts 425,000/mm3 vs. 285,000/mm3, p=0.0015). There was a trend towards higher platelet counts in WT mice receiving EPO but this did not reach statistical significance. No difference in platelet counts was observed in mice injected with vehicle. In one experiment c-Mpl-null or WT mice were injected with EPO for 4 weeks and the platelet response in the c-Mpl-null animals was sustained for the duration of the experiment. Western blotting showed that murine MKs express the EPO receptor. To determine if EPO stimulates MK production directly we stimulated WT murine MKs in vitro with either 6 u/ml EPO, 100 ng/ml rhTPO or both and monitored activation of ERK and STAT5 signaling by immunoblotting. Stimulation of MKs with EPO resulted in phosphorylation of ERK and STAT5 (15- and 14-fold above baseline, respectively), compared to TPO (97- and 75-fold above baseline). Stimulation with EPO and TPO together had an additive effect (phospho-ERK increased 121-fold and phospho-STAT5 increased 100-fold). To determine if EPO acts primarily on early or late MKs, we harvested bone marrow from c-Mpl-null mice after 2 weeks of treatment with EPO or vehicle and measured CFU-MK frequency and MK ploidy. Although there was a small increase in the frequency of CFU-MK in mice treated with EPO compared to vehicle, these differences were not significant (n=3, p=0.7), possibly due to the difficulty in assaying CFU-MK in vitro without TPO. In addition, EPO did not significantly enhance MK ploidy in c-Mpl-null mice, although MKs in the 32N and greater peaks were slightly more numerous. Therefore, we conclude that EPO can augment platelet production in the absence of c-Mpl signaling, although it is not yet clear if EPO primarily acts on early or late cells. Additional experiments are underway to determine if ablating EPO receptors in a TPO-null mouse model will eliminate residual thrombopoiesis. These findings may have clinical relevance for treating patients with congenital amegakaryocytic thrombocytopenia and other causes of thrombocytopenia in which c-Mpl signaling is impaired.

1996 ◽  
Vol 183 (3) ◽  
pp. 837-845 ◽  
Author(s):  
X Sui ◽  
K Tsuji ◽  
S Tajima ◽  
R Tanaka ◽  
K Muraoka ◽  
...  

Erythropoietin (EPO) is the primary humoral regulator of erythropoiesis and no other factor has previously been reported to support proliferation and terminal maturation of erythroid cells from hemopoietic stem cells. Here we show that stimulation of glycoprotein (gp130) by a combination of recombinant human soluble interleukin 6 receptor (sIL-6R) and IL-6 but not sIL-6R or IL-6 alone can support proliferation, differentiation, and terminal maturation of erythroid cells in the absence of EPO from purified human CD34+ cells in suspension culture containing stem cell factor (SCF). A number of erythroid bursts and mixed erythroid colonies also developed in methylcellulose culture under the same combination. The addition of anti-gp130 monoclonal antibodies but not anti-EPO antibody to the same culture completely abrogated the generation of erythroid cells. These results clearly demonstrate that mature erythroid cells can be emerged from hemopoietic progenitors without EPO in vitro. Together with the previous reports that human sera contain detectable levels of sIL-6R, IL-6, and SCF, current data suggest that gp130 signaling in association with c-kit activation may play a role in human erythropoiesis in vivo.


Blood ◽  
2009 ◽  
Vol 114 (22) ◽  
pp. 3622-3622
Author(s):  
Frederick Karl Racke ◽  
Maureen E Baird ◽  
Rolf Barth ◽  
Tianyao Huo ◽  
Weilian Yang ◽  
...  

Abstract Abstract 3622 Poster Board III-558 Despite recent advances in our understanding of megakaryocytic growth and platelet production, thrombocytopenia remains a difficult problem in the clinical management of patients with hematologic malignancies. Thrombopoietin (TPO) is the major cytokine involved in the normal production of platelets. However, the use of TPO has been relatively unsuccessful for the treatment of these patients and platelet transfusions remain the primary treatment for thrombocytopenia despite their significant cost and relatively short-lived responses. Thus, there remains an important clinical need for the development of novel approaches to generate platelets. Despite numerous reports on protein kinase C (PKC) agonists as promoters of megakaryocytic differentiation in leukemic cell lines and primary cells, little is known about their in vitro effects on primary CD34-selected progenitors or when administered in vivo. In the present study, we examine that effects of the novel PKC isoform agonist ingenol 3,20 dibenzoate (IDB) on megakaryocyte differentiation from CD34+ cells cultured in TPO and stem cell factor (SCF) or erythropoietin/SCF and its effects on platelet production in BALB/c mice. IDB potently stimulates early megakaryopoiesis and redirects the specificity of EPO to favor megakaryopoiesis over erythropoiesis. In contrast, broad spectrum PKC agonists such as phorbol myristate acetate, mezerein, and indolactam V fail to promote megakaryopoiesis. In vitro, IDB stimulates early expression of the promegakaryopoietic transcription factors egr1 and fli-1 and downregulates the proerythropoietic factors KLF1 and c-myb. Induction of the early megakaryocytic marker, CD9, was observed within the first 24 hrs of treatment with IDB and CD9 induction was blocked by the PKC inhibitor bisindolylmaleimide, which inhibits both novel and conventional PKC isoforms. In contrast, an inhibitor of conventional PKC isoforms, Gö6976, failed to block CD9 induction. In vivo, single intraperitoneal injections of IDB selectively increased platelet counts in BALB/c mice by 50% (plt= 630,000 vs. 985,000/μl; p<.005) at day 7 without affecting hemoglobin (Hgb) concentration or white counts (WBC). Mice treated with low dose radiation (2-4 Gy) had a transient drop in both platelet and WBC counts. Pretreatment with IDB 3 hrs prior to irradiation increased the platelet counts without improving WBC. More severe radiation exposure (6-8 Gy) causes pancytopenia. IDB treatment 3 hrs prior to 6 Gy irradiation significantly reduced the thrombocytopenia (plt=192,000 vs 594,000/μl; p<0.005) and anemia (hemoglobin=11.9 vs. 13.5gm/dl); p<0.005) without affecting the drop in WBC (WBC=1,200 vs. 1,300/μl; p=NS) at 14 days following irradiation. For mice treated with 8 Gy radiation, IDB pretreatment resulted in similar improvements in platelet counts (plt=111,000 vs. 443,000/μl; p<0.005) and hemoglobin (hgb=8.2 vs. 12.7 gm/dl; p<0.005) at 21 days following irradiation. The mitigation of thrombocytopenia is accompanied by marked increases in the megakaryocyte content in both the spleens and bone marrows of IDB-treated mice. Most importantly, IDB mitigated radiation-induced thrombocytopenia, even when administered 24 hrs after irradiation (plt=80,000 vs. 241,000/μl at 14 days following 6 Gy irradiation; p<0.01). Finally, IDB improved the survival of lethally irradiated mice. Our data suggest that the novel PKC isoform agonist IDB promotes the early differentiation of megakaryocytes from hematopoietic progenitors at the resulting in a significant improvement in platelet recovery following irradiation. IDB also improved Hgb levels following higher radiation doses. This may be due to improved hemostasis secondary to increased platelet numbers; however, an additional radioprotective effect on erythroid precursors cannot be excluded. These results strongly support our hypothesis that the novel PKC agonist IDB may be useful for the treatment of radiation and possibly drug-induced thrombocytopenia. Disclosures: No relevant conflicts of interest to declare.


Blood ◽  
2011 ◽  
Vol 118 (21) ◽  
pp. 3279-3279 ◽  
Author(s):  
Ann Janssens ◽  
Michael D. Tarantino ◽  
Robert Bird ◽  
Maria Gabriella Mazzucconi ◽  
Ralph Vincent V. Boccia ◽  
...  

Abstract Abstract 3279 Background: ITP is an autoimmune disorder characterized by increased platelet destruction and suboptimal platelet production. Romiplostim stimulates platelet production via the TPO-receptor, and is recommended for second- and third-line treatment of chronic ITP in adults. We report final data from a large prospective study of romiplostim in adults with ITP of varying duration and severity. Methods: Eligibility criteria were broad: patients ≥18 years of age, who had received prior ITP therapies (final protocol amendment: ≥1, previous amendments: ≥3), with low platelet counts (final amendment: ≤ 30 × 109/L, previous amendments: ≤ 10, ≤ 20 × 109/L) or experiencing uncontrolled bleeding. The only excluded comorbidities were: hematological malignancy, myeloproliferative neoplasms, MDS and bone marrow stem cell disorder. Romiplostim was initiated at 1 (final amendment) or 3 (previous amendments) μg/kg/week, with dose adjustments allowed to maintain platelet counts ≥50 × 109/L. Patients could continue on study until they had access to commercially available romiplostim. Rescue medications were allowed at any time; concurrent ITP therapies could be reduced when platelet counts were > 50 × 109/L. Primary endpoint was incidence of adverse events (AEs) and antibody formation. Secondary endpoint was platelet response, defined as either (1) doubling of baseline count and ≥ 50 × 109/L or (2) ≥20 × 109/L increase from baseline. Results: A total of 407 patients received romiplostim, 60% of whom were female. Median (Q1, Q3) time since ITP diagnosis was 4.25 (1.20, 11.40) years (maximum 57.1 years), with 51% of patients splenectomised and 39% receiving baseline concurrent ITP therapies. Seventy-one percent of patients completed the study, with requirement for alternative therapy and withdrawn consent the most common reasons for discontinuation (5% each). Median (Q1, Q3) on-study treatment duration was 44.29 (20.43, 65.86) weeks (maximum 201 weeks), with a total of 20,201 subject-weeks on study. Incidence and type of AEs were consistent with previous studies. The most common serious treatment-related AEs were cerebrovascular accident, headache, bone marrow reticulin fibrosis (with no evidence of positive trichrome staining for collagen and no evidence suggesting primary idiopathic myelofibrosis), nausea, deep vein thrombosis, hemorrhage and pulmonary embolism, with each reported in 2 of 407 (0.5%) patients. All other serious treatment-related AEs were each reported in one patient. Eighteen patients died; 3 deaths (hemolysis, intestinal ischaema, aplastic anemia) were considered treatment-related. No neutralizing antibodies to romiplostim or TPO were reported. Approximately 90% of patients achieved each of the platelet response definitions, regardless of splenectomy status. Overall, median (Q1, Q3) time to response was 2 (1, 4) weeks for response definition 1, and 1 (1, 3) week for response definition 2. Median (Q1, Q3) baseline platelet count was 14 (8, 21) × 109/L. After 1 week of treatment median (Q1, Q3) platelet count had increased to 42 (18, 101) × 109/L. From week 8 onwards, and excluding counts within 8 weeks of rescue medication use, median platelet counts were consistently above 100 × 109/L (range 101.0–269.5 × 109/L). Median (Q1, Q3) average weekly romiplostim dose was 3.62 (1.99, 6.08) μg/kg. Summary/conclusions: This is the largest prospective study in adult ITP reported to date. The data reported here are similar to those reported for previous romiplostim studies, with romiplostim able to safely induce a rapid platelet response in adult ITP patients with low platelet counts or bleeding symptoms. Romiplostim is an important, well-tolerated, treatment option for adult ITP patients, which significantly increases and maintains platelet counts. Adverse Event Subject Incidence Platelet Response Disclosures: Janssens: Amgen: Consultancy; Roche: Speakers Bureau; GSK: Membership on an entity's Board of Directors or advisory committees. Tarantino:Cangene corporation: Research Funding; Baxter: Research Funding; Talecris: Honoraria, Speakers Bureau; Up-to-date: Patents & Royalties; The Bleeding and Clotting Disorders Institute: Board Member. Bird:Amgen: Honoraria, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; Novartis: Honoraria, Membership on an entity's Board of Directors or advisory committees, Speakers Bureau; GSK: Membership on an entity's Board of Directors or advisory committees. Boccia:Amgen: Equity Ownership, Honoraria, Speakers Bureau. Lopez-Fernandez:Amgen: Membership on an entity's Board of Directors or advisory committees, Speakers Bureau. Kozak:Amgen: Membership on an entity's Board of Directors or advisory committees; Novartis: Membership on an entity's Board of Directors or advisory committees. Steurer:Amgen: Honoraria. Dillingham:Amgen Limited: Employment, Equity Ownership. Lizambri:Amgen: Employment, Equity Ownership.


Blood ◽  
2011 ◽  
Vol 118 (21) ◽  
pp. 2376-2376
Author(s):  
Minh-Ha T Do ◽  
Wei Zhang ◽  
Kyle Chiang ◽  
Chi-Fang Wu ◽  
Chulho Park ◽  
...  

Abstract Abstract 2376 Thrombopoietin (TPO) is recognized as the main regulator of platelet production, yet its genetic ablation in mice does not completely obliterate thrombopoiesis, suggesting that alternate pathways could lead to platelet formation. We recently identified a naturally-occurring protein that acts as a potent agonist of platelet production by a mechanism distinct from that of TPO. This protein belongs to a novel class of human extracellular signaling proteins called physiocrines that are generated from tRNA synthetases by alternative splicing or proteolysis. Physiocrines interact with several classes of receptors through unique mechanisms to modulate cellular differentiation and tissue homeostasis in normal and pathological processes. The newly identified thrombopoietic physiocrine, termed ATYR0030, is an engineered version of a naturally-occurring physiocrine derived from the tyrosyl tRNA synthetase (YRS). In vivo, systemic administration of ATYR0030 or YRS physiocrine to rats led to an increase in platelets counts comparable to that seen with TPO treatment, but with a greater effect in animals with low baseline platelet levels. When injected into normal animals preselected for low platelet counts, ATYR0030 treatment resulted in an increase in platelets up to, but not beyond, normal levels (Figure 1), suggesting a role in platelet homeostasis and differentiating its effects from the known activity of TPO. Intravenous administration of ATYR0030 also accelerated recovery of platelet counts in carboplatin-treated rats, indicating a possible role in bone marrow reconstitution after chemical insult. Consistent with homeostatic properties, no toxicity was seen in a repeat-dose 28-day non-GLP safety study in rats dosed up to 100-fold above the efficacious range. Histopathology assessment revealed no tissue abnormalities, no increase in bone marrow reticulin and no hyperplasia of myeloid precursors. Clinical chemistry and hematology parameters were in the normal range with a modest increase in platelet counts, as anticipated in animals with normal platelet levels. Our in vitro data suggest that ATYR0030 may play a role in megakaryopoiesis by facilitating cell migration and adhesion to the vasculature. In contrast to TPO, ATYR0030 does not directly signal through the TPO receptor and does not activate the JAK/STAT pathway but rather appears to engage specific G-protein coupled receptors. In vitro, ATYR0030 does not stimulate proliferation of cultured M07e human megakaryoblasts or primary bone marrow cells isolated from AML patients (Figure 2). The parent synthetase is present in human platelets and is secreted in response to platelet activation, perhaps providing a feedback mechanism to stimulate the release of new platelets. In an effort to link the biological activity of ATYR0030 and the role that the parent synthetase plays in human physiology, we have begun to analyze samples from patients with abnormal platelets counts to determine circulating levels of the parent synthetase. The unique thrombopoietic activity of ATYR0030 may lead to an orthogonal approach to restoring normal platelet levels in thrombocytopenic patients who currently have limited treatment options. For example, in the myelodysplastic syndrome population, TPO-receptor agonists carry a risk of stimulating blast proliferation and accelerating disease progression to acute myeloid leukemia (AML). The distinct proliferation profile of ATYR0030 may translate into important safety benefits by reducing the risk of progression to AML. In addition, the potential role of ATYR0030 in regulating platelet homeostasis may provide a greater safety margin in the normalization of platelet levels, thereby also limiting the risk of thrombosis. Leveraging the therapeutic potential of this thrombopoietic physiocrine may lead to the development of a novel treatment option with a favorable safety profile. Disclosures: Do: aTyr Pharma: Employment, Equity Ownership, Patents & Royalties. Zhang:aTyr Pharma: Employment, Equity Ownership. Chiang:aTyr Pharma: Employment, Equity Ownership. Wu:aTyr Pharma: Employment, Equity Ownership, Patents & Royalties. Park:aTyr Pharma: Equity Ownership. Yang:aTyr Pharma: Consultancy, Equity Ownership, Patents & Royalties, Research Funding. Kunkel:aTyr Pharma: Consultancy, Stock Ownership. Ashlock:aTyr Pharma: Employment, Equity Ownership. Mendlein:aTyr Pharma: Employment, Equity Ownership, Membership on an entity's Board of Directors or advisory committees, Patents & Royalties. Belani:Atyr Pahrma: Consultancy, Equity Ownership, Patents & Royalties. Vasserot:aTyr Pharma: Employment, Equity Ownership, Patents & Royalties. Watkins:aTyr Pharma: Employment, Equity Ownership, Membership on an entity's Board of Directors or advisory committees.


1992 ◽  
Vol 4 (1) ◽  
pp. 119 ◽  
Author(s):  
AM Lewis ◽  
PL Kaye ◽  
R Lising ◽  
RD Cameron

Present evidence indicates that insulin may act as a growth factor during preimplantation development. This hypothesis has been tested on pig blastocysts by determining the effect of insulin on protein synthesis and blastocyst expansion over 24 h. Blastocysts were collected from superovulated gilts or sows on Day 5 or 6 and incubated overnight in a modified BMOC2 medium. Those that were cultured with 1.7 nM insulin had 14% larger radii, and were 36% more active in their incorporation of [3H]leucine (protein synthesis) than those that had been cultured in non-supplemented medium. There was a significant linear correlation between the rate of protein synthesis and the radius of blastocysts when all blastocysts and only those cultured with insulin were examined, but the correlation for the blastocysts in non-supplemented medium was just outside statistical significance. The regression coefficient for the insulin-treated blastocysts was 132% of that for blastocysts cultured in unsupplemented medium; this suggests that insulin increased the size of blastocysts and the rate of protein synthesis per unit size. The results indicate that pig blastocysts respond to physiological levels of insulin in similar fashion to those of mice and cattle, supporting the hypothesis that insulin may act as a general embryonic growth factor. Because of the cross reaction between the insulin receptor and the ligands, insulin and insulin-like growth factor 1 (IGF-1), the results also suggest that IGF-1, reported to be present in pig uterine fluid, could be involved in this stimulation in utero.


Blood ◽  
2014 ◽  
Vol 124 (21) ◽  
pp. 1463-1463
Author(s):  
Karen Arkam ◽  
Sameer Doshi ◽  
Bing-Bing Yang

Abstract Background: Chronic Immune thrombocytopenia (ITP) is characterized by low platelet counts, resulting from increased platelet destruction and inadequate platelet production. Romiplostim is a 59 kDa peptibody which binds to and activates the thrombopoietin (TPO) receptor on platelet precursors in the bone marrow, and increases platelet counts. This analysis integrates the pharmacokinetic (PK) and pharmacodynamic (PD) properties of romiplostim in animals, healthy volunteers and patients with ITP, and describes its intricate PK-PD inter-relationship. Methods and Results: In healthy subjects, over a wide range of doses examined, the PK and PD (platelet response) of romiplostim were dependent on both the dose administered and the baseline platelet counts. Following SC administration, platelet counts increased in a dose-dependent fashion after 4 to 9 days, peaking at 12 to 16 days (Wang Clin Pharmacol Ther. 2004;76:628-38). When romiplostim binds to the TPO receptor on megakaryocytes and platelets, the peptibody-receptor complex is internalized and degraded inside the cells. Therefore, as platelet counts increase, a higher number of free receptors are available to clear romiplostim (Wang AAPS J. 2010;12:729-40). Results from rodent studies suggest that as the dose increases, the TPO receptors become saturated and the contribution of the kidney to clearance increases. Additionally, proteolysis plays a role in the clearance of romiplostim; however, the cytochrome P450 enzymes are not involved in protein catabolism (Wang Pharm Res. 2011;28:1931-8), hence there are no known drug-drug interactions or dietary restrictions (Nplate Prescribing Information 2014). Following SC administration, serum concentrations of romiplostim were markedly lower, however, platelet response was similar after the same dose of intravenous (IV) and SC administration (Wang Clin Pharmacol Ther. 2004;76:628-38). This suggests that the PD response is driven by the length of time that the romiplostim concentrations remained above a threshold rather than by the magnitude of concentrations achieved. This effect was verified in a mechanistic PK-PD modeling study in animals (Krzyzanski Pharm Res. 2013;30:655-69). In patients with ITP receiving SC romiplostim at a dose of 1 mcg/kg, the peak platelet response was achieved at 18 days (range 8 to 43; Bussel N Engl J Med. 2006;355:1672-81). Pharmacodynamic model analysis showed that compared with healthy subjects, patients with ITP had a shorter platelet life span and a decreased rate of production of progenitor cells, but no major difference in the time to maturation of megakaryocytes. The PD response in this modeling analysis was not notably affected by age, body weight, sex, and race (Perez-Ruixo J Clin Pharmacol. 2012;52:1540-51). The frequency of once-weekly dosing was selected because once every 2 weeks dosing was determined to be inadequate to achieve and maintain platelet counts in the therapeutic range (Bussel N Engl J Med. 2006;355:1672-81). A mechanistic PK-PD model based on data from the healthy subjects further suggested that weekly dosing resulted in a sustained platelet response while dosing less frequently resulted in high fluctuation of platelet counts (Wang AAPS J. 2010;12:729-40). Large inter- and intra-individual variability in the PD response was observed at a given dose; therefore, dose adjustments should be made based on a patient's platelet counts, using a titrated dosing scheme to prevent having platelet counts over 400 x 109/L (Perez-Ruixo J Clin Pharmacol. 2012;52:1540-51). Conclusion: Romiplostim is a peptibody that binds and activates the TPO receptor, and consequently increases platelet production in individuals with chronic ITP. The peptibody-receptor complex is internalized and degraded inside the cells, without involvement of the liver. Romiplostim's PD response is driven by the length of time that its concentrations remained above a threshold rather than by the magnitude of concentrations achieved. Moreover, weekly dosing has demonstrated a sustained platelet response while less frequent dosing resulted in fluctuating platelet counts. Disclosures Arkam: Amgen Inc.: Employment, Equity Ownership. Off Label Use: Romiplostim is a thrombopoietin receptor agonist indicated for the treatment of thrombocytopenia in patients with chronic immune thrombocytopenia (ITP) who have had an insufficient response to corticosteroids, immunoglobulins, or splenectomy. This abstract also describes PK data from healthy volunteers.. Doshi:Amgen Inc.: Employment, Equity Ownership. Yang:Amgen Inc.: Employment, Equity Ownership.


Blood ◽  
1990 ◽  
Vol 76 (8) ◽  
pp. 1473-1480
Author(s):  
AM Vannucchi ◽  
A Grossi ◽  
D Rafanelli ◽  
PR Ferrini

Murine recombinant granulocyte-macrophage colony-stimulating factor (rGM-CSF) was injected in mice, and the effects on bone marrow, splenic megakaryocytes, megakaryocyte precursors (megakaryocyte colony-forming units [CFU-Meg]) were evaluated. In mice injected three times a day for 6 days with 12,000 to 120,000 U rGM-CSF, no significant modification of both platelet levels and mean platelet volume was observed, while there was a twofold increase in blood neutrophils. However, the rate of platelet production, as assessed by the measurement of 75selenomethionine incorporation into blood platelets, was On the contrary, administration of up to 384,000 U rGM-CSF two times a day for 2 days, as for a typical “thrombopoietin assay,” failed to modify platelet production. A significant dose-related increase in the number of splenic megakaryocytes occurred in mice receiving 60,000 to 120,000 U rGM-CSF, while a slight increase in the number of bone marrow megakaryocytes was observed in mice injected with 120,000 U rGM-CSF. The proportion of bone marrow megakaryocytes with a size less than 18 microns and greater than 35 microns resulted significantly higher in mice receiving rGM-CSF in comparison with controls; an increase in the percentage of splenic megakaryocytes greater than 35 microns was also observed. A statistically significant increase in the total spleen content of CFU-Meg was observed after administration of 90,000 and 120,000 U rGM-CSF three times a day for 6 days, while no effect on bone marrow CFU-Meg was recorded, irrespective of the dose delivered. Finally, 24 hours after a single intravenous injection of rGM-CSF, there was a significant increase in the proportion of CFU-Meg in S- phase, with the splenic progenitors being more sensitive than bone marrow-derived CFU-Meg. These data indicate that rGM-CSF has in vivo megakaryocyte stimulatory activity, and are consistent with previous in vitro observations. However, an effective stimulation of megakaryocytopoiesis in vivo, bringing about an increase in the levels of blood platelets, may require interaction of rGM-CSF with other cytokines.


2006 ◽  
Vol 24 (18_suppl) ◽  
pp. 2541-2541
Author(s):  
M. Renshaw ◽  
S. Frederickson ◽  
B. Lin ◽  
X. Su ◽  
Y. Wang ◽  
...  

2541 Background: Thrombocytopenia, or low platelet count, is a significant clinical problem associated with ITP, cancer chemotherapy, or other clinical settings. To avoid transfusions, agents that stimulate platelet production are in development. Clinical trials with recombinant versions of thrombopoietin (TPO), which stimulates platelet production in a lineage specific manner by binding to cMpl receptor (cMpl-R) on megakaryocytic progenitors, were demonstrated to increase platelet counts in humans. However, the possible generation of an anti-TPO immune response that cross-reacts with and impairs the function of the endogenous cytokine is a significant disadvantage of this treatment approach. Alternate thrombopoietic agents that lack native TPO primary sequences have been developed to address this concern. Methods: Using rational design, antibody fragments (Fabs) that mimic TPO were created. A peptide with cMpl-R binding capability was grafted into different CDRs of a fully human Fab scaffold. Functional presentation of the peptide was optimized using phage display and cell-based panning. Select antibodies and fragments containing two grafted peptides were assayed for their ability to stimulate cMpl-R in vitro. In vivo stimulation of platelet production was tested in normal mice injected daily for five days with either rhTPO (90 μg/kg) or Fab59 (0.2 mg/kg, 2 mg/kg or 5 mg/kg) or negative control Fab (5mg/kg). Results: Several candidates demonstrated agonist activity in an in vitro cMpl-R signaling reporter assay, including Fab59 which was estimated to be equipotent to TPO. In vivo, a rise in peripheral platelet counts comparable to rhTPO was seen with Fab59 at 2 mg/kg and 5 mg/kg. Serum antibodies generated in response to dosing with Fab59 did not cross-react with murine or human TPO. Conclusion: These rationally-designed mimetic Fabs may provide a therapeutic intervention for thrombocytopenia while avoiding the potential generation of neutralizing antibodies to endogenous TPO. [Table: see text]


Blood ◽  
1990 ◽  
Vol 76 (8) ◽  
pp. 1473-1480 ◽  
Author(s):  
AM Vannucchi ◽  
A Grossi ◽  
D Rafanelli ◽  
PR Ferrini

Abstract Murine recombinant granulocyte-macrophage colony-stimulating factor (rGM-CSF) was injected in mice, and the effects on bone marrow, splenic megakaryocytes, megakaryocyte precursors (megakaryocyte colony-forming units [CFU-Meg]) were evaluated. In mice injected three times a day for 6 days with 12,000 to 120,000 U rGM-CSF, no significant modification of both platelet levels and mean platelet volume was observed, while there was a twofold increase in blood neutrophils. However, the rate of platelet production, as assessed by the measurement of 75selenomethionine incorporation into blood platelets, was On the contrary, administration of up to 384,000 U rGM-CSF two times a day for 2 days, as for a typical “thrombopoietin assay,” failed to modify platelet production. A significant dose-related increase in the number of splenic megakaryocytes occurred in mice receiving 60,000 to 120,000 U rGM-CSF, while a slight increase in the number of bone marrow megakaryocytes was observed in mice injected with 120,000 U rGM-CSF. The proportion of bone marrow megakaryocytes with a size less than 18 microns and greater than 35 microns resulted significantly higher in mice receiving rGM-CSF in comparison with controls; an increase in the percentage of splenic megakaryocytes greater than 35 microns was also observed. A statistically significant increase in the total spleen content of CFU-Meg was observed after administration of 90,000 and 120,000 U rGM-CSF three times a day for 6 days, while no effect on bone marrow CFU-Meg was recorded, irrespective of the dose delivered. Finally, 24 hours after a single intravenous injection of rGM-CSF, there was a significant increase in the proportion of CFU-Meg in S- phase, with the splenic progenitors being more sensitive than bone marrow-derived CFU-Meg. These data indicate that rGM-CSF has in vivo megakaryocyte stimulatory activity, and are consistent with previous in vitro observations. However, an effective stimulation of megakaryocytopoiesis in vivo, bringing about an increase in the levels of blood platelets, may require interaction of rGM-CSF with other cytokines.


Blood ◽  
2010 ◽  
Vol 116 (21) ◽  
pp. 3696-3696
Author(s):  
Kazunori Murai ◽  
Shugo Kowata ◽  
Akiko Abo ◽  
Tatsuo Oyake ◽  
Kenichi Nomura ◽  
...  

Abstract Abstract 3696 Background: Bortezomib is potent and reversible proteasome inhibitor that has been extensively used for multiple myeloma. Several clinical studies demonstrated that overall response rates using bortezomib alone to relapsed or refractory patients with multiple myeloma were 33 to 50%. The most common grade 3 adverse event was a cyclic thrombocytopenia, which was reported in 20–30% of patients in several clinical studies. The mechanism by which bortezomib causes thrombocytopenia remains unknown. In this study, we evaluated the effect of bortezomib on megakaryocytic progenitor cells, megakaryocytopoiesis, megakaryocyte and platelet production in mice. Method: All animal procedures were approved by the Institutional Animal Care and Use Committee in Iwate Medical University. Male ddY at 8 weeks of age mice were used in all experiments. In vivo experiments: (a) The mice received 2.5 mg/kg bortezomib via tail-vein injection. Blood was obtained and the following experiments were carried out at day 2, 4, 6, 8, 10 after intravenous injection (n=9, each group). Complete blood counts were measured. Reticulated platelet (RP) was analyzed by flow cytometry using thiazole orange (TO) to evaluate platelet kinetics. Plasma TPO level were measured by ELISA. Bone marrow megakaryocyte's number and morphology from femur in bortezomib- and control-treated mice were observed by microscopy. Femur was fixed in 10% buffered formalin, decalcificated, embedded in paraffin and stained for Hematoxylin-Eosin (H-E). (b) Bortezomib (2.5 mg/kg) was administrated via tail-vein to mice. After 24hr, bone marrow cells were cultured in MegaCult®-C at 5% CO2 and 20% O2for 7 days. The megakaryocytic colonies (CFU-Megs-in vivo) were counted. In vitro experiments: (c) Bone marrow cells, obtained from non-treated mice, were cultured at 37°C in 5% CO2and 20% O2 for 7 days with bortezomib (0.01, 0.1, 1, 10, 100 ng/ml). CFU-Megs were counted (CFU-Megs-in vitro). (d) Proplatelet formation: Murine megakarocytes were partially purified from bone marrow using BSA gradient. They were plated in 96 micro-well culture plates (300 megakaryocyte)well) and cultured in IMDM in duplicates, supplemented with 1 × ITS-G (Life technologies) and each concentration of bortezomib (0.01, 0.1, 1, 10, 100 ng/ml), at 37°C in 5% CO2and 20% O2. After 24 hr incubation, the megakaryocytes with proplatelets in each well were counted. Results: (a) Control mice did not have any significant change in platelet counts, % reticulated platelets and plasma TPO levels at days 0, 2, 4, 6, 8, 10. While, bortezomib treated mice (2.5mg/kg) had a significant reduction in platelet counts at day 2 (470 ± 210 × 109/L. P<0.001), at day 4 (667 ± 118 × 109/L, P<0.001). The platelet counts returned to normal value at day 6 (903 ± 548 × 109/L) and day 10 (1122 ± 187 × 109/L). RP (%) began to increase at day 6 (8.8 ± 4.0 %). Plasma TPO levels tend to increase at day 4. Means megakaryocytes's number in one field of femur was similar in between bortezomib non-treated and –treated mice. The megakaryocytes were similar in morphology at each day, too. (b) CFU-Megs-in vivo were similar in number between bortezomib non-treated and –treated mice (38.0 ± 6.1 vs 34.5 ± 5.6 per 1 × 105 bone marrow cells respectively). (c) CFU-Megs-in vitro were not decreased significantly at 0.001 to 1 ng/ml and decreased significantly (p<0.01) at 10 and 100 ng/ml of bortezomib. (d) Proplatelet formation (PPF) were decreased significantly at 0.01, 0.1, 1, 10, 100 ng/ml bortezomib (0 mg)ml: 25.2 ± 4.8%, 0.01ng/ml: 23.8 ± 4.9%, 0.1 ng/ml: 18.4 ± 3.1% p<0.01, 1 ng/ml: 13.2 ± 3.8% p<0.001, 10 ng/ml: 13.3 ± 2.1% p<0.001, 100ng/ml: 5.9 ± 1.4 % p<0.001). Discussion & Conclusion: Bortezomib did not adversely affect on megakaryocytic prognitors nor megakaryocytes. It did inhibit PPF, that is, the step of platelet production, even when bortezomib plasma concentration levels have gone down. Plasma TPO level showed an inverse relationship against circulating platelet counts. Based on the evidence in which Cmax of plasma bortezomib concentration was under 100 ng/ml in bortezomib-injected mice (2.5mg/kg), bortezomib induced thrombocytopenia might be due to the inhibition of proplatelet formation of megakaryocyte. Disclosures: No relevant conflicts of interest to declare.


Sign in / Sign up

Export Citation Format

Share Document