Posts Tagged ‘china-’

Addition of ruxolitinib to standard graft-versus-host disease prophylaxis for allogeneic stem cell transplantation in … – Nature.com

Georges GE, Doney K, Storb R. Severe aplastic anemia: allogeneic bone marrow transplantation as first-line treatment. Blood Adv. 2018;2:20208.

Article CAS PubMed PubMed Central Google Scholar

Zhao J, Ma L, Zheng M, Su L, Guo X. Meta-analysis of the results of haploidentical transplantation in the treatment of aplastic anemia. Ann Hematol. 2023;102:256587.

Article CAS PubMed Google Scholar

Esteves I, Bonfim C, Pasquini R, Funke V, Pereira NF, Rocha V, et al. Haploidentical BMT and post-transplant Cy for severe aplastic anemia: a multicenter retrospective study. Bone Marrow Transplant. 2015;50:6859.

Article CAS PubMed Google Scholar

Xu LP, Jin S, Wang SQ, Xia LH, Bai H, Gao SJ, et al. Upfront haploidentical transplant for acquired severe aplastic anemia: registry-based comparison with matched related transplant. J Hematol Oncol. 2017;10:25.

Article PubMed PubMed Central Google Scholar

Verstovsek S, Mesa RA, Gotlib J, Levy RS, Gupta V, DiPersio JF, et al. A double-blind, placebo-controlled trial of ruxolitinib for myelofibrosis. N Engl J Med. 2012;366:799807.

Article CAS PubMed PubMed Central Google Scholar

Mart-Carvajal AJ, Anand V, Sol I. Janus kinase-1 and Janus kinase-2 inhibitors for treating myelofibrosis. Cochrane Database Syst Rev. 2015;2015:CD010298.

PubMed PubMed Central Google Scholar

Spoerl S, Mathew NR, Bscheider M, Schmitt-Graeff A, Chen S, Mueller T, et al. Activity of therapeutic JAK 1/2 blockade in graft-versus-host disease. Blood. 2014;123:383242.

Article CAS PubMed Google Scholar

Schroeder MA, Choi J, Staser K, DiPersio JF. The role of Janus kinase signaling in graft-versus-host disease and graft versus leukemia. Biol Blood Marrow Transplant. 2018;24:112534.

Article CAS PubMed Google Scholar

Hechinger AK, Smith BA, Flynn R, Hanke K, McDonald-Hyman C, Taylor PA, et al. Therapeutic activity of multiple common -chain cytokine inhibition in acute and chronic GVHD. Blood. 2015;125:57080.

Article CAS PubMed PubMed Central Google Scholar

Choi J, Ziga ED, Ritchey J, Collins L, Prior JL, Cooper ML, et al. IFNR signaling mediates alloreactive T-cell trafficking and GVHD. Blood. 2012;120:4093103.

Article CAS PubMed PubMed Central Google Scholar

Carniti C, Gimondi S, Vendramin A, Recordati C, Confalonieri D, Bermema A, et al. Pharmacologic inhibition of JAK1/JAK2 signaling reduces experimental murine acute GVHD while preserving GVT effects. Clin Cancer Res. 2015;21:37409.

Article CAS PubMed Google Scholar

Zeiser R, von Bubnoff N, Butler J, Mohty M, Niederwieser D, Or R, et al. Ruxolitinib for glucocorticoid-refractory acute graft-versus-host disease. N Engl J Med. 2020;382:180010.

Article PubMed Google Scholar

Zeiser R, Polverelli N, Ram R, Hashmi SK, Chakraverty R, Middeke JM, et al. Ruxolitinib for glucocorticoid-refractory chronic graft-versus-host disease. N Engl J Med. 2021;385:22838.

Article CAS PubMed Google Scholar

Zeiser R, Burchert A, Lengerke C, Verbeek M, Maas-Bauer K, Metzelder SK, et al. Ruxolitinib in corticosteroid-refractory graft-versus-host disease after allogeneic stem cell transplantation: a multicenter survey. Leukemia. 2015;29:20628.

Article CAS PubMed PubMed Central Google Scholar

Krger N, Shahnaz Syed Abd Kadir S, Zabelina T, Badbaran A, Christopeit M, Ayuk F. et al. Peritransplantation ruxolitinib prevents acute graft-versus-host disease in patients with myelofibrosis undergoing allogenic stem cell transplantation. Biol Blood Marrow Transplant. 2018;24:21526.

Article PubMed Google Scholar

Gooptu M, Antin JH. GVHD prophylaxis 2020. Front Immunol. 2021;12:605726.

Article CAS PubMed PubMed Central Google Scholar

Borowitz MJ, Craig FE, Digiuseppe JA, Illingworth AJ, Rosse W, Sutherland DR, et al. Guidelines for the diagnosis and monitoring of paroxysmal nocturnal hemoglobinuria and related disorders by flow cytometry. Cytom B Clin Cytom. 2010;78:21130.

Article Google Scholar

Donohue RE, Marcogliese AN, Sasa GS, Elghetany MT, Redkar AA, Bertuch AA, et al. Standardized high-sensitivity flow cytometry testing for paroxysmal nocturnal hemoglobinuria in children with acquired bone marrow failure disorders: a single center US study. Cytom B Clin Cytom. 2018;94:699704.

Article CAS Google Scholar

Zhang Y, Huo J, Liu L, Shen Y, Chen J, Zhang T, et al. Comparison of hematopoietic stem cell transplantation outcomes using matched sibling donors, haploidentical donors, and immunosuppressive therapy for patients with acquired aplastic anemia. Front Immunol. 2022;13:837335.

Article CAS PubMed PubMed Central Google Scholar

Han TT, Xu LP, Liu DH, Liu KY, Wang FR, Wang Y, et al. Recombinant human thrombopoietin promotes platelet engraftment after haploidentical hematopoietic stem cell transplantation: a prospective randomized controlled trial. Ann Hematol. 2015;94:11728.

Article CAS PubMed Google Scholar

Tang B, Huang L, Liu H, Cheng S, Song K, Zhang X, et al. Recombinant human thrombopoietin promotes platelet engraftment after umbilical cord blood transplantation. Blood Adv. 2020;4:382939.

Article CAS PubMed PubMed Central Google Scholar

Chinese Society of Hematology, Chinese Medical Association. [Chinese expert consensus on the management of hemorrhagic complications after hematopoietic stem cell transplantation(2021)]. Zhonghua Xue Ye Xue Za Zhi. 2021;42:27680.

Wang J, Zhou M, Xu JY, Zhou RF, Chen B, Wan Y. Comparison of antifungal prophylaxis drugs in patients with hematological disease or undergoing hematopoietic stem cell transplantation: a systematic review and network meta-analysis. JAMA Netw Open. 2020;3:e2017652.

Article PubMed PubMed Central Google Scholar

Przepiorka D, Weisdorf D, Martin P, Klingemann HG, Beatty P, Hows J, et al. 1994 Consensus Conference on Acute GVHD Grading. Bone Marrow Transplant. 1995;15:8258.

CAS PubMed Google Scholar

Filipovich AH, Weisdorf D, Pavletic S, Socie G, Wingard JR, Lee SJ, et al. National Institutes of Health consensus development project on criteria for clinical trials in chronic graft-versus-host disease: I. Diagnosis and staging working group report. Biol Blood Marrow Transplant. 2005;11:94556.

Article PubMed Google Scholar

Champlin RE, Horowitz MM, van Bekkum DW, Camitta BM, Elfenbein GE, Gale RP, et al. Graft failure following bone marrow transplantation for severe aplastic anemia: risk factors and treatment results. Blood. 1989;73:60613.

Article CAS PubMed Google Scholar

Lin F, Han T, Zhang Y, Cheng Y, Xu Z, Mo X, et al. The incidence, outcomes, and risk factors of secondary poor graft function in haploidentical hematopoietic stem cell transplantation for acquired aplastic anemia. Front Immunol. 2022;13:896034.

Article CAS PubMed PubMed Central Google Scholar

McLornan DP, Hernandez-Boluda JC, Czerw T, Cross N, Joachim Deeg H, Ditschkowski M, et al. Allogeneic haematopoietic cell transplantation for myelofibrosis: proposed definitions and management strategies for graft failure, poor graft function and relapse: best practice recommendations of the EBMT Chronic Malignancies Working Party. Leukemia. 2021;35:244559.

Article PubMed Google Scholar

Li J, Wang Y, Zhang Y, Zhang X, Pang A, Yang D, et al. Haematopoietic stem cell transplantation for hepatitis-associated aplastic anaemia and non-hepatitis-associated aplastic anaemia: a propensity score-matched analysis. Br J Haematol. 2023;201:117991.

Article PubMed Google Scholar

Aggarwal N, Manley AL, Chen J, Groarke EM, Feng X, Young NS. Effects of ruxolitinib on murine regulatory T cells are immune-context dependent. Exp Hematol. 2023;125126:169.

Article PubMed Google Scholar

Xu ZL, Huang XJ. Optimizing outcomes for haploidentical hematopoietic stem cell transplantation in severe aplastic anemia with intensive GVHD prophylaxis: a review of current findings. Expert Rev Hematol. 2021;14:44955.

Article CAS PubMed Google Scholar

Williams L, Cirrone F, Cole K, Abdul-Hay M, Luznik L, Al-Homsi AS. Post-transplantation cyclophosphamide: from HLA-haploidentical to matched-related and matched-unrelated donor blood and marrow transplantation. Front Immunol. 2020;11:636.

Article CAS PubMed PubMed Central Google Scholar

Bourgeois AL, Jullien M, Garnier A, Peterlin P, Bn MC, Guillaume T, et al. Post-transplant cyclophosphamide as sole GHVD prophylaxis after matched reduced-intensity conditioning allotransplant. Clin Transl Med. 2023;13:e1242.

Article CAS PubMed PubMed Central Google Scholar

Prata PH, Eikema DJ, Afansyev B, Bosman P, Smiers F, Diez-Martin JL, et al. Haploidentical transplantation and posttransplant cyclophosphamide for treating aplastic anemia patients: a report from the EBMT Severe Aplastic Anemia Working Party. Bone Marrow Transplant. 2020;55:10508.

Article CAS PubMed Google Scholar

Chang YJ, Zhao XY, Huang XJ. Granulocyte colony-stimulating factor-primed unmanipulated haploidentical blood and marrow transplantation. Front Immunol. 2019;10:2516.

Article CAS PubMed PubMed Central Google Scholar

Yao D, Tian Y, Li J, Li B, Lu J, Ling J, et al. Association between haploidentical hematopoietic stem cell transplantation combined with an umbilical cord blood unit and graft-versus-host disease in pediatric patients with acquired severe aplastic anemia. Ther Adv Hematol. 2022;13:20406207221134409.

Article CAS PubMed PubMed Central Google Scholar

Xu ZL, Huang XJ. Haploidentical transplants with a G-CSF/ATG-based protocol: experience from China. Blood Rev. 2023;62:101035.

Article CAS PubMed Google Scholar

Xu LP, Lu DP, Wu DP, Jiang EL, Liu DH, Huang H, et al. Hematopoietic stem cell transplantation activity in China 20202021 during the SARS-CoV-2 pandemic: a report from the Chinese Blood and Marrow Transplantation Registry Group. Transplant Cell Ther. 2023;29:136 e1e7.

Article PubMed Google Scholar

Yoshimi A, Baldomero H, Horowitz M, Szer J, Niederwieser D, Gratwohl A, et al. Global use of peripheral blood vs bone marrow as source of stem cells for allogeneic transplantation in patients with bone marrow failure. JAMA. 2016;315:198200.

Article PubMed Google Scholar

Kennedy GA, Varelias A, Vuckovic S, Le Texier L, Gartlan KH, Zhang P, et al. Addition of interleukin-6 inhibition with tocilizumab to standard graft-versus-host disease prophylaxis after allogeneic stem-cell transplantation: a phase 1/2 trial. Lancet Oncol. 2014;15:14519.

Article CAS PubMed Google Scholar

Chen YB, Shah NN, Renteria AS, Cutler C, Jansson J, Akbari M, et al. Vedolizumab for prevention of graft-versus-host disease after allogeneic hematopoietic stem cell transplantation. Blood Adv. 2019;3:413646.

Article CAS PubMed PubMed Central Google Scholar

Lin Y, Gu Q, Lu S, Pan Z, Yang Z, Li Y, et al. Ruxolitinib improves hematopoietic regeneration by restoring mesenchymal stromal cell function in acute graft-versus-host disease. J Clin Investig. 2023;133:e162201.

Ryu DB, Lim JY, Kim TW, Shin S, Lee SE, Park G, et al. Preclinical evaluation of JAK1/2 inhibition by ruxolitinib in a murine model of chronic graft-versus-host disease. Exp Hematol. 2021;98:3646.e2.

Article CAS PubMed Google Scholar

Hlsdnker J, Ottmller KJ, Neeff HP, Koyama M, Gao Z, Thomas OS, et al. Neutrophils provide cellular communication between ileum and mesenteric lymph nodes at graft-versus-host disease onset. Blood. 2018;131:185869.

Article PubMed PubMed Central Google Scholar

Heine A, Held SA, Daecke SN, Wallner S, Yajnanarayana SP, Kurts C, et al. The JAK-inhibitor ruxolitinib impairs dendritic cell function in vitro and in vivo. Blood. 2013;122:1192202.

Article CAS PubMed Google Scholar

Sylvine P, Thomas S, Pirayeh E. Infections associated with ruxolitinib: study in the French Pharmacovigilance database. Ann Hematol. 2018;97:9134.

Article PubMed Google Scholar

Abedin S, McKenna E, Chhabra S, Pasquini M, Shah NN, Jerkins J, et al. Efficacy, toxicity, and infectious complications in ruxolitinib-treated patients with corticosteroid-refractory graft-versus-host disease after hematopoietic cell transplantation. Biol Blood Marrow Transplant. 2019;25:168994.

Article CAS PubMed Google Scholar

Hong X, Chen Y, Lu J, Lu Q. Addition of ruxolitinib in graft-versus-host disease prophylaxis for pediatric -Thalassemia major patients after allogeneic stem cell transplantation: a retrospective cohort study. Pediatr Transplant. 2023;27:e14466.

Article CAS PubMed Google Scholar

Zhang B, Chen L, Zhou J, Zu Y, Gui R, Li Z, et al. Ruxolitinib early administration reduces acute GVHD after alternative donor hematopoietic stem cell transplantation in acute leukemia. Sci Rep. 2021;11:8501.

Article CAS PubMed PubMed Central Google Scholar

Zhao Y, Shi J, Luo Y, Gao F, Tan Y, Lai X, et al. Calcineurin inhibitors replacement by ruxolitinib as graft-versus-host disease prophylaxis for patients after allogeneic stem cell transplantation. Biol Blood Marrow Transplant. 2020;26:e128e33.

Article CAS PubMed Google Scholar

Morozova EV, Barabanshikova MV, Moiseev IS, Shakirova AI, Barhatov IM, Ushal IE, et al. A prospective pilot study of graft-versus-host disease prophylaxis with post-transplantation cyclophosphamide and ruxolitinib in patients with myelofibrosis. Acta Haematol. 2021;144:15865.

Article CAS PubMed Google Scholar

See the rest here:
Addition of ruxolitinib to standard graft-versus-host disease prophylaxis for allogeneic stem cell transplantation in ... - Nature.com

China Genetic Testing Analysis Report 2024: Market to Reach $14.9 Billion by 2032 from $4.3 Billion in 2023, Driven … – PR Newswire

DUBLIN, March 27, 2024 /PRNewswire/ -- The"China Genetic Testing Market Report by Test Type, Disease, Technology, Service Provider, Testing Sample 2024-2032" report has been added toResearchAndMarkets.com's offering.

The China genetic testing market size reached US$ 4.3 billion in 2023. The market is projected to reach US$ 14.9 billion by 2032, exhibiting a growth rate (CAGR) of 14.9% during 2023-2032.

Genetic testing is becoming popular in China. It may benefit many different interest groups, such as individuals and families with a history of genetic disorder, pregnant women, employers, and health or life insurance. This market is currently driven by a number of factors such as rising awareness regarding the benefits of genetic testing, availability of direct to consumer tests and increasing incidences of genetic disorders.

Over the past few years, there has been a significant rise in the awareness levels regarding the benefits of genetic testing in China. Genetic testing provides various technologies that help in the early detection of various chronic diseases and ensures its treatment and prevention. Moreover, a rise in the availability of Direct to consumer tests (DTC) which has increased the convenience and accessibility of such tests is also creating a positive impact in the growth of the market.

Moreover, In October, 2015, China announced that the iconic one-child policy had finally been replaced by a universal two-child policy. This is expected to increase the number of babies born each year and create a positive impact on the demand of the new born genetic testing segment. Other major factors that are expected to drive this market include growing middle class, aging population, and expanding healthcare system.

This report provides a deep insight into the China genetic testing market covering all its essential aspects. This ranges from macro overview of the market to micro details of the industry performance, recent trends, key market drivers and challenges, SWOT analysis, Porter's five forces analysis, value chain analysis, etc. This report is a must-read for entrepreneurs, investors, researchers, consultants, business strategists, and all those who have any kind of stake or are planning to foray into the China genetic testing industry in any manner.

Key Questions Answered in This Report

Competitive Landscape

Key Market Segmentation:

Breakup by Test Type:

Breakup by Disease:

Breakup by Technology:

Breakup by Service Provider:

Breakup by Testing Sample:

For more information about this report visit https://www.researchandmarkets.com/r/ob8wjd

About ResearchAndMarkets.com ResearchAndMarkets.com is the world's leading source for international market research reports and market data. We provide you with the latest data on international and regional markets, key industries, the top companies, new products and the latest trends.

Media Contact:Research and Markets Laura Wood, Senior Manager [emailprotected] For E.S.T Office Hours Call +1-917-300-0470 For U.S./CAN Toll Free Call +1-800-526-8630 For GMT Office Hours Call +353-1-416-8900 U.S. Fax: 646-607-1907 Fax (outside U.S.): +353-1-481-1716

Logo: https://mma.prnewswire.com/media/539438/Research_and_Markets_Logo.jpg

SOURCE Research and Markets

Read the original here:
China Genetic Testing Analysis Report 2024: Market to Reach $14.9 Billion by 2032 from $4.3 Billion in 2023, Driven ... - PR Newswire

Microplastics dampen the self-renewal of hematopoietic stem cells by disrupting the gut microbiota-hypoxanthine-Wnt … – Nature.com

Mice

C57BL/6J (CD45.2) and C57BL6.SJL (CD45.1) mice were purchased from The Jackson Laboratory and housed under specific pathogen-free conditions. Male and female mice from 8 to 12 weeks were used in experiments and provided with a suitable environment and sufficient water and food. After a week of acclimatization, each mouse was randomly divided into groups, given 100L pure water, 0.01mg/100L, or 0.1mg/100L MPs by oral gavage every two days for five weeks in a gavage experiment (n=5 for each group). For the intravenous injection experiment, MPs were administered into mouse blood via the tail vein at a rate of 0.1g/100L per week for a duration of 4 weeks (n=5 for each group). All animal experiments were first approved by the Laboratory Animal Welfare and Ethics Committee of Zhejiang University (AP CODE: ZJU20220108).

Indocyanine green polystyrene (ICG-PS), polystyrene (PS) and polymethyl methacrylate (PMMA) particles were obtained from Suzhou Mylife Advanced Material Technology Company (China). Polyethylene (PE) particles were purchased from Cospheric (USA). Scanning electron microscopy (SEM, Nova Nano 450, FEI) was used to characterize the primary sizes and shapes of different MPs20. MPs were dispersed in ultrapure water with sonication before dynamic light scattering analysis (Zetasizer, Malvern, UK) to determine the hydrodynamic sizes and zeta potentials49.

Mice were sacrificed and organs were removed within six hours of ICG-PS gavage, including the heart, lung, kidney, spleen, liver, gastrointestinal tissues and bone marrow. Feces were collected 1h before the mice were sacrificed. Both organs and feces were monitored by ex vivo bioluminescence imaging with a small-animal imaging system50 (IVIS Spectrum, PerkinElmer).

For flow cytometry analysis and isolation of hematopoietic stem and progenitor cells, cells were stained with relevant antibodies51 in PBS with 2% fetal bovine serum for 3045min on ice. Antibody clones that were used: Sca-1-PE-Cy7, c-Kit-APC, CD150-PE, CD48-BV421, CD45.1-FITC, CD45.2 PE-Cy5, Gr-1-PE-Cy5, Mac1-PE-Cy5, IgM-PE-Cy5, CD3-PE-Cy5, CD4- PE-Cy5, CD8-PE-Cy5, CD45R-PE-Cy5 and Ter-119-PE-Cy5. Detailed antibody information is summarized in Supplementary Table S6. HSPCs were stained with a lineage antibody cocktail (Gr-1, Mac1, CD3, CD4, CD8, CD45R, TER119 and B220), Sca-1, c-Kit, CD150 and CD48. Cell types were defined as followed: LSK compartment (LinSca-1+c-Kit+), LT-HSC (LSK CD150+CD48), ST-HSC (LSK CD150CD48), MPP2 (LSK CD150+CD48+) and MPP3/4 (LSK CD150CD48+). B cells (CD45.2+Mac1Gr-1+B220+), T cells (CD45.2+Mac1Gr-1+CD3+) and myeloid cells (CD45.2+Mac1+Gr-1). Samples were analyzed on a flow cytometer (CytoFLEX LX, Beckman). For sorting HSCs, lineage antibody cocktail-conjugated paramagnetic microbeads and MACS separation columns (Miltenyi Biotec) were used to enrich Lin cells before sorting. Stained cells were re-suspended in PBS with 2% FBS and sorted directly using the Beckman moflo Astrios EQ (Beckman). Flow cytometry data were analyzed by FlowJo (BD) software.

Apoptosis of cells was detected by Annexin V staining (Yeason, China). After being extracted from the bone marrow of mice, 5106 cells were labeled with different surface markers for 30 to 45min at 4C and then twice rinsed with PBS. Subsequently, the cells were reconstituted in binding buffer and supplemented with Annexin V. After 30min of incubation, flow cytometry was detected in the FITC channel. Cell cycle analysis was performed with the fluorescein Ki-67 set (BD Pharmingen, USA), following the directions provided by the manufacturer. Briefly, a total of 5106 bone marrow cells were labeled with corresponding antibodies, as previously stated. Afterward, the cells were pre-treated with a fixation/permeabilization concentrate (Invitrogen, USA) at 4C overnight and subsequently rinsed with the binding buffer. The cells were stained with Ki-67 antibody for 1h in the dark and then with DAPI (Invitrogen) for another 5min at room temperature. Flow cytometry data were collected by a flow cytometer (CytoFLEX LX, Beckman, USA).

HSCs were sorted by flow cytometry according to the experimental group (ctrl and PSH mice, Rikenellaceae treatment or hypoxanthine treatment). 150 HSCs were seeded in triplicate on methylcellulose media52 (M3434, Stemcell Technologies, Inc.). After 8 days, the number of colonies was counted by microscopy. In addition, 5000 BM cells were seeded and analyzed the same way as HSCs. The cell culture media was diluted in PBS and subjected to centrifugation at 400g for 5min to determine the total cell number.

Recipient mice (CD45.1) were administered drinking water with Baytril (250mg/L) for 7 days pre-transplant and 10 days post-transplant. The day before transplantation, recipients received a lethal dose of radiation (4.5Gy at a time, divided into two times with an interval of 4h). In primary transplantation, 2105 bone marrow cells from the ctrl or PS group (CD45.2) mice and 2105 recipient-type (CD45.1) bone marrow cells were transplanted into recipient mice (CD45.1) mice. Cells were injected into recipients via tail vein injection. Donor chimerism was tracked using peripheral blood cells every 4 weeks for at least 16 weeks after transplantation. For secondary transplantation, donor BM cells were collected from primary transplant recipients sacrificed at 16 weeks after transplantation and transplanted at a dosage of 2106 cells into irradiated secondary recipient mice (9Gy). Analysis of donor chimerism and the cycle of transplantation in secondary transplantation were the same as in primary transplantation.

For limiting dilution assays52, 1104, 5104 and 2105 donor-derived bone marrow cells were collected from ctrl or PS mice (CD45.2) and transplanted into irradiated (9Gy) CD45.1 recipient mice with 2105 recipient-type (CD45.1) bone-marrow cells. Limiting dilution analysis was performed using ELDA software53. 16 weeks after transplantation, recipient mice with more than 1% peripheral-blood multilineage chimerism were defined as positive engraftment. On the other hand, recipient mice undergoing transplantation that had died before 16 weeks post transplantation were likewise evaluated as having failed engraftment54.

For histological analysis, small intestines were collected and fixed in 4% paraformaldehyde and embedded in paraffin, sectioned (5m thickness), and stained with H&E at ZJU Animal Histopathology Core Facility (China). We used Chius scores33,34 to evaluate the damage for each sample. The grade was as follows: 0, normal mucosa; 1, development of subepithelial Gruenhagens space at the tip of villus; 2, extension of the Gruenhagens area with moderate epithelial lifting; 3, large epithelial bulge with a few denuded villi; 4, denuded villi with lamina propria and exposed capillaries; and 5, disintegration of the lamina propria, ulceration, and hemorrhage. For TEM analysis, slices of the small intestine were fixed with 2.5% glutaraldehyde for ultra-microstructure observation of intestinal epithelial cells. The samples were postfixed for one hour at 4C with 1% osmium tetroxide and 30min with 2% uranyl acetate, followed by dehydration with a graded series of alcohol solutions (50%, 70%, 90% and 100% for 15min each) and acetone (100% twice for 20min). Subsequently, they were embedded with epon (Sigma-Aldrich, MO, US) and polymerized. Ultrathin sections (6080nm) were made, and examined using TEM (Tecnai G2 Spirit 120kV, Thermo FEI).

In the short-term and long-term mouse models for MP ingestion, mice were fasted for 4h before oral gavage of FITC-dextran (4kD, Sigma). The fluorescence intensity of FITC-dextran (50mg/100g body weight) was measured in the peripheral blood after 2h of gavage. Fluorescence was measured using a microplate reader (Molecular Devices, SpectraMax iD5) with excitation at 490nm and emission at 520 nm29.

Fecal samples (about 3050mg per sample) were collected from the ctrl, PSL and PSH mice, quickly frozen in liquid nitrogen, and stored at 80C. DNA samples for the microbial community were extracted using E.Z.N.A. Stool DNA Kit (Omega, USA), according to the manufacturers instructions. In brief, polymerase chain reaction (PCR) amplification of prokaryotic 16S rDNA gene V3V4 region was performed using the forward primer 341F (5-CCTACGGGNGGCWGCAG-3) and the reverse primer 805R (5-GACTACHVGGGTATCTAATCC-3)55. After 35 cycles of PCR, sequencing adapters and barcodes were included to facilitate amplification. The PCR products were detected by 1.5% agarose gel electrophoresis and were further purified using AMPure XT beads (Beckman Coulter Genomics, Danvers, MA, USA), while the target fragments were recovered using the AxyPrep PCR Cleanup Kit (Axygen, USA). In addition, the amplicon library was quantified with the Library Quantification Kit for Illumina (Kapa Biosciences, Woburn, MA, USA), and sequenced on the Illumina NovaSeq PE250 platform. In bioinformatics pipeline29,56, the assignment of paired-end reads to samples was determined by their unique barcode, and subsequently shortened by cutting off the barcode and primer sequence. The paired-end reads were combined by FLASH (v1.2.8). Quality filtering on the raw reads was carried out under precise parameters to obtain high-quality clean tags according to fqtrim (v0.94). The chimeric sequences were filtered by Vsearch software (v2.3.4). After the dereplication process using DADA2, we acquired a feature table and feature sequence. The bacterial sequence fragments obtained were grouped into Operational Taxonomic Units (OTUs) and compared to the Greengenes microbial gene database using QIIME2. Alpha diversity and beta diversity were generated by QIIME2, and pictures were drawn by R (v3.2.0). The species annotation sequence alignment was performed by Blast, with the SILVA and NT-16S databases as the alignment references. Additional sequencing results are provided in Supplementary Table S1. The experiment was supported by Lc-Bio Technologies.

The methods for the analysis of feces from HSCT donors were slightly different from those used for mice. All samples were stored in the GUHE Flora Storage buffer (GUHE Laboratories, China). The bacterial genomic DNA was extracted with the GHFDE100 DNA isolation kit (GUHE Laboratories, China) and quantified using a NanoDrop ND-1000 spectrophotometer (Thermo Fisher Scientific, USA). The V4 region of the bacterial 16S rDNA genes was amplified by PCR, with the forward primer 515F (5-GTGCCAGCMGCCGCGGTAA-3) and the reverse primer 806R (5-GGACTACHVGGGTWTCTAAT-3). PCR amplicons were purified with Agencourt AMPure XP Beads (Beckman Coulter, IN) and quantified by the PicoGreen dsDNA Assay Kit (Invitrogen, USA). Following the previously reported steps57, the paired-end 2150bp sequencing was performed on the Illumina NovaSeq6000 platform. The details of bacterial OTUs are summarized in Supplementary Table S5. Sequence data analyses were performed using QIIME2 and R packages (v3.2.0).

For metabolite evaluation, samples from mice feces were prepared and detected as previously described55,58,59. In a nutshell, metabolites were extracted from feces through precooled 50% methanol buffer and stored at 80C before the LCMS analysis. All chromatographic separations were conducted using an ultra-performance liquid chromatography (UPLC) system (SCIEX, UK). A reversed phase separation was performed using an ACQUITY UPLC T3 column (100mm * 2.1mm, 1.8m, Waters, UK). The temperature of the column oven was maintained at 35C and the flow rate was 0.4mL/min. Both positive (the ionspray voltage floating set at 5000V) and negative ion modes (4500V) were analyzed using a TripleTOF 5600 Plus high-resolution tandem mass spectrometer (SCIEX, UK). The mass spectrometry data were obtained in Interactive Disassembler Professional (IDA) mode, with a time-of-flight (TOF) mass range of 60 to 1200Da. The survey scans were acquired in 150 milliseconds and product ion scans with a charge state of 1+ and 100 counts per second (counts/s) were recorded up to 12. Cycle duration was 0.56s. Stringent quality assurance (QA) and quality control (QC) procedures were applied, as the mass accuracy was calibrated every 20 samples and a QC sample was obtained every 10 samples. LCMS raw data files underwent processing in XCMS (Scripps, La Jolla, CA) to perform peak picking, peak alignment, gap filling, and sample normalization. Online KEGG was adopted to annotate metabolites through the matching between the precise molecular mass data (m/z) of samples and those from the database. PCA and volcano plot were utilized to identify ion characteristics that exhibit significant differences between the groups. The details of metabolomes can be found in Supplementary Table S2. The experiment was supported by Lc-Bio Technologies.

Before FMT, SPF mice received a 200L antibiotic treatment (1g/L ampicillin, 0.5g/L neomycin, 0.5g/L vancomycin and 1g/L metronidazole) for three consecutive days by oral gavage. Fresh feces were collected from ctrl or PS mice and resuspended in reduced PBS (0.5g/L cysteine and 0.2g/L Na2S in PBS) at a ratio of about 120mg feces/mL reduced PBS. Feces were then centrifuged at 500g for 1min to remove insolubilize particles25. Recipients (C57BL/6J mice) were administered 100mL of the supernatant from different groups by oral gavage twice every week for 4 weeks. 2 days after the last FMT, recipients were euthanized to analyze the changes in the hematopoietic system.

The Rikenellaceae strain (ATCC BAA-1961), purchased from ATCC, was cultured in an anaerobic chamber using BD Difco Dehydrated Culture Media: Reinforced Clostridial Medium at a temperature of 37C with a gas mixture of 80% N2 and 20% CO2. The final concentration of Rikenellaceae was 2108 viable c.f.u. per 100L and hypoxanthine (200mg/kg, Sigma, Germany) was dissolved in double distilled water29. Mice first received antibiotic treatment (same as FMT) and were then treated by oral gavage with 100L of either Rikenellaceae or hypoxanthine suspension three times a week for 4 weeks. Reinforced Clostridial Medium or double distilled water was used as a vehicle control, respectively. 2 days after the last administration, recipients were euthanized to analyze the changes in the hematopoietic system. To examine the impact of hypoxanthine on HSCs, we exposed bone marrow cells to direct co-culture with hypoxanthine at a concentration of 100pg/mL for a period of 3 days.

Mouse bone marrow cells were harvested by flushing the mices tibia and femur in phosphate buffered saline (PBS) with 2% fetal bovine serum (GIBCO). Harvested cells were grown into 96-well u-bottom plates containing freshly made HSC culture medium (StemSpanTM SFEM, Stemcell Tec.) with SCF (50ng/mL; PeproTech) and TPO (50ng/mL; PeproTech), at 37C with 5% CO2. For HSC culture, the medium was changed every 3 days by manually removing half of the conditioned medium and replacing it with fresh medium60. To assess the effects of WNT10A, IL-17, TNF and NF-kappa B on hematopoiesis, we cultured HSCs in a basic medium and supplemented them with related proteins (10ng/mL; Cosmo Bio, USA) or PBS as a control for two days, followed by flow cytometry analysis. Different concentrations of PS were added to the medium and tested using CCK-8 and FACS to detect the effect of MPs on cultured HSCs.

1104 HSCs were obtained in triplicate from mouse bone marrow cells from the ctrl or PSH group by flow cytometry sorting and RNA was extracted with RNAiso Plus (Takara, Japan) according to the manufacturers protocol. The concentration and integrity of RNA were examined by Qubit 2.0 and Agilent 2100 (Novogene, China), respectively. Oligo (dT)-coated magnetic beads (Novogene, China) were used to enrich eukaryotic mRNA. After cDNA synthesis and PCR amplification, the PCR product was purified using AMPure XP beads (Novogene, China) to obtain the final library. The Illumina high-throughput sequencing platform NovaSeq 6000 was used for sequencing. Analysis of gene expression was calculated by R or the DESeq2 package61. Detailed information regarding RNA-seq is listed in Supplementary Table S3.

For RNA expression analysis, total RNA from bone marrow cells was extracted using Trizol (Invitrogen, US) and resuspended in nuclease-free water. Reverse transcription was performed using the QuantiTect Reverse Transcription kit (Qiagen NV). qPCR was conducted using cDNA, primers and SYBR-green (Takara, Japan) in 20L using the ABI 7500 Q-PCR system62. Results were calculated using the RQ value (RQ=2Ct). Mouse Actin was chosen as the normalization control. Gene-specific primer sequences are shown in Supplementary Table S7.

Bone marrow and Rikenellaceae supernatant in different groups were obtained by centrifugation. Fecal supernatant was obtained from human samples. Hypoxanthine (LANSO, China) and WNT10A (EIAab, China) were measured by ELISA with respective kits according to the manufacturers protocols.

Human feces and peripheral blood samples were obtained from 14 subjects who provided grafts for HSCT patients. They were divided into graft success group and graft failure (GS)/poor graft function (GF/PGF) group, with 7 participants in each group. Research involving humans was approved by the Clinical Research Ethics Committee of the First Affiliated Hospital, College of Medicine, Zhejiang University (IIT20230067B). All participants read and signed the informed consent. Detailed information on patients was listed in Supplementary Table S4.

The Agilent 8700 Laser Direct Infrared Imaging system was utilized for fast and automated analysis of MPs in feces received from donors. An excessive nitric acid concentration (68%) was added to the sample and heated to dissolve the protein. Large particles were first intercepted with a large aperture filter and then filtered by vacuum extraction. After rinsing with ultra-pure water and ethanol several times, the materials, including MPs, were dispersed in the ethanol solution. The LDIR test was carried out when the ethanol was completely volatilized63. The sample of MPs was positioned on the standard sample stage. The stage was then put into the sample stage, and the Agilent Clarity was initiated to advance the sample stage into the sample chamber. The software rapidly scanned the chosen test area using a constant wave number of 1800cm1, and accurately detected and pinpointed the particles within the selected area. The unoccupied area devoid of particles was automatically designated as the background. The background spectrum was gathered and readjusted, followed by the visualization of detected particles and the collection of the whole infrared spectrum. After obtaining the particle spectrum, the spectrum library was utilized to carry out qualitative analysis automatically, including the inclusion picture, size, and area of each particle. The test was supported by Shanghai WEIPU Testing Technology Group.

MPs in peripheral blood from donors were tested by Py-GC/MS. Nitric acid was added to samples for digestion at 110C for 12h, and then used deionized water to make the solution weakly acidic. After concentration, the solution was dribbled into the sampling crucible of Py-GCMS and tested when the solvent in the crucible was completely volatilized17. Various standards of MPs were prepared and analyzed using Py-GCMS in order to construct the quantitative curve. PY-3030D Frontier was employed for lysis, with a lysis temperature set at 550 C. The chromatographic column dimensions were 30m in length, 0.25mm inner diameter, and 0.25m film thickness. The sample was subjected to a heat preservation period of 2min at 40C, followed by a gradual increase in temperature at a rate of around 20C per minute until it reached 320C. The sample was maintained at this temperature for 14min and the entire process takes a total of 30min. The carrier gas utilized was helium, with the ion source temperature of 230C. The split ratio employed was 5:1, and the m/z scan range spanned from 40 to 60064. The experiment was supported by Shanghai WEIPU Testing Technology Group.

Each animal experiment was tested using at least 56 replicates and each in vitro experiment was at least three replicates. Specific replication details are provided in relevant figure captions. Statistical significance was ascertained through unpaired two-tailed t-tests by GraphPad Prism when the P value was less than 0.05. Error bars in all figures indicate the standard deviation (SD).

Link:
Microplastics dampen the self-renewal of hematopoietic stem cells by disrupting the gut microbiota-hypoxanthine-Wnt ... - Nature.com

Advanced Therapy Medicinal Products CDMO Industry is Rising Rapidly – BioSpace

According to latest study, the global advanced therapy medicinal products CDMO Market size was valued at USD 6.10 billion in 2023 and is projected to reach USD 34.53 billion by 2033, growing at a CAGR of 18.93% from 2024 to 2033.

Key Takeaways:

Request PDF Sample Copy of Report: (Including Full TOC, List of Tables & Figures, Chart)@ https://www.novaoneadvisor.com/report/sample/8439

owing to risingclinical trialsfor advanced therapy medicinal products and the increasing awareness among researchers about the benefits of advanced therapies, driving the advanced therapy medicinal products (ATMP) CDMO market growth. Tissue engineering has greatly benefited in recent years from technological development. The damaged tissues and organ function are replaced or restored using this technique. Similarly, gene and cell therapy are attracting a lot of patients for the treatment of rare diseases, whose incidence is rising globally.

With rising demand for robust disease treatment therapies, key players have focused their efforts to ramp up research and development for effective gene therapies that target the cause of disorder at a genomic level. According to ASGCT, the number of cell and gene therapies in the U.S. pipeline programs (phase I-III trials) increased from 483 in 2021 to 529 in 2022. Furthermore, the FDA delivers constant support for innovations in the gene therapy field via a number of policies with regard to product manufacturing. In January 2020, the agency released six final guidelines on the manufacturing and clinical development of safe & efficient gene therapy products.

Moreover, awareness about ATMP treatment options is being driven by initiatives aimed at informing the public about the benefits of these products, which, in turn, is leading to increased adoption of advanced therapies and fueling market growth for CDMOs. For instance, Alliance for Regenerative Medicine Foundation for Cell and Gene Medicine prioritizes activities for increasing public awareness through educational programs, underlining the clinical & societal benefits of regenerative medicine.

Increasing clinical trial activity along with new product launches generates growth opportunities for the market. As of 2022, there are 1451 ATMPs in preclinical stages and 535 are being studied in Phase 1 to 3 studies. Since August 2020, EMA has approved six of these additional ATMPs, and five more will be approved by 2023. In the UK, there were approximately 168 advanced therapy medicinal product trials underway in 2021, up from the 154 studies reported the year before, which is a 9% increase. 2021 saw a 32% increase in phase 1 trials, indicating a significant shift from experimental medicines to first-in-human studies.

On the other hand, key players are undertaking various strategic initiatives to introduce novel products, which is expected to propel market growth. For instance, in March 2021, CureVac N.V. signed a partnership agreement with Celonic Group, engaged in the manufacture of CVnCoV, CureVacs mRNA-based COVID-19 vaccine candidate. CureVac's COVID-19 vaccine candidate is manufactured at Celonic's commercial manufacturing unit for ATMPs and biologics in Heidelberg, Germany. Under the terms of the commercial supply agreement, the Celonic facility could produce over 100 million doses of CVnCoV.

Immediate Delivery Available, Get Full Access@ https://www.novaoneadvisor.com/report/checkout/8439

Advanced Therapy Medicinal Products CDMO Market Trends

Segments Insights:

Product Insights

The gene therapy segment held the largest share of over 49.11% in 2023. Increase in financial support and rise in number of clinical trials for gene therapies are driving demand for gene therapy segment. In 2020, in the first three quarters, gene therapies attracted financing of over USD 12 billion globally, with around 370 clinical trials underway. Additionally, in mid-2022, approximately 2,000 gene therapies were in development, targeting several therapeutic areas, such as neurological, cancer, cardiovascular, blood, and infectious diseases.

The cell therapy segment is expected to show lucrative growth over the forecast period. The field of cellular therapeutics is constantly advancing with inclusion of new cell types, which, in turn, provides ample opportunities for companies to enhance their market positions. Furthermore, the market is attracting new entrants due to high unmet demand for cell therapy manufacturing, the recent approval of advanced therapies, and proven effectiveness of these products.

Indication Insights

The oncology segment accounted for the largest revenue share in 2023. The segments dominance is attributed to disease burden, strategic initiatives undertaken by key players, and availability of advanced therapies used for treating various cancer indications. In January 2021, around 18,000 to 19,000 patients and 124,000 patients were estimated to be potential patients for treating cancer using cell & gene therapy products Kymriah (Novartis AG) and Yescarta (Gilead Sciences, Inc.), respectively. Furthermore, a publication on PubMed reports that as of the conclusion of the first quarter of 2023, there have been over 100 distinct gene, cell, and RNA therapies approved globally, along with an additional 3,700-plus in various stages of clinical and preclinical development.

The cardiology segment is estimated to register the fastest CAGR over the forecast period. This is attributed to the increasing prevalence of cardiovascular diseases and research collaboration for development of advanced therapies. For instance, in October 2023, Cleveland Clinic administered a novel gene therapy to the first patient globally as part of a clinical trial, aiming to deliver a functional gene to combat the primary cause of hypertrophic cardiomyopathy (HCM). Similarly, in February 2021, Trizell GmbH entered into partnership with Catalent, Inc. for development of phase 1 cell therapy to treat micro- and macroangiopathy. Trizell's medication is an Advanced Therapy Medicinal Product (ATMP) that employs regulatory macrophagesa platform technology developed in Germany.

Phase Insights

The phase I segment dominated the market in 2023 due to growing R&D activities and increasing number of human trials for advanced therapies. Phase 1 helps ensure the safety levels of a drug at different doses and dosage forms administered to a small number of patients. This phase is mainly conducted to determine the highest dose a patient can take without any adverse effects. Around 70% of drugs in phase 1 move to the next phase.

The phase II segment has been anticipated to show lucrative growth over the forecast period. Phase II clinical studies comprise the largest number of developing ATMPs, due to the high clearance rate of phase I clinical studies. According to data published by Alliance for Regenerative Medicine, as of June 2022, more than 2,093 clinical trials are ongoing globally, out of which 1,117 are under phase II clinical trials accounting for 53%. Thus, the increase in number of products in phase II is driving the segment.

Regional Insights

North America dominated the overall market share of 49.11% in 2023. This can be attributed to increasing outsourcing activities and rising awareness about advanced therapy. North America has consistently been a leader in R&D for advanced treatments, and it is anticipated that it will keep this position during the forecast period. Recent approvals of products such as Kymriah and Yescarta have propelled investments in the regional market. Moreover, in March 2021, the U.S. FDA approved Abecma, the first approval of CAR-T cells to fight against cancer. Similarly, in December 2023, Casgevy and Lyfgenia, the initial cell-based gene therapies for sickle cell disease (SCD) in patients aged 12 and above, received approval from the U.S. Food and Drug Administration, marking a significant milestone.

The U.S. accounted for the largest share of the global market in the North America region in 2023. The U.S. maintains dominance in this sector due to the presence of a robust and highly advanced biopharmaceutical industry with a considerable focus on research and development. Additionally, the continuous presence of numerous pharmaceutical and biotechnology companies, along with academic and research institutions, generates a sustained demand for rigorous safety testing, further reinforcing the country's leadership in the field.

The Asia Pacific region is expected to grow at the fastest CAGR over the forecast period due to the increasing demand for novel ATMPs and rising R&D activities to develop novel therapies. Moreover, the market growth is driven by continuously expanding CDMO Cell Therapy in the country, a number of domestic players have collaborated with biotech companies from other countries involved in mesenchymal stem cell research and therapy development. In addition, in September 2022 Takara Bio, Inc. launched CDMO Cell Therapy for gene therapy products using siTCR technology for its genetically modified T-cell therapy products.

China accounted for the largest share of the global market in the Asia Pacific region in 2023 due to its strategic focus on advancing research and development capabilities, particularly in the pharmaceutical and biotechnology sectors. Additionally, with a rapidly growing biopharmaceutical industry and supportive government initiatives, China has become a key market for advanced therapy medicinal products (CDMO) market.

Recent Developments

Key Companies & Market Share Insights

Some of the key players operating in the market include AGC Biologics,WuXi Advanced Therapies and Celonic

Minaris Regenerative Medicine and BlueReg are some of the emerging market players in the global market.

Key Advanced Therapy Medicinal Products CDMO Companies:

Segments Covered in the Report

This report forecasts revenue growth at country levels and provides an analysis of the latest industry trends in each of the sub-segments from 2021 to 2033. For this study, Nova one advisor, Inc. has segmented the Advanced Therapy Medicinal Products CDMO market.

By Product

By Phase

By Indication

By Region

Immediate Delivery Available | Buy This Premium Research: https://www.novaoneadvisor.com/report/checkout/8439

You can place an order or ask any questions, please feel free to sales@novaoneadvisor.com| USA - +1 9197 992 333

About US

Nova One Advisor is a worldwide market research and consulting organization. We give unmatched nature of offering to our customers present all around the globe across industry verticals. Nova One Advisor has expertise in giving deep-dive market insight along with market intelligence to our customers spread crosswise over various undertakings.

Read more from the original source:
Advanced Therapy Medicinal Products CDMO Industry is Rising Rapidly - BioSpace

Sound of Success, Gene Therapy Breakthrough Grants Hearing to Deaf Children in China-Harvard Study – Hoodline

Sound of Success, Gene Therapy Breakthrough Grants Hearing to Deaf Children in China-Harvard Study  Hoodline

View post:
Sound of Success, Gene Therapy Breakthrough Grants Hearing to Deaf Children in China-Harvard Study - Hoodline

Archives