Recipients of hematopoietic stem cell transplantation (HCT) suffe

Recipients of hematopoietic stem cell transplantation (HCT) suffer from a prolonged post-transplant immune deficiency that results in significant morbidity and mortality [8]. Reconstitution of the T cell population involves both thymus-dependent de novo T cell generation as well as extrathymic expansion of mature, donor-derived T cells and studies in mice indicate that IL-7 may be critically involved in both of these processes [9]. Based on the known functions of IL-7 and TSLP, we hypothesized that polymorphisms in exons of the IL-7Rα gene might influence the process of immune reconstitution after selleck chemicals HCT impacting the risk of infections, acute and chronic graft versus host disease (GvHD) and treatment-related mortality

(TRM). In a previously published study of a Danish HCT cohort, we found an association between donor rs1494555G and rs1494558T and increased TRM after HLA-matched unrelated donor (MUD) HCT [10]. The aim of this study was to validate these findings in an independent, larger and more homogeneous cohort of adults receiving MUD HCT for haematological malignancies. In addition, we evaluated the significance of rs6897932 genotypes in relation to HCT because this SNP has previously been associated with autoimmune disease and allergy [11, 12]. Established in 2004, the Center for

International Blood and Marrow Transplant Research (CIBMTR) is a research affiliation of the International Bone Marrow Transplant Registry (IBMTR), Autologous Blood and Marrow Transplant Registry (ABMTR) and the National Marrow Donor Program (NMDP) and is comprised of a voluntary working group of more than 450 transplantation DAPT mouse centres worldwide that contribute detailed data on consecutive allogeneic and autologous HCT to a Statistical Centre at the Medical College of Wisconsin

in Milwaukee (WI, USA) and the NMDP Coordinating Center many in Minneapolis (MN, USA). Participating centres are required to report all transplants consecutively; compliance is monitored by on-site audits. Patients are followed longitudinally, with yearly follow-up. Computerized checks for discrepancies, physicians’ review of submitted data and on-site audits of participating centres ensure data quality. Observational studies conducted by the CIBMTR are performed in compliance with the privacy rule (HIPAA) as a Public Health Authority and in compliance with all applicable federal regulations pertaining to the protection of human research participants as determined by continuous review of the Institutional Review Boards (IRB) of the NMDP and the Medical College of Wisconsin. The study population consisted of 590 donor/recipients pairs receiving a bone marrow (BM) or growth factor–mobilized peripheral blood stem cell (PBSC) transplant following a myeloablative conditioning regimen between 1988 and 2004 facilitated through the National Marrow Donor Program (NMDP). All donors and recipients were Caucasian and over 18 years old.

One-sided tests were used for comparison of small sample sizes (n

One-sided tests were used for comparison of small sample sizes (n < 5). A P-value of < 0·05 was considered significant in call cases. Elevated Treg numbers have been observed in response to H. pylori infection, both at the site of infection and circulating in the periphery [20, 21]. To determine whether the elevated number of Tregs Metformin nmr was due to active proliferation at the site of infection, we stained gastric biopsy specimens from patients with and without confirmed H. pylori infection for FoxP3 and the proliferation marker Ki67 (four sections

from each patient and four patients). As expected from previous publications, H. pylori-positive biopsy specimens had greater numbers of FoxP3+ cells than H. pylori-negative specimens (Fig. 1a). In the presence of H. pylori, a greater percentage of Tregs stained positively for Ki67 (10·2 ± 1·5% versus 2·4 ± 2·0% of FoxP3+ cells, P < 0·05; Fig. 1a,b), suggesting that Tregs proliferate in vivo in the presence of H. pylori. DCs play a critical role in presenting pathogens to the adaptive immune response. Murine selleck chemical models have indicated that pathogen-stimulated DCs can alter Treg function [22, 26] and their presence in the gastric mucosa indicates that they have the opportunity to influence Treg function [13]. To determine whether H. pylori-stimulated DCs (HpDCs) can influence Treg proliferation

and can, at least in part, explain the expansion of Tregs seen at biopsy sites of H. pylori-infected individuals [10, 13], DCs were generated from peripheral blood monocytes using GM-CSF and IL-4, and incubated with H. pylori [106−4 cfu/ml corresponding to multiplicity of infection (MOI) of 0·75] for 8 h before being washed and placed in co-culture with allogeneic

Tregs for 5 days (Fig. 2), as described previously by us [10]. Allogeneic Tregs were used, as published previously [10], to ensure that the frequency of responding Tregs was not dependent on previous H. pylori exposure and relied purely on the high frequency of alloreactive Tregs [30]. HpDC-induced Treg proliferation was assessed by [3H]-thymidine incorporation; an example is shown in Fig. 2a. This was confirmed through cumulative Venetoclax experiments with HpDCs (106 cfu/ml), in which the differences between Treg proliferation in the presence and absence of H. pylori were found to be statistically significant (Fig. 2b). Tregs were enriched using magnetic beads and, although the purity reached 90%, to ensure further that proliferation measured was not due to non-Treg (e.g. CD4+CD25int T cells) ‘contamination’ of Treg preparations, Tregs were purified to >98% purity by FACS sorting (to ensure that only the CD25hi cells were selected) and cultured with DCs as before. HpDCs expanded allogeneic CD25hi cells, confirming that the proliferation observed was not due to impurities (Fig. 2c). We also ruled out the possibility that H.

Thus, TCRβ diversity is important for optimal TCRαβ pairing and f

Thus, TCRβ diversity is important for optimal TCRαβ pairing and function when TCRα is limiting. Immune T cells play a key role in limiting viral, bacterial, and parasitic infections. Both the CD8+ and the CD4+ cells use specific TCR to recognize epitopes composed of peptide (p) bound to MHC glycoproteins expressed on the surface of infected cells. Following TCR-mediated activation, T cells proliferate, and produce anti-viral cytokines (e.g. IFN-γ and TNF) and cytotoxic effector molecules that function to destroy the pMHC-marked cells. Epitope-specific TCR are selected from pools of naïve precursors that consists of ∼107 (in mice) and ∼108 (in humans) distinct

TCRαβ heterodimers 1, 2 assembled from variable (Vα and Vβ)

and constant (Cα and Cβ) regions. As expected, immune T cells are often characterized by reproducible pMHC-specific biases in TCR Vβ usage 3 and, less frequently, by a limited spectrum of TCR Vα selection 4, 5. The extent PD0325901 order of TCR diversity in an immune repertoire has been related to CTL-mediated control and pathogen escape in CD8+ T-cell Selleckchem CHIR 99021 responses to viruses 6, 7. Most of the diversity in TCR/pMHCI interactions rests in the hypervariable complementarity-determining regions (CDR1, CDR2, and CDR3) involved in TCR-pMHCI binding 8. CDR3β provides the predominant contact in at least some of the antigenic peptides bound inside the groove of the MHC molecule 9, 10. However, the CDR1α, CDR2α, and CDR3α loops also contribute greatly to TCR repertoire diversity and mediate important interactions with antigenic peptides and/or MHC determinants 5, 11, 12. The CDR3β and CDR3α regions reflect the clonal characteristics of immune TCR repertoires. In general, TCR repertoires can be either broad, consisting of numerous clonotypes of different CDR3 aa sequences, CDR3 length, and J regions, or restricted

to a few clonotypes that show similar Jβ and CDR3 characteristics. GNE-0877 TCR repertoires can be also defined as “public” (same clonotypes found in all individuals) or completely “private” (unique to the individual) 3. The exact mechanisms underlying generation of public and private TCR repertoires are far from clear. Influenza virus infection of C57BL/6 (B6, H2b) mice elicits immunodominant CD8+ T-cell responses to peptides from the viral influenza nucleoprotein (NP) and influenza acid polymerase (PA) complexed with the H2Db (DbNP366 and DbPA224), and subdominant CD8+ sets, including those toward the basic polymerase (PB) peptide presented by H2Kb (KbPB1703). Analysis of TCR-CDR3β sequence variability and clone prevalence showed predominantly private and diverse TCRβ sequences for DbPACD8+ T cells 13, but a limited, and substantially public, TCRβ repertoire for the DbNPCD8+ set 14, 15. Thus, influenza infection of B6 mice provides a readily accessible experimental system for dissecting the nexus between TCR repertoire diversity and antiviral efficacy for immune CD8+ T cells.

The ‘typical’ presentation describes the majority of patients wit

The ‘typical’ presentation describes the majority of patients with AD who suffer a broad spectrum of clinical symptoms characterized by early episodic memory loss followed by a combination of attention-executive, language and visuospatial impairments. The ‘focal’ presentations describes those cases where one aspect of brain function (cognition) declines

disproportionately to the others. The focal presentations could be subtyped further into a ‘memory (amnestic)’ variant (n = 6), ‘frontal’ variant (n = 3), ‘visual’ variant (n = 3) and a ‘language’ variant (n = 5). Because of the small number of subtypes available, however, it was only possible to assess whether the distribution of ‘typical’ or ‘focal’ presentations differed Metformin in vitro among the pathological phenotypes. APOE genotyping from DNA extracted from frozen brain tissue (cerebellum or frontal cortex) by phenol chloroform had been previously performed on 127 cases, according to method of Wenham et al. [13]. Sections of frontal (Brodmann areas 8/9), temporal cortex (Brodmann areas 21/22 to include hippocampus) and occipital cortex (Brodmann areas 17/18) were cut at 6 μm thickness from formalin fixed, paraffin embedded blocks and mounted on to glass slides. Sections were first hydrated through successive baths of xylene, alcohols

of decreasing concentration find more and distilled water. The rehydrated sections were transferred to a ceramic holder and subject to chemical antigen retrieval with 90% formic acid at room temperature for 20 min. Sections were incubated for 30 min at room temperature in 0.3% peroxide in methanol to quench endogenous peroxidise activity, and then for a further 30 min at room temperature in Vectastain Elite PK-6100 horse serum as blocking buffer. Sections were then incubated for 1 h at room temperature in mouse monoclonal antibody directed against Amyloid-β17–24 (4G8) (Signet Labs Inc., Dedham, MA, USA), which recognizes D-malate dehydrogenase all molecular forms of

Aβ, at a concentration of 1:3000. The sections were incubated for 30 min in a biotinylated secondary antibody followed by 30 min in avidin–biotin complex (ABC) reagent (both Vectastain Elite PK-6100 Mouse IgG), both at room temperature. Sites of immunoreactions were visualized by incubating in DAB (3,3′-diaminobenzidine tetrahydrochloride) for 5 min, followed by light counterstaining with haematoxylin (Vector H-3401). Sections were dehydrated and mounted for analysis under the microscope. The histological sections were examined under a Leica DMR light microscope. All assessments were made by a single observer (NA). The three topographical regions (frontal, temporal, occipital) for each case were scored consecutively (see below). Sections were scored twice to increase objectivity, and any discrepancies reconciled by consultation with a second observer (D.M.A.M.). The system used to grade CAA was similar to that originated by Olichney et al.

Most assays today employ PR3 isolated from

human neutroph

Most assays today employ PR3 isolated from

human neutrophils [40] by a method that preserves the conformation of the molecule, and attachment of PR3 molecules is accomplished either directly by coating onto some plastic surface (microwells, beads or other particles) or indirectly through attachment via bound specific mouse monoclonal antibody or a linker molecule that does not interfere with important epitopes for human PR3-ANCA reactivity [41]. Less common is the use of recombinant PR3 as antigen. There are data to suggest that ELISAs based on indirect binding of PR3 by a capture technique PARP inhibitor is superior to direct ELISAs in predicting flares of vasculitis [42], but there is no general agreement about this. Such monitoring would most probably have to involve weekly or biweekly testing to be able to catch an ANCA rise and thus predict imminent flares. A P-ANCA staining pattern on neutrophils (Fig. 2) and monocytes is found commonly in patients with different chronic inflammatory diseases, e.g. rheumatoid arthritis, ulcerative colitis and chronic hepatitis, and verification that such antibodies are directed specifically to MPO is mandatory to be useful for diagnosing vasculitis [35]. Even then, it is important to emphasize that P-ANCA directed against MPO is not a specific marker for any of the small vessel

vasculitides, as anti-MPO positivity occurs in many non-vasculitic disorders. The P-ANCA staining pattern can thus be caused by antibodies to several selleck chemicals llc hydrophilic autoantigens in neutrophils that dislocate from their original site of placement onto neighbouring structures, e.g. the nucleus and its adjacent structures upon fixation Ribonucleotide reductase of the cells in ethanol or acetone. A P-ANCA staining pattern can be produced with autoantibodies to MPO, leucocyte elastase, cathepsin G, lactoferrin, azurocidin and lysozyme. If a P-ANCA is not caused by MPO-ANCA, the other specificities may be looked for by separate assays [43], but in practice this is not conducted unless

there is a firm suspicion of a drug-induced condition, e.g. lupus-like syndrome or drug-induced vasculitis, where ANCA directed to one or more of these antigens are common [44]. Pathogenicity of ANCA.  Although ANCA do not fulfil traditional immunological criteria for pathogenicity of autoantibodies, there is substantial evidence attesting to the biological activity of ANCA in terms of stimulation of the neutrophil respiratory burst, induction of cytokine release and increased adhesion to cultured endothelium [45]. However, the occurrence of ANCA in a variety of non-vasculitic disorders suggests that ANCA are heterogeneous in their biological activity and, consequently, their pathogenicity. Animal models offer support for a direct pathogenic role for ANCA IgG in human glomerulonephritis and vasculitis.

Statistical analysis   The statistical significance of differenti

Statistical analysis.  The statistical significance of differential findings between experimental groups was determined by Student’s test. Data were considered statistically significant

at P < 0.05. The recombinant pcDNA3-Ag85A plasmid was confirmed by BamHI and XbaI digestion. The UbGR-Ag85A fusion DNA vaccine was confirmed, respectively, by HindIII and XbaI, BamHI and XbaI, Hind III and XbaI digestion. Finally, the sequences of the two DNA vaccines were shown to be correct by sequencing. After a large scale of preparation, the plasmids were suspended selleck chemicals in endotoxin-free PBS. DNA was quantified by spectrophotometry at 260 nm, and the final concentration of the solution was adjusted to 1 μg/μl of DNA in PBS. The purpose of transfection experiment was to obtain the specific target cells for cytotoxicity assay. After selection by G418 (800 μg/ml), fifteen clones were obtained, and five clones of transfected cells were randomly chosen and screened for Ag85A mRNA by reverse transcription-PCR. After electrophoresis, a specific single band about 1.0 kb in length Akt inhibitor was

observed in clones I, II, III, IV and V. And then, the expression of Ag85A was further examined in clone I by immunocytochemistry. The immunostaining was restricted to the cytoplasm of the cells transfected with pcDNA3-Ag85A plasmid. However, no staining was detected in P815 cells. No staining Adenosine signals were detected with the sera from healthy control people, which indicated that the staining is specific. Those results demonstrated that Ag85A antigen could be expressed stably in P815 cells and that the clone I could be used as target cells in the cytotoxicity assay. To determine the level of Ag85A-specific IgG elicited by different vaccines, mice of different groups were immunized three times at 3-week intervals. Three weeks after the last immunization, the sera from mice were collected by retro-orbital bleeding, and

antigen-specific antibodies were detected by ELISA. As shown in Fig. 1, compared with the pcDNA3 vector group or pcDNA3-ub group, the Ag85A DNA vaccine elicited a significantly higher level of IgG (P < 0.01). However, the IgG level in the UbGR-Ag85A fusion DNA vaccine group was lower than that in Ag85A DNA vaccine group (P < 0.01). The IgG subclasses give an indication of the Th1 versus Th2 nature of the immune response. We also detected the relative ratio of IgG2a/IgG1. As shown in Fig. 2, although the IgG level decreased in the ub fusion DNA vaccine group, the relative ratio of IgG2a/IgG1 increased significantly in the fusion DNA vaccine (P < 0.01), compared with the Ag85A DNA vaccine group. T helper cells play an important role in eliciting both humoral and cellular immune responses via expansion of antigen-stimulated B cells and expansion of CD8+ T cells.

Amongst the upregulated genes, the p62 (also known as sequestosom

Amongst the upregulated genes, the p62 (also known as sequestosome 1) (SQSTM1) is an adaptor protein that has a role in inflammation, neurogenesis, osteoclastogeneis, adipogenesis and T-cell differentiation [21]. Our data indicated that p62 is induced by TLR-2 and NOD-1 activation at both mRNA and protein levels. Elucidating the pathways that control HCS assay p62 levels in MSC will add another layer of detail to our understanding of the cell differentiation cascades in which p62

is involved. In addition to p62, VEGF and CXCL-10 were upregulated in response to NOD-1 and TLR-2 signalling. Human MSC released VEGF in response to TLR-2 and NOD-1 ligands as a potentially beneficial paracrine response. It will be interesting to investigate which mechanisms are involved in VEGF upregulation and secretion in MSC. Notably, previous studies have suggested a direct contribution of MSC to the blood vessel formation, as differentiation of MSC

into endothelial cells has been demonstrated [22, 23]. In contrast to NOD-1, TLR-2 signalling BGB324 cell line also upregulated the expression of several important genes such as interleukin-1 receptor-associated kinase 2 (IRAK-2), involved in TLR signalling, NOTCH-1 and Gal-3 involved in innate and adaptive immunity. Notably, Notch pathway is highly conserved in evolution and is generally involved in cell fate decisions during cell differentiation [24]. A recent study showed that the inhibition of Notch signalling in MSC can hinder their suppressive activity on T-cell proliferation [13]. In addition to binding to glycan structures that are expressed by host cells, galectins can also recognize β-galactoside carbohydrates that are common structures on many pathogens [25], and therefore they are considered as a soluble pathogen recognition receptor. Within

the immune system, galectins are expressed Cepharanthine by virtually all immune cells, either constitutively or in an inducible fashion [17]. Also, they can be expressed by a spectrum of normal and tumour cells. As found in this study, Gal-3 is constitutively expressed by MSC and upregulated in response to TLR-2 ligation. Of note, high levels of Gal-3 protein are found in MSC culture supernatants; thus, it may participate in extra cellular matrix (ECM)-cell interactions and modulation of surrounding immune cells. Results from knockdown experiments showed that the immunosuppressive effects of MSC on T cells was lower than that from cells expressing Gal-3, suggesting a possible involvement of Gal-3 in MSC immunosuppressive function. This observation would fit with the demonstrated inhibitory effect of Gal-3 on T-cell proliferation [19, 20]. Also, a more recent study showed that tumour-associated Gal-3 contributes to tumour immune escapes by inhibiting the function of tumour-reactive T cells [26]. Some studies demonstrated that the MSC immunoregulatory properties are at least in part mediated by the production of cytokines, such TGF-β and hepatocyte growth factors [27].

The membrane was then washed three times in TBST, incubated with

The membrane was then washed three times in TBST, incubated with an anti-rabbit horseradish peroxidase-conjugated secondary antibody (1 : 3000, Bio-Rad Laboratories) for 1 h at room temperature and washed again. Proteins were visualized using

the SuperSignal West Pico Chemiluminescent Substrate (Pierce). To ensure equal amounts of FAK in all samples, the membrane was stripped and reprobed with rabbit anti-FAK antibody C-309 (1 : 200 in blocking buffer, Santa Cruz Biotechnologies). Digital images of the membrane were analyzed using imagej software (NIH, Bethesda, MD, http://rsb.info.nih.gov/ij/). Data were analyzed by one-way anova with the Newman–Keuls multiple comparison post test using the graphpad prism version 5.02 (GraphPad Software,

San Diego, CA). BMS-777607 in vitro Differences with P-values <0.05 were considered statistically Paclitaxel cost significant. The expression of EpoR in nonerythropoietic tissue is debated (Ghezzi et al., 2010; Sinclair et al., 2010; Swift et al., 2010; Xiong et al., 2010). A prediction for our hypothesis was, however, that EpoR, as part of the heteromer with CD131, is expressed in the bladder epithelium. We therefore tested the bladder epithelial cell lines and primary bladder epithelial cells used in our cell infection model for EpoR expression. We could detect low constitutive levels of EpoR-specific mRNA in all three bladder cell types investigated in this study (Fig. 1) as well as in the monocytic cell line THP-1. Discrepancies among the findings these reported by others might result from the different sensitivities of methods or interpretation criteria (Ghezzi et al., 2010). Contact between E. coli and bladder epithelial cells induces a general inflammatory response. In other nonerythropoietic tissues, TNF-α-dependent upregulation of EpoR has been described to mediate the tissue-protective action of Epo (Brines & Cerami, 2008). To investigate whether this also applies for bladder epithelial cells, we exposed cells to bacterial stimuli, E. coli NU14, and determined the mRNA expression of EpoR at different time points after stimulation. The expression of EpoR was induced in a bimodal

manner, with a first peak at three (5637 cells) or 6 h (primary cells) and a second upregulation after 24 h of stimulation (Fig. 2a). This first peak was very low in T24 cells stimulated with bacteria alone. When, however, these cells were costimulated with ARA290, EpoR expression was upregulated 3 h after costimulation (P<0.05; Fig. 2b). Enhanced and earlier EpoR upregulation in the presence of ARA290 was also observed for 5637 and primary bladder epithelial cells, although the effect was less pronounced (data not shown). In the monocytic cell line THP-1, a similar pattern was observed, but expression peaked earlier, after 1 and 12 h of stimulation, respectively (Fig. 2a). Additional stimulation with ARA290 showed no obvious additive effect.

25×104 in a final volume of 50 μL A total of 721 221 target cell

25×104 in a final volume of 50 μL. A total of 721.221 target cells were added (1×104 in 50 μL) to each well (quadruplicate wells were assayed per point) and the plate was centrifuged at 500 rpm for 1 min and incubated for 4 h at 37°C. At the end of this incubation period, 50 μL of assay buffer was added to each well. The substrate (50 μL/well) was added and the samples incubated in dark for 15 min. The plates were read using Synergy4 microplate reader (BioTeK® Instruments). Maximum cell lysis was determined by treating 1×104 target cells with 0.1% digitonin in assay buffer for 3 min at RT. Freshly harvested YTS, control vector-transduced YTS, and IQGAP1 shRNA-transduced YTS cells washed and resuspended

in 0.5% BSA in PBS (PBS-BSA). The cells were fixed for 10 min Erlotinib cost at room temperature in PBS containing 2% paraformaldehyde, washed three times in Venetoclax cost PBS-BSA, and permeabilized

with 0.1% Tween-20 in PBS-BSA for 5 min. The cells were washed three times in PBS-BSA and incubated with primary Ab against IQGAP1 or Alexa fluor phalloidin 488 for 45 min. The cells were washed and incubated with secondary goat anti-rabbit Alexa fluor 488 for 45 min. Cells were washed and staining was assessed (10 000 cells/sample) using a BD FACS Array system. First, the live cells were gated to exclude debris, and then the number of cells positive for Alexa fluor 488 was assessed within this population. The assay was performed according to the method described in 26. YTS cells were prelabeled with 1.5 μM Cell Tracker™ Green CMFDA (Invitrogen cat no. C2925) and target cells were labeled with 5 μM Cell Tracker™ Orange (Invitrogen cat no. C34551). The for cells were combined

at an effector to target ratio of 2:1 and incubated for the indicated times. The samples were gently vortexed for 3 s at maximum vortex speed and immediately fixed with 2% PFA. Samples were run in triplicates and 30 000 events were counted for every replicate. The frequency of double-positive events was determined within the Cell Tracker™ Green-positive population using Summit V5.2.0.7477 software. The following gating strategy was used: First, the live cells were gated to exclude debris. Compensation adjustments were made on this population using single-positive cells stained for either of the two dyes. Gates were set to differentiate between the double positives, represented in G2, from the single positives and double negatives in the experimental cells. This research was supported by a grant from the Canada Institutes for Health Research (J. A. W.) and the Health Sciences Research Department (N. K.). The authors thank Qiujang Du for preparation of the shRNA-mir constructs and Monroe Chan for flow cytometry. Conflict of interest: The authors declare no financial or commercial conflict of interest. Detailed facts of importance to specialist readers are published as ”Supporting Information”.

Results: The mean daily salt excretion was 9 9 ± 2 6 g BP and eG

Results: The mean daily salt excretion was 9.9 ± 2.6 g. BP and eGFR were not different among for groups. However, incidence of overt proteinuria was significantly higher in the first quartile (Q1: 23%, Q2: 9%, Q3: 2%, Q4: 2%, p < 0.001). Conclusion: Low daily salt excretion was correlated with proteinuria in non-diabetic patients. Although the cause and effect relationship between salt intake and proteinuria could not be determined in BVD-523 this study, low daily

salt excretion could be a marker for proteinuria in non-diabetic outpatients. AHMAD ISABEL1, YANG YATING ADONSIA1,2, LAU TITUS1,2, SETHI SUNIL1,2, TEO BOON WEE1,2, LIN TINGXUAN1,2, TOH QI CHUN1,2, CHONG YUE TING1,2, LI JIALING1,2 1National University Health System, Singapore; 2National University of Singapore, Singapore Introduction: Clinical practice guidelines recommend using 2 or more anti-hypertensive agents to control blood pressure (BP) to targets in chronic kidney disease (CKD) patients. The impact of the number of medications on the components of BP (systolic, SBP; diastolic, DBP) in Asian CKD patients is unclear. We assessed the effects of the number of anti-hypertensive agents on BP components

RG-7204 in a multi-ethnic Asian population of stable CKD patients. Methods: We prospectively recruited 613 patients (male 55.1%, Chinese 74.7%, Indian 6.4%, Malay 11.4%, Others 7.5%; 35.7% diabetes) with mean age 57.8 ± 14.5 years. BP was measured according to guidelines using calibrated automatic manometers. Glomerular filtration rate (GFR) was estimated using the CKD-EPI equation. ANOVA was used to compare means of BP components with number of anti-hypertensive medications, and Tukey-Kramer HSD for pairwise comparisons. Linear regression was used to assess associations of BP with continuous variables. Non-normally distributed data was natural log-transformed for analyses. Results: The mean SBP was 139 ± 21 mmHg, DBP

74 ± 11 mmHg, serum creatinine 166 ± 115 μmol/L, and GFR 53 ± 32 mL/min/1.73 m2. SBP increased with lower GFR (p < 0.001), whereas DBP decreased with lower GFR (p = 0.0052). Mean SBP increased with increasing number of antihypertensive agents used (p < 0.001), whereas mean DBP decreased with ≥3 antihypertensive Rapamycin supplier agents used (p = 0.0020, Table 1). Conclusions: Clinical practice guidelines recommend different component BP targets for CKD patients. Increasing number of antihypertensive agents use results in a divergence in the achievement of targets. Further research into improved methods of monitoring and treatment is required to better achieve targets. SHIMIZU HIDEKI, KANAME SHINYA, KAWASHIMA SOKO, IKEGAYA NORIKO, HAYAKAWA SATOSHI, FUKUOKA KAZUHITO, KARUBE MIHO, KOMAGATA YOSHINORI, ARIMURA YOSHIHIRO, YAMADA AKIRA First Department of Internal Medicine, Kyorin University School of Medicine Introduction and Purpose: We aimed to examine the hypothesis that renoprotective effect of angiotensin II (AngII) receptor blocker telmisartan may be associated with obesity.