首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Leu-2/T8 is a cell surface glycoprotein expressed by most cytotoxic and suppressor T lymphocytes. Its expression on T cells correlates best with recognition of class I major histocompatibility complex antigens, and it has been postulated to be a receptor for these proteins. We have determined the complete primary structure of Leu-2/T8 from the nucleotide sequence of its cDNA. The protein contains a classical signal peptide, two external domains, a hydrophobic transmembrane region, and a cytoplasmic tail. The N-terminal domain of the protein has striking homology to variable regions of immunoglobulins and the T cell receptor. The membrane-proximal domain appears to be a hinge-like region similar to that of immunoglobulin heavy chains. The superfamily of immunologically important surface molecules can now be extended to include Leu-2/T8.  相似文献   

2.
A memory response is established following primary antigen exposure that stays more or less constant. It appears to adopt a set-point in magnitude but upon re-exposure the response is quicker and better and there is an upward shift in memory frequency that varies with individuals based on the exposure pattern to other microbes or its components. Our investigations were designed to test such differences of non-specific stimulation by PAMPs in lowering the threshold of activation. Neonatal mice were pre-exposed to TLR-ligands intermittently and later analyzed for its resilience to challenge with virus during adult-life. Secondly, adult mice with pre-existing memory to virus were exposed to various TLR-ligands and analyzed for their quality of memory response. The TLR-ligands exposed animals were better responders to a new agent exposure compared to the animals kept in sterile surroundings. Moreover, immune memory recall and the viral specific CD8+ T cells response with TLR-ligands were comparable to the recall response with the cognate antigen. The results provide insights into the role of hyper-sanitized environment versus PAMPs mediated signaling in adaptive immunity and long-term immune memory.  相似文献   

3.
Activation of peripheral blood T cells, and the leukemic T cell line Jurkat, as measured by mobilization of intracellular calcium, by an anti-TCR antibody is blocked by mAb (T191) to the leukocyte common Ag (CD45). T191 also blocked down-regulation of the CD3-TCR complex induced by an anti-CD3 mAb. Vanadate, a phosphotyrosine phosphatase inhibitor, partially blocks the effect of T191 and restored mobilization of intracellular calcium. Assays of the immunoprecipitates of T191 and CD45 from both Jurkat-BM1 and peripheral T cells showed that the immune complexes had intrinsic phosphatase activity. A parallel immunoprecipitate using a mAb (4-10) against HLA class I showed no such activity. Further analysis of the T191 immunocomplex revealed activity against phosphotyrosine, p-nitrophenylphosphate, and [32P-poly-glu-tyr, but not against phosphoserine. Phosphatase activity was inhibited by Vanadate, but not by Zn2+ or F-. These results show that CD45 is a phosphotyrosine phosphatase, and strongly suggest that tyrosine phosphorylation/dephosphorylation is critically involved in activation of T cells through the TCR-CD3 complex.  相似文献   

4.
Thy-1- T cells expressing CD4 and the alpha beta-TCR have been identified in murine lymphoid tissues. These cells are particularly prevalent in Peyer's patches (PP), representing 17 +/- 3% of PP CD4 T cells, whereas they are much less prevalent in spleen, lymph nodes, lamina propria, or peritoneum. Phenotypic studies of fresh-isolated PP T cells demonstrate that all PP CD4 T cells (both Thy-1- and Thy-1+) express CD3, alpha beta-TCR, and CD5 (Lyt-1), whereas none coexpress CD8 (Lyt-2). Thy-1- and Thy-1+ CD4 T cell lines generated from PP also coexpress CD3 and alpha beta-TCR, but are heterogeneous in expression of CD5 and again do not coexpress CD8. Further studies revealed that Thy-1- CD4+ T cells were not present in nude mice. Short term stimulation of Thy-1+ CD4+ PP T cells with anti-CD3 resulted in loss of Thy-1 in a substantial fraction of these cells. Functional studies of Thy-1- and Thy-1+ CD4+ PP T cells indicate that fresh-isolated Thy-1- CD4+ cells do not proliferate in response to insoluble anti-CD3 but do proliferate when stimulated with soluble anti-CD3 in the presence of feeder cells. In contrast, Thy-1+ CD4+ cells proliferate well to both stimuli. However, Thy-1- CD4+ PP T cells adapted to in vitro culture exhibit vigorous proliferative responses when stimulated with either form of anti-CD3. Evaluation of lymphokine secretion by fresh-isolated Thy-1- and Thy-1+ CD4+ PP T cells revealed that both make substantial amounts of IL-2; however, Thy-1- T cells made less IL-4 than their Thy-1+ counterparts. Neither population made IL-5 or IFN-gamma. Similarly, Thy-1- and Thy-1+ CD4 T cell lines made similar amounts of IL-2; again Thy-1- T cells made less IL-4; and in this case Thy-1- T cells made IL-5 albeit significantly less than the Thy-1+ cells. Finally, immunohistochemical studies suggested that many of the CD4+ T cells in PP germinal centers were Thy-1-, indicating that Thy-1- and Thy-1+ CD4 T cells differ in their distribution within the PP. These studies thus define a phenotypically and functionally distinct T cell population which is most prevalent in murine Peyer's patches.  相似文献   

5.
Fetal thymic organ cultures (FTOC) were tested as a model system to induce, in a polyclonal fashion, negative and positive thymic selection events. By flow cytometry, thymocytes developed in FTOC differed in several parameters from their in vivo differentiated counterparts. In particular, no clear distinction was possible between CD4+CD8+ immature cells with low TCR expression and mature CD4+ or CD8+ cells with high TCR expression. Thymocyte development in FTOC was manipulated with three different antibody reagents: anti-V beta 8 (F23.1), anti-Lyt-2.2 (19/178) and the quadroma derived bifunctional antibody HPHT-2, carrying one binding site of each. This antibody served also as a monovalent anti-V beta 8 reagent in FTOC from Lyt-2.1 mouse strains. Antibody 19/178 suppressed the development of single positive CD8+ cells, but only at very high concentrations. F23.1 and HPHT-2 suppressed the development of CD4+V beta 8+ and CD8+V beta 8+ thymocytes at relatively low concentrations giving rise to V beta 8 occupancies from about 2% upwards. Suppression was equally pronounced in cells with low and high TCR densities. Moreover, V beta 8 suppression occurred upon divalent and monovalent V beta 8 binding and was not significantly influenced by V beta 8-CD8 cross-linking. This suggests that ligation of the TCR alone is sufficient for clonal deletion. The data do not exclude a role for CD8 as an accessory adhesion molecule but suggest that exogenous cross-linking of CD8 to the TCR is not essential in transmembrane signaling for clonal deletion. At lower antibody concentrations giving rise to V beta 8 occupancies below detection, V beta 8-CD8 cross-linking by HPHT-2, but no divalent and monovalent V beta 8 ligation, induced an increase of CD8+V beta 8+ cells at the expense of CD4+ V beta 8+ cells with no change in the proportion of total V beta 8+ thymocytes. The latter effect was quantitatively of borderline significance but reproducible. These latter results are compatible with the hypothesis that cross-linking of the alpha beta TCR and CD8 on the thymocyte surface provides a maturation signal resulting in loss of CD4 from CD4+ CD8+ double positive immature thymocytes.  相似文献   

6.
The TCR for Ag, on the majority of human T cells, is a disulfide-linked heterodimer composed of TCR-alpha and -beta chains noncovalently associated with the monomorphic CD3 complex composed of the CD3-gamma, -delta, -epsilon, and -zeta chains. The interactions involved in the assembly of the various components of this multimeric protein complex are not fully understood. In this report, a variant of the human leukemic T cell line Jurkat that synthesized all of the known components of the TCR/CD3 complex but fails to express the TCR/CD3 complex at the cell surface is further characterized. This variant, J79, has a mutated TCR-alpha chain that does not affect the assembly of the pentameric form (TCR-alpha beta-CD3-gamma delta epsilon) of the TCR/CD3 complex but inhibits the assembly of the CD3-zeta homodimer with the rest of the complex (TCR-alpha beta-CD3-gamma delta epsilon----TCR-alpha beta-CD3-gamma delta epsilon zeta 2). Transfecting a wild-type TCR-alpha gene into J79 reconstituted expression of a complete functionally competent TCR/CD3 complex at the cell surface. The results indicate that the TCR-alpha chain plays a crucial role in the assembly of the CD3-zeta homodimer with the pentameric form of the TCR/CD3 complex.  相似文献   

7.
The mAb F23.1, specific for V beta 8-related determinants on the TCR, was used to study the requirements for TCR cross-linking and for accessory cells (AC) in the induction of proliferation or IL-2 responsiveness in L3T4+ (CD4+) and Lyt-2+ (CD8+) T cells. T cells were exposed in vitro to soluble native F23.1 antibody, to heteroconjugates composed of the Fab fragments of F23.1 linked to Fab fragments of antibodies specific for Ia determinants on AC, or to F23.1 immobilized on an insoluble matrix. Soluble F23.1 antibody-induced proliferation in naive T cells only in the presence of both AC and exogenous IL-2, and these responses were confined to Lyt-2+ T cells. In contrast, heteroconjugates capable of crosslinking F23.1+ TCR to AC surface Ia determinants were capable of inducing proliferation in both L3T4+ and Lyt-2+ T cells in the absence of added lymphokine. Moreover, binding to and presumably multi-valent crosslinking of the TCR by immobilized F23.1 was sufficient to induce proliferation in both Lyt-2+ and L3T4+ T cells in the absence of AC or exogenous IL-2. Further, it was found that the conditions necessary for T cell growth factor secretion paralleled closely those required for induction of T cell proliferation in the absence of added lymphokine, suggesting that production of endogenous lymphokine might be the limiting process for triggering of T cell proliferation. Taken together, these findings suggest that under optimal conditions of TCR cross-linking, TCR occupancy and cross-linking is sufficient to deliver all of the signals necessary to initiate proliferation in naive populations of both L3T4+ and Lyt-2+ T cells. However, when conditions for TCR signaling are suboptimal, as may be the case for normal Ag-mediated stimulation, a role for second signals delivered by AC or exogenous lymphokines can become critical for T cell activation.  相似文献   

8.
The CD8 coreceptor enhances T cell function by stabilizing the TCR/peptide/MHC complex and/or increasing T cell avidity via interactions with the intracellular kinases Lck and LAT. We previously reported a CD4(+) T cell (TIL 1383I), which recognizes the tumor-associated Ag tyrosinase in the context of HLA-A2. To determine whether CD8 independent tumor cell recognition is a property of the TCR, we used retroviral transduction to express the TIL 1383I TCR in the CD8(-) murine lymphoma, 58 alpha(-)/beta(-). Immunofluorescent staining of TCR-transduced cells with human TCR V beta subfamily-specific and mouse CD3-specific Abs confirmed surface expression of the transferred TCR and coexpression of mouse CD3. Transduced effector cells secreted significant amounts of IL-2 following Ag presentation by tyrosinase peptide-pulsed T2 cells as well as stimulation with HLA-A2(+) melanoma lines compared with T2 cells alone or HLA-A2(-) melanoma cells. Further analysis of TCR-transduced clones demonstrated a correlation between T cell avidity and cell surface expression of the TCR. Therefore, the TIL 1383I TCR has sufficient affinity to mediate recognition of the physiologic levels of Ag expressed by tumor cells in the absence of CD8 expression.  相似文献   

9.
Most mature human T lymphocytes express both the multichain T3 (CD3)/Ti T cell receptor for antigen (TCR), and the biochemically distinct 55-kDa T11 (CD2) glycoprotein. Stimulating the T11 molecule causes profound T cell proliferation and functional activation in vitro, but the relationship of T11-mediated activation to antigenic stimulation of T lymphocytes in vivo remains unknown. We now present evidence that T11 function is directly linked to TCR components in T3/Ti+ T11+ human T cells. First, we found that stimulating peripheral blood T cells with the mitogenic combination of anti-T11(2) cells with the mitogenic combination of anti-T11(2) plus anti-T11(3) monoclonal antibodies caused the phosphorylation of TCR T3 chains. The predominance of T3-gamma-phosphorylation that occurred in anti-T11(2) plus anti-T11(3)-treated T cells is similar to the pattern previously observed in antigen-stimulated T cell clones. Second, T11 function depended upon concurrent cell-surface expression of the TCR. Thus, when peripheral blood T cells were deprived of cell surface T3/Ti by anti-T3 modulation, anti-T11(2) plus anti-T11(3)-induced mitogenesis and transmembrane signal generation in the form of calcium mobilization were inhibited. The mechanism of TCR-T11 interdependence was investigated in a series of TCR-deficient variants of a T cell lymphoblastoid cell line. T3/Ti negative variants expressed cell surface T11, but anti-T11(2) plus anti-T11(3) failed to cause detectable calcium mobilization. The TCR-deficient variants also failed to express T11(3) activation epitopes after incubation with anti-T11(2) antibodies, suggesting that T11(3) expression is an essential and TCR-dependent intermediate in the T11 activation mechanism in these cells. Taken together, our results suggest that T11 function depends upon cell-surface expression of TCR in many T3/Ti+ T11+ T lymphocytes, and T11-mediated activation is intimately interconnected with TCR activation mechanisms. A model in which stimulating signals delivered via T11 may be a part of antigenic activation of T lymphocytes is presented.  相似文献   

10.
This study follows our previous investigation describing the production of four cytokines (IL-2, IL-4, IFN-gamma, and TNF-alpha) by subsets of thymocytes defined by the expression of CD3, 4, 8, and 25. Here we investigate in greater detail subpopulations of CD4-CD8- double negative (DN) thymocytes. First we divided immature CD25-CD4-CD8-CD3- (CD25- triple negative) (TN) thymocytes into CD44+ and CD44- subsets. The CD44+ population includes very immature precursor T cells and produced high titers of IL-2, TNF-alpha, and IFN-gamma upon activation with calcium ionophore and phorbol ester. In contrast, the CD44- subset of CD25- TN thymocytes did not produce any of the cytokines studied under similar activation conditions. This observation indicates that the latter subset, which differentiates spontaneously in vitro into CD4+CD8+, already resembles CD4+CD8+ thymocytes (which do not produce any of the tested cytokines). We also subdivided the more mature CD3+ DN thymocytes into TCR-alpha beta- and TCR-gamma delta-bearing subsets. These cells produced cytokines upon activation with solid phase anti-CD3 mAb. gamma delta TCR+ DN thymocytes produced IL-2, IFN-gamma and TNF-alpha, whereas alpha beta TCR+ DN thymocytes produced IL-4, IFN-gamma, and TNF-alpha but not IL-2. We then studied alpha beta TCR+ DN T cells isolated from the spleen and found a similar cytokine production profile. Furthermore, splenic alpha beta TCR+ DN cells showed a TCR V beta gene expression profile reminiscent of alpha beta TCR+ DN thymocytes (predominant use of V beta 8.2). These observations suggest that at least some alpha beta TCR+ DN splenocytes are derived from alpha beta TCR+ DN thymocytes and also raises the possibility that these cells may play a role in the development of Th2 responses through their production of IL-4.  相似文献   

11.
12.
T cell lines with a novel phenotype (CD3+ TCR-alpha/beta+ CD4- CD8-) were developed from the peripheral blood of a patient with a combined immunodeficiency and tissue injury resembling graft-vs-host disease. One of these IL-2-dependent T cell lines demonstrated non-MHC-restricted cytolytic function against tumor targets, syngeneic and allogeneic fibroblasts, and PHA blasts from allogeneic donors. The other cell line only became cytotoxic in the presence of lectin or anti-CD3 antibody. The two cell lines also differed in their expression of the T-200 gene products CD45RO (gp180) and CD45RA (gp220). Both cell lines produced tumor necrosis factor-alpha and -beta and IFN-gamma activity when activated with mitogens or PMA and IL-1. The in vitro functions of these T-cell lines suggest a potential role for alpha/beta double-negative T lymphocytes in tissue injury resembling graft-vs-host disease.  相似文献   

13.
The outcome of viral infections is dependent on the function of CD8+ T cells which are tightly regulated by costimulatory molecules. The NK cell receptor 2B4 (CD244) is a transmembrane protein belonging to the Ig superfamily which can also be expressed by CD8+ T cells. The aim of this study was to analyze the role of 2B4 as an additional costimulatory receptor regulating CD8+ T cell function and in particular to investigate its implication for exhaustion of hepatitis C virus (HCV)-specific CD8+ T cells during persistent infection. We demonstrate that (i) 2B4 is expressed on virus-specific CD8+ T cells during acute and chronic hepatitis C, (ii) that 2B4 cross-linking can lead to both inhibition and activation of HCV-specific CD8+ T cell function, depending on expression levels of 2B4 and the intracellular adaptor molecule SAP and (iii) that 2B4 stimulation may counteract enhanced proliferation of HCV-specific CD8+ T cells induced by PD1 blockade. We suggest that 2B4 is another important molecule within the network of costimulatory/inhibitory receptors regulating CD8+ T cell function in acute and chronic hepatitis C and that 2B4 expression levels could also be a marker of CD8+ T cell dysfunction. Understanding in more detail how 2B4 exerts its differential effects could have implications for the development of novel immunotherapies of HCV infection aiming to achieve immune control.  相似文献   

14.
D Redelman 《Cytometry》1987,8(2):170-183
The E-rosette receptor (CD2, T11) is a differentiation antigen expressed on immature and mature human T lymphocytes. Activation of T cells from human peripheral blood with phytohemagglutinin (PHA) or with monoclonal antibody to the CD3-Ti complex (anti-Leu-4) caused the expression of CD2 to increase 10- to 20-fold. Dual parameter correlated analyses with antibody to the T cell growth factor (TCGF) receptor (anti-Tac) and anti-CD2 antibody demonstrated that the increase in CD2 expression occurred at the same time and on the same cells that expressed the TCGF receptor after stimulation with PHA. The increased expression of CD2 and the initial expression of Tac were totally inhibited by cycloheximide, but were not affected by sufficient actinomycin-D to block the T cell proliferative response. The expression of CD2 was compared with the expression of CD4 and CD8, i.e., T cell differentiation antigens on cytotoxic/suppressor or helper T cells, respectively. Although virtually all of the small percentage of freshly isolated Tac+ peripheral blood cells belonged to the CD4+, CD8- subset, both CD4+ and CD8+ T cells were equivalently activated by PHA to express Tac. By 20-30 hr after activation, the expression of CD4 or CD8 was initially decreased 10-50%. Subsequently, the expression of CD4 and CD8 returned to the levels on resting T cells but did not increase further. Therefore, the increase in CD2 expression does not reflect a universal property of cell surface antigens on activated T lymphocytes.  相似文献   

15.
An immunohistological study of L3T4(CD4)+ and LYT-2(CD8)+ lymphocytes, Mac-1(CD11b)+ monocytes and granulocytes in experimental murine cryptococcal meningoencephalitis was conducted. To assess the concomitant inflammatory reaction in an extracerebral site, livers were examined in parallel. Mice were infected i.v. withCryptococcus neoformans, group A/D, and organs were examined immunohistologically for CD4-, CD8- and monocyteand granulocyte-specific CD11b-phenotypic leukocytes over a period of 60 days. Intracerebrally, agglomerations of cryptococci formed pseudocysts that were surrounded by CD4+ and CD8+ lymphocytes at the end of the second week post-infection, followed by the invasion of monocytes and granulocytes into the lesions. After the fourth week post-infection, most of the invaded lesions were transformed into glious scars. Meningitis was usually marked and showed a homogenous distribution of CD4-, CD8- and CD11b-phenotypic cells, with a predominance of monocytes and CD4+ lymphocytes. Inflammatory infiltrates in the liver were found already 4 days post-infection. CD4+ lymphocytes and monocytes were distributed homogenously in the infiltrates, with a lower number of CD8+ lymphocytes being located rather in the periphery of the infiltrates. Comparing leukocyte kinetics in brain and liver, an important observation was the delayed immigration of immune cells at the intracerebral cryptococcal lesions as compared with the liver, and the different migration patterns of T-lymphocyte subgroups and macrophages. These results suggest that there are differential leukocyte migration patterns in the liver and brain following disseminated cryptococcosis. The immunological aspects of the observed leukocyte kinetics are discussed.  相似文献   

16.
Recent studies in basic immunology have been directed towards the understanding of the mechanism of T cell activation. T cells can be activated to proliferatevia the classical pathway through the antigen receptor (CD3-Ti) orvia the alternate pathway through the CD2 receptor. Since immunologic unresponsiveness in lepromatous leprosy is considered to be due to the inability of T cells to proliferate upon stimulation, we have been interested in the nature of these receptors and the activation pathways in lymphocytes of leprosy patients. In the present investigation we demonstrate: (i) CD2 receptor (E-receptor) is downregulated in bacterial index positive lepromatous leprosy patients. (ii) The alternate pathway of T cell activation is impaired in lepromatous patients as revealed by the inability of their lymphocytes to proliferate in response to a pair of mitogenic anti-CD2 monoclonals. (iii) The addition of recombinant interleukin 2in vitro restores the ability of lymphocytes from lepromatous patients to proliferate in response to anti-CD2 antibodies. (iv) Interestingly, CD2 modulation and the associated functional impairment could be brought about in peripheral blood lymphocytes from normal subjects by prior treatment withMycobacterium leprae in vitro. This approach would be useful in understanding the molecular events leading to the defective T cell functions in leprosy.  相似文献   

17.
Cellular FLIP long form (c-FLIP(L)) was originally identified as an inhibitor of Fas (CD95/Apo-1). Subsequently, additional functions of c-FLIP(L) were identified through its association with receptor-interacting protein (RIP)1 and TNFR-associated factor 2 to activate NF-kappaB, as well as by its association with and activation of caspase-8. T cells from c-FLIP(L)-transgenic (Tg) mice manifest hyperproliferation upon activation, although it was not clear which of the various functions of c-FLIP(L) was involved. We have further explored the effect of c-FLIP(L) on CD8(+) effector T cell function and its mechanism of action. c-FLIP(L)-Tg CD8(+) T cells have increased proliferation and IL-2 responsiveness to cognate Ags as well as to low-affinity Ag variants, due to increased CD25 expression. They also have a T cytotoxic 2 cytokine phenotype. c-FLIP(L)-Tg CD8(+) T cells manifest greater caspase activity and NF-kappaB activity upon activation. Both augmented proliferation and CD25 expression are blocked by caspase inhibition. c-FLIP(L) itself is a substrate of the caspase activity in effector T cells, being cleaved to a p43(FLIP) form. p43(FLIP) more efficiently recruits RIP1 than full-length c-FLIP(L) to activate NF-kappaB. c-FLIP(L) and RIP1 also coimmunoprecipitate with active caspase-8 in effector CD8(+) T cells. Thus, one mechanism by which c-FLIP(L) influences effector T cell function is through its activation of caspase-8, which in turn cleaves c-FLIP(L) to allow RIP1 recruitment and NF-kappaB activation. This provides a partial explanation of why caspase activity is required to initiate proliferation of resting T cells.  相似文献   

18.
To examine the bases for CD8 T cell functional heterogeneity, we analyzed responses to partial vs full agonist Ag. An extended period of interaction with APCs was required to set the threshold required for cell division in response to partial as compared with full agonist Ag. Acquisition of cytolytic function was restricted to the divided T cell population. In contrast, the threshold for commitment to produce IFN-gamma and express some activation markers appeared lower and independent of cell division. Indeed, we characterized a T cell population stimulated in response to the partial agonist that was committed to produce IFN-gamma, but failed to divide or secrete IL-2. Importantly, this activated nondivided population behaved as "primed" rather than "anergized," indicating 1) that priming of CD8 T cells may be induced by suboptimal stimulation independent of cell division and 2) that encounter with Ag does not always induce a complete differentiation program in naive CD8 T cells, as previously reported.  相似文献   

19.
Epitope-specific CD8+ T lymphocytes may play an important role in controlling human immunodeficiency virus (HIV)/simian immunodeficiency virus replication. Unfortunately, standard cellular assays do not measure the antiviral efficacy (the ability to suppress virus replication) of CD8+ T lymphocytes. Certain epitope-specific CD8+ T lymphocytes may be better than others at suppressing viral replication. We compared the antiviral efficacy of two immunodominant CD8+ T lymphocyte responses--Tat(28-35)SL8 and Gag(181-189)CM9--by using a functional in vitro assay. Viral suppression by Tat-specific CD8+ T lymphocytes was consistently greater than that of Gag-specific CD8+ T lymphocytes. Such differences in antigen-specific CD8+-T-lymphocyte efficacy may be important for selecting CD8+ T lymphocyte epitopes for inclusion in future HIV vaccines.  相似文献   

20.
T4 DNA ligase is more than an effective trap of cyclized dsDNA   总被引:1,自引:1,他引:0  
T4 DNA ligase is used in standard cyclization assays to trap double-stranded DNA (dsDNA) in low-probability, cyclic or highly bent conformations. The cyclization probability, deduced from the relative yield of cyclized product, can be used in conjunction with statistical mechanical models to extract the bending stiffness of dsDNA. By inserting the base analog 2-aminopurine (2-AP) at designated positions in 89bp and 94bp dsDNA fragments, we find that T4 DNA ligase can have a previously unknown effect. Specifically, we observe that addition of T4 ligase to dsDNA in proportions comparable to what is used in the cyclization assay leads to a significant increase in fluorescence from 2-AP. This effect is believed to originate from stabilization of local base-pair opening by formation of transient DNA-ligase complexes. Non-specific binding of T4 ligase to dsDNA is also confirmed using fluorescence correlation spectroscopy (FCS) experiments, which reveal a systematic reduction of dsDNA diffusivity in the presence of ligase. ATP competes with regular DNA for non-covalent binding to the T4 ligase and is found to significantly reduce DNA-ligase complexation. For short dsDNA fragments, however, the population of DNA-ligase complexes at typical ATP concentrations used in DNA cyclization studies is determined to be large enough to dominate the cyclization reaction.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号