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 31 毫秒
1.
Apoptotic caspases, such as caspase-7, are stored as inactive protease zymogens, and when activated, lead to a fate-determining switch to induce cell death. We previously discovered small molecule thiol-containing inhibitors that when tethered revealed an allosteric site and trapped a conformation similar to the zymogen form of the enzyme. We noted three structural transitions that the compounds induced: (i) breaking of an interaction between Tyr-223 and Arg-187 in the allosteric site, which prevents proper ordering of the catalytic cysteine; (ii) pinning the L2′ loop over the allosteric site, which blocks critical interactions for proper ordering of the substrate-binding groove; and (iii) a hinge-like rotation at Gly-188 positioned after the catalytic Cys-186 and Arg-187. Here we report a systematic mutational analysis of these regions to dissect their functional importance to mediate the allosteric transition induced by these compounds. Mutating the hinge Gly-188 to the restrictive proline causes a massive ∼6000-fold reduction in catalytic efficiency. Mutations in the Arg-187–Tyr-223 couple have a far less dramatic effect (3–20-fold reductions). Interestingly, although the allosteric couple mutants still allow binding and allosteric inhibition, they partially relieve the mutual exclusivity of binding between inhibitors at the active and allosteric sites. These data highlight a small set of residues critical for mediating the transition from active to inactive zymogen-like states.Caspases are a family of dimeric cysteine proteases whose members control the ultimate steps for apoptosis (programmed cell death) or innate inflammation among others (for reviews, see Refs. 1 and 2). During apoptosis, the upstream initiator caspases (caspase-8 and -9) activate the downstream executioner caspases (caspase-3, -6, and-7) via zymogen maturation (3). The activated executioner caspases then cleave upwards of 500 key proteins (46) and DNA, leading to cell death. Due to their pivotal role in apoptosis, the caspases are involved both in embryonic development and in dysfunction in diseases including cancer and stroke (7). The 11 human caspases share a common active site cysteine-histidine dyad (8), and derive their name, cysteine aspartate proteases, from their exquisite specificity for cleaving substrate proteins after specific aspartate residues (913). Thus, it has been difficult to develop active site-directed inhibitors with significant specificity for one caspase over the others (14). Despite difficulties in obtaining specificity, there has been a long-standing correlation between efficacy of caspase inhibitors in vitro and their ability to inhibit caspases and apoptosis in vivo (for review, see Ref. 31). Thus, a clear understanding of in vitro inhibitor function is central to the ability control caspase function in vivo.Caspase-7 has been a paradigm for understanding the structure and dynamics of the executioner caspases (1521). The substrate-binding site is composed of four loops; L2, L3, and L4 are contributed from one-half of the caspase dimer, and L2′ is contributed from the other half of the caspase dimer (Fig. 1). These loops appear highly dynamic as they are only observed in x-ray structures when bound to substrate or substrate analogs in the catalytically competent conformation (1719, 22) (Fig. 1B).Open in a separate windowFIGURE 1.Allosteric site and dimeric structure in caspase-7. A, the surface of active site-bound caspase-7 shows a large open allosteric (yellow) site at the dimer interface. This cavity is distinct from the active sites, which are bound with the active site inhibitor DEVD (green sticks). B, large subunits of caspase-7 dimers (dark green and dark purple) contain the active site cysteine-histidine dyad. The small subunits (light green and light purple) contain the allosteric site cysteine 290. The conformation of the substrate-binding loops (L2, L2′, L3, and L4) in active caspase-7 (Protein Data Bank (PDB) number 1f1j) is depicted. The L2′ loop (spheres) from one-half of the dimer interacts with the L2 loop from the other half of the dimer. C, binding of allosteric inhibitors influences the conformation of the L2′ loop (spheres), which folds over the allosteric cavity (PDB number 1shj). Subunit rendering is as in panel A. Panels A, B, and C are in the same orientation.A potential alternative to active site inhibitors are allosteric inhibitors that have been seeded by the discovery of selective cysteine-tethered allosteric inhibitors for either apoptotic executioner caspase-3 or apoptotic executioner caspase-7 (23) as well as the inflammatory caspase-1 (24). These thiol-containing compounds bind to a putative allosteric site through disulfide bond formation with a thiol in the cavity at the dimer interface (Fig. 1A) (23, 24). X-ray structures of caspase-7 bound to allosteric inhibitors FICA3 and DICA (Fig. 2) show that these compounds trigger conformational rearrangements that stabilize the inactive zymogen-like conformation over the substrate-bound, active conformation. The ability of small molecules to hold mature caspase-7 in a conformation that mimics the naturally occurring, inactive zymogen state underscores the utility and biological relevance of the allosteric mechanism of inhibition. Several structural changes are evident between these allosterically inhibited and active states. (i) The allosteric inhibitors directly disrupt an interaction between Arg-187 (next to the catalytic Cys-186) and Tyr-223 that springs the Arg-187 into the active site (Fig. 3), (ii) this conformational change appears to be facilitated by a hinge-like movement about Gly-188, and (iii) the L2′ loop folds down to cover the allosteric inhibitor and assumes a zymogen-like conformation (Fig. 1C) (23).Open in a separate windowFIGURE 2.Structure of allosteric inhibitors DICA and FICA. DICA and FICA are hydrophobic small molecules that bind to an allosteric site at the dimer interface of caspase-7. Binding of DICA/FICA is mediated by a disulfide between the compound thiol and Cys-290 in caspase-7.Open in a separate windowFIGURE 3.Movement of L2′ blocking arm. The region of caspase-7 encompassing the allosteric couple Arg-187 and Tyr-223 is boxed. The inset shows the down orientation of Arg-187 and Tyr-223 in the active conformation with DEVD substrate mimic (orange spheres) in the active site. In the allosteric/zymogen conformation, Arg-187 and Tyr-223 are pushed up by DICA (blue spheres).Here, using mutational analysis and small molecule inhibitors, we assess the importance of these three structural units to modulate both the inhibition of the enzyme and the coupling between allosteric and active site labeling. Our data suggest that the hinge movement and pinning of the L2-L2′ are most critical for transitioning between the active and inactive forms of the enzyme.  相似文献   

2.
The gene rapL lies within the region of the Streptomyces hygroscopicus chromosome which contains the biosynthetic gene cluster for the immunosuppressant rapamycin. Introduction of a frameshift mutation into rapL by ΦC31 phage-mediated gene replacement gave rise to a mutant which did not produce significant amounts of rapamycin. Growth of this rapL mutant on media containing added l-pipecolate restored wild-type levels of rapamycin production, consistent with a proposal that rapL encodes a specific l-lysine cyclodeaminase important for the production of the l-pipecolate precursor. In the presence of added proline derivatives, rapL mutants synthesized novel rapamycin analogs, indicating a relaxed substrate specificity for the enzyme catalyzing pipecolate incorporation into the macrocycle.Rapamycin is a 31-member macrocyclic polyketide produced by Streptomyces hygroscopicus NRRL 5491 which, like the structurally related compounds FK506 and immunomycin (Fig. (Fig.1),1), has potent immunosuppressive properties (24). Such compounds are potentially valuable in the treatment of autoimmune diseases and in preventing the rejection of transplanted tissues (16). The biosynthesis of rapamycin requires a modular polyketide synthase, which uses a shikimate-derived starter unit (11, 20) and which carries out a total of fourteen successive cycles of polyketide chain elongation that resemble the steps in fatty acid biosynthesis (2, 27). l-Pipecolic acid is then incorporated (21) into the chain, followed by closure of the macrocyclic ring, and both these steps are believed to be catalyzed by a pipecolate-incorporating enzyme (PIE) (18), the product of the rapP gene (8, 15). Further site-specific oxidations and O-methylation steps (15) are then required to produce rapamycin. Open in a separate windowFIG. 1Structures of rapamycin, FK506, and immunomycin.The origin of the pipecolic acid inserted into rapamycin has been previously established (21) to be free l-pipecolic acid derived from l-lysine (although the possible role of d-lysine as a precursor must also be borne in mind) (9). Previous work with other systems has suggested several alternative pathways for pipecolate formation from lysine (22), but the results of the incorporation of labelled lysine into the pipecolate moiety of immunomycin (Fig. (Fig.1)1) clearly indicate loss of the α-nitrogen atom (3). More recently, the sequencing of the rap gene cluster revealed the presence of the rapL gene (Fig. (Fig.2),2), whose deduced gene product bears striking sequence similarity to two isoenzymes of ornithine deaminase from Agrobacterium tumefaciens (25, 26). Ornithine deaminase catalyzes the deaminative cyclization of ornithine to proline, and we have proposed (15) that the rapL gene product catalyzes the analogous conversion of l-lysine to l-pipecolate (Fig. (Fig.3).3). Open in a separate windowFIG. 2A portion of the rapamycin biosynthetic gene cluster which contains ancillary (non-polyketide synthase) genes (15, 27). PKS, polyketide synthase.Open in a separate windowFIG. 3(A) The conversion of l-ornithine to l-proline by ornithine cyclodeaminase (17). (B) Proposed conversion of l-lysine to l-pipecolic acid by the rapL gene product.Here, we report the use of ΦC31 phage-mediated gene replacement (10) to introduce a frameshift mutation into rapL and the ability of the mutant to synthesize rapamycins in the absence or presence of added pipecolate or pipecolate analogs.  相似文献   

3.
FTY720, a sphingosine analog, is in clinical trials as an immunomodulator. The biological effects of FTY720 are believed to occur after its metabolism to FTY720 phosphate. However, very little is known about whether FTY720 can interact with and modulate the activity of other enzymes of sphingolipid metabolism. We examined the ability of FTY720 to modulate de novo ceramide synthesis. In mammals, ceramide is synthesized by a family of six ceramide synthases, each of which utilizes a restricted subset of acyl-CoAs. We show that FTY720 inhibits ceramide synthase activity in vitro by noncompetitive inhibition toward acyl-CoA and uncompetitive inhibition toward sphinganine; surprisingly, the efficacy of inhibition depends on the acyl-CoA chain length. In cultured cells, FTY720 has a more complex effect, with ceramide synthesis inhibited at high (500 nm to 5 μm) but not low (<200 nm) sphinganine concentrations, consistent with FTY720 acting as an uncompetitive inhibitor toward sphinganine. Finally, electrospray ionization-tandem mass spectrometry demonstrated, unexpectedly, elevated levels of ceramide, sphingomyelin, and hexosylceramides after incubation with FTY720. Our data suggest a novel mechanism by which FTY720 might mediate some of its biological effects, which may be of mechanistic significance for understanding its mode of action.FTY720 (2-amino-(2-2-[4-octylphenyl]ethyl)propane 1,3-diol hydrochloride), also known as Fingolimod, is an immunosuppressant drug currently being tested in clinical trials for organ transplantation and autoimmune diseases such as multiple sclerosis (1). FTY720 is a structural analog of sphingosine, a key biosynthetic intermediate in sphingolipid (SL)2 metabolism (see Fig. 1). In vivo, FTY720 is rapidly phosphorylated by sphingosine kinase 2 (2, 3) to form FTY720 phosphate (FTY720-P), an analog of sphingosine 1-phosphate (S1P) (see Fig. 1A). FTY720-P binds to S1P receptors (S1PRs) (4, 5) and thereby induces a variety of phenomena such as T-lymphocyte migration from lymphoid organs (69); accordingly, FTY720 treatment results in lymphopenia as lymphocytes (especially T-cells) become sequestered inside lymphoid organs (1012). The ability of FTY720 to sequester lymphocytes has stimulated its use in treatment of allograft rejection and autoimmune diseases (13), and FTY720 is currently under phase III clinical trials for treatment of relapsing-remitting multiple sclerosis (14).Open in a separate windowFIGURE 1.SL structure and metabolism. A, structures of SLs and SL analogs used in this study. B, metabolic inter-relationships between SLs and the metabolism of FTY720. The enzymes are denoted in italics. LPP3, lipid phosphate phosphatase 3; LPP1α, lipid phosphate phosphatase 1α.Apart from the binding of FTY720-P to S1PRs, the ability of FTY720 to inhibit S1P lyase (15) (see Fig. 1B), and its inhibitory effect on cytosolic phospholipase A2 (16), whose activity can be modulated by ceramide 1-phosphate (17), little is known about whether FTY720 or FTY720-P can modulate the activity of other enzymes of SL metabolism. Because FTY720 is an analog of sphingosine, one of the two substrates of ceramide synthase (CerS) (see Fig. 1), we now examine whether FTY720 can modulate CerS activity. CerS utilizes fatty acyl-CoAs to N-acylate sphingoid long chain bases. Six CerS exist in mammals, each of which uses a restricted subset of acyl-CoAs (1823). We demonstrate that FTY720 inhibits CerS activity and that the extent of inhibition varies according to the acyl chain length of the acyl-CoA substrate. Surprisingly, FTY720 inhibits CerS activity toward acyl-CoA via noncompetitive inhibition and toward sphinganine via uncompetitive inhibition. Finally, the mode of interaction of FTY720 with CerS in cultured cells depends on the amount of available sphinganine. Together, we show that FTY720 modulates ceramide synthesis, which may be of relevance for understanding its biological effects in vivo and its role in immunomodulation.  相似文献   

4.
Folding and stability are parameters that control protein behavior. The possibility of conferring additional stability on proteins has implications for their use in vivo and for their structural analysis in the laboratory. Cyclic polypeptides ranging in size from 14 to 78 amino acids occur naturally and often show enhanced resistance toward denaturation and proteolysis when compared with their linear counterparts. Native chemical ligation and intein-based methods allow production of circular derivatives of larger proteins, resulting in improved stability and refolding properties. Here we show that circular proteins can be made reversibly with excellent efficiency by means of a sortase-catalyzed cyclization reaction, requiring only minimal modification of the protein to be circularized.Sortases are bacterial enzymes that predominantly catalyze the attachment of surface proteins to the bacterial cell wall (1, 2). Other sortases polymerize pilin subunits for the construction of the covalently stabilized and covalently anchored pilus of the Gram-positive bacterium (35). The reaction catalyzed by sortase involves the recognition of short 5-residue sequence motifs, which are cleaved by the enzyme with the concomitant formation of an acyl enzyme intermediate between the active site cysteine of sortase and the carboxylate at the newly generated C terminus of the substrate (1, 68). In many bacteria, this covalent intermediate can be resolved by nucleophilic attack from the pentaglycine side chain in a peptidoglycan precursor, resulting in the formation of an amide bond between the pentaglycine side chain and the carboxylate at the cleavage site in the substrate (9, 10). In pilus construction, alternative nucleophiles such as lysine residues or diaminopimelic acid participate in the transpeptidation reaction (3, 4).When appended near the C terminus of proteins that are not natural sortase substrates, the recognition sequence of Staphylococcus aureus sortase A (LPXTG) can be used to effectuate a sortase-catalyzed transpeptidation reaction using a diverse array of artificial glycine-based nucleophiles (Fig. 1). The result is efficient installation of a diverse set of moieties, including lipids (11), carbohydrates (12), peptide nucleic acids (13), biotin (14), fluorophores (14, 15), polymers (16), solid supports (1618), or peptides (15, 19) at the C terminus of the protein substrate. During the course of our studies to further expand sortase-based protein engineering, we were struck by the frequency and relative ease with which intramolecular transpeptidation reactions were occurring. Specifically, proteins equipped with not only the LPXTG motif but also N-terminal glycine residues yielded covalently closed circular polypeptides (Fig. 1). Similar reactivity using sortase has been described in two previous cases; however, rigorous characterization of the circular polypeptides was absent (16, 20). The circular proteins in these reports were observed as minor components of more complex reaction mixtures, and the cyclization reaction itself was not optimized.Open in a separate windowFIGURE 1.Protein substrates equipped with a sortase A recognition sequence (LPXTG) can participate in intermolecular transpeptidation with synthetic oligoglycine nucleophiles (left) or intramolecular transpeptidation if an N-terminal glycine residue is present (right).Here we describe our efforts toward applying sortase-catalyzed transpeptidation to the synthesis of circular and oligomeric proteins. This method has general applicability, as illustrated by successful intramolecular reactions with three structurally unrelated proteins. In addition to circularization of individual protein units, the multiprotein complex AAA-ATPase p97/VCP/CDC48, with six identical subunits containing the LPXTG motif and an N-terminal glycine, was found to preferentially react in daisy chain fashion to yield linear protein fusions. The reaction exploited here shows remarkable similarities to the mechanisms proposed for circularization of cyclotides, small circular proteins that have been isolated from plants (2123).  相似文献   

5.
A decade after its inception, MALDI imaging mass spectrometry has become a unique technique in the proteomics arsenal for biomarker hunting in a variety of diseases. At this stage of development, it is important to ask whether we can consider this technique to be sufficiently developed for routine use in a clinical setting or an indispensable technology used in translational research. In this report, we consider the contributions of MALDI imaging mass spectrometry and profiling technologies to clinical studies. In addition, we outline new directions that are required to align these technologies with the objectives of clinical proteomics, including: 1) diagnosis based on profile signatures that complement histopathology, 2) early detection of disease, 3) selection of therapeutic combinations based on the individual patient''s entire disease-specific protein network, 4) real time assessment of therapeutic efficacy and toxicity, 5) rational redirection of therapy based on changes in the diseased protein network that are associated with drug resistance, and 6) combinatorial therapy in which the signaling pathway itself is viewed as the target rather than any single “node” in the pathway.MS has become a versatile tool that we are familiar with in large part due to important electronic and informatics advancements. The ability to obtain the molecular weight is one of the first steps in the identification of a molecule. With the addition of primary structural information mass spectrometry has become a useful technique to identify molecules within complex mixtures.Biological specimens, such as tissues, urine, or plasma, are complex and highly heterogeneous, which makes them inherently difficult to analyze. Further research and developments are necessary to achieve reliable biological models for understanding and studying pathologies. Therefore, it is of primary importance to identify the constituents of these systems and subsequently understand how they function within the framework of the tissue. With regard to clinical proteomics, there is the added dimension of disease, and therefore, the main goal is to characterize the cellular circuitry with a focus on the impact of the disease and/or therapy on these cellular networks.Mass spectrometry has become a centerpiece technology predominantly in the field of proteomics. Nonetheless a more comprehensive understanding of the constituents of biological systems will be aided by determining the constituent distribution. This anatomical dimension has been added through mass spectrometry imaging (MSI)1 especially using MALDI-MSI.MALDI is an ion source that is well compatible with the introduction of raw materials and surfaces. Shortly after its introduction, MALDI was used for direct tissue profiling. The first applications were neurobiological studies on dissected organs from the mollusk Lymnaea stagnalis (18), crustaceans (9), and other mollusks (10, 11). More recently, MALDI was used to generate profiles from tissue sections and ion images using a scanning method to analyze the surface (12) (Fig. 1). This led to the first MALDI MS tissue section imaging micrographs in 1997 (1315). These studies were followed by 10 years of intense efforts to improve the sensitivity, reproducibility, data processing, tissue preservation, and preparation treatments to fully characterize the proteome leading to a clear improvement of molecular images (1639) (Fig. 2).Open in a separate windowFig. 1.Schematic representation of the MALDI-MSI work flow. After tissue sectioning and transfer onto a conductive and transparent sample plate, the MALDI matrix is deposited, and data are acquired by recording mass spectra according to a raster of points covering the surface to be analyzed. Mass spectra recorded with their coordinates on the tissue are processed, and molecular images of the localization of molecules can be reconstructed. a.u., arbitrary units; ITO, idium tin oxide.Open in a separate windowFig. 2.Ten years'' evolution from one of the first MALDI images presented in 1999 at the 47th ASMS Conference on Mass Spectrometry and Allied Topics (left) (reprinted with permission of Caprioli and co-workers (84)) and molecular images obtained by our group for mouse stem cells injected in brain tissue sections (right) (M. Wisztorski, C. Meriaux, M. Salzet, and I. Fournier, unpublished results).These developments led to clinical studies using MALDI-MSI technology. Clinical proteomics has many objectives including 1) diagnosis based on signatures as a complement to histopathology, 2) early disease detection, 3) individualized selection of therapeutic combinations that best target the patient''s entire disease-specific protein network, 4) real time assessment of therapeutic efficacy and toxicity, 5) rational redirection of therapy based on changes in the diseased protein network that are associated with drug resistance, and 6) combinatorial therapy in which the signaling pathway itself is viewed as the target rather than any single “node” in the pathway.Based on these key objectives, can we consider MALDI-MSI a mature technology for use in clinical studies? What is the potential impact of this technology in anatomy/pathology and disease? By reviewing each objective, do we have sufficient evidence that MALDI-MSI satisfies the criteria imposed by clinical proteomics? We will now specifically address each of these key points.  相似文献   

6.
7.
8.
9.
The study of metabolically labeled or probe-modified proteins is an important area in chemical proteomics. Isolation and purification of the protein targets is a necessary step before MS identification. The biotin-streptavidin system is widely used in this process, but the harsh denaturing conditions also release natively biotinylated proteins and non-selectively bound proteins. A cleavable linker strategy is a promising approach for solving this problem. Though several cleavable linkers have been developed and tested, an efficient, easily synthesized, and inexpensive cleavable linker is a desirable addition to the proteomics toolbox. Here, we describe the chemical proteomics application of a vicinal diol cleavable linker. Through easy-to-handle chemistry we incorporate this linker into an activity-based probe and a biotin alkyne tag amenable for bioorthogonal ligation. With these reagents, background protein identifications are significantly reduced relative to standard on-bead digestion.The covalent modification of proteins by small molecules within a complex proteome is a major theme in chemical biology and proteomics. An effective method for the detection of posttranslational modifications of proteins is the metabolic incorporation of modified biomolecules such as tagged carbohydrates or lipids (1). Reversible interactions of enzyme inhibitors, natural products, or drugs can be detected when one appends photocrosslinking agents, thereby facilitating target discovery (2, 3). A particularly interesting example of protein labeling is activity-based protein profiling (ABPP)1 (4, 5), which utilizes the intrinsic catalytic activity of a target enzyme for the covalent attachment of an affinity or visualization tag. ABPP makes use of small molecules (activity-based probes (ABPs)) that react with the active form of a specific enzyme or enzyme class by means of a “warhead,” which is often derived from a mechanism-based enzyme inhibitor (Fig. 1A). DCG-04, for example, is based on the naturally occurring inhibitor E-64 and targets the papain family of cysteine proteases via covalent attachment of the epoxysuccinate group to the active site cysteine (Fig. 1B) (6).Open in a separate windowFig. 1.The cleavable linker strategy in ABPP. A, the elements of an ABP. B, the example ABP DCG-04, an epoxysuccinate-containing probe for clan CA cysteine proteases. DCG-04 is based on the naturally occurring protease inhibitor E-64. C, schematic strategy of cleavable linker-mediated target identification. D, the cleavage mechanism of a vicinal diol.Bulky fluorophore or biotin tags on chemical probes might interfere with efficient protein binding. Moreover, they can negatively influence the cell permeability of probes, which therefore limits their applicability in in vitro experiments. Bioorthogonal chemistries, such as the Bertozzi-Staudinger ligation (7) and the 1,3-bipolar cycloaddition of an azide and an alkyne (click chemistry) (8), allow tandem labeling strategies in which a biotin or a fluorophore is attached to an enzyme probe complex in a separate step. Consequently, the probes themselves only carry azide or alkyne groups as “mini-tags.” Tandem labeling using bioorthogonal chemistry has now become a widely used strategy to label biomolecules in lysates and in live cells (911).An essential step in ABPP, as well as in other chemical proteomics approaches, is the elucidation of the tagged proteins. This usually involves a biotin-mediated enrichment step followed by mass-spectrometry-based identification. Although the streptavidin-biotin interaction allows efficient enrichment as a result of the strong binding affinity (Kd ∼ 10−15 m), it also has limitations. The quantitative elution of biotinylated proteins requires harsh conditions (12), which lead to contamination of the sample by endogenous biotinylated and non-specifically bound proteins. These other proteins will be identified together with the real protein targets. Given that subsequent target validation with secondary assays can be a costly and time-consuming process, a reduction in false positive identifications is highly desirable. For cleaner protein identification, cleavable linker strategies (13) that allow the selective release of target proteins have been developed (Fig. 1C). The commercially available disulfide linker can be cleaved under mild conditions, but it suffers from premature cleavage in reducing media such as the intracellular environment and reducing buffers used for click chemistry and in vitro reactions of cysteine proteases. Therefore, a variety of alternative linkers for proteomics applications have been reported, including a sterically hindered disulfide (14), diazobenzenes (1519), hydrazones (20, 21), silanes (22), light sensitive linkers (2325), tobacco etch virus protease sensitive linkers (26, 27), and a levulinoyl-based linker (28). The synthesis of some of these linkers is lengthy or difficult to scale up, which limits their general application in chemical proteomics.Ideally, a cleavable linker is stable under a wide variety of conditions, is efficiently and selectively cleaved, and can be synthesized in a low number of easy chemical transformations. We aimed to meet these requirements by using a vicinal diol as a cleavable linker system. When vicinal diols are treated with sodium periodate (NaIO4), the carbon–carbon bond is cleaved (Fig. 1D). Periodate treatment of proteins can result in side-reactions, such as the cleavage of linked carbohydrates or the oxidation of N-terminal serine and threonine residues. However, these N-termini rarely occur in proteins and are therefore of minor concern. In general, the mild, neutral conditions of periodate cleavage are compatible with proteins. This has been illustrated in the past, for example, by its application in the detection of protein–protein interactions (29) and the creation of unliganded MHC class I molecules (30). In this article, we report the chemical proteomics application of diol cleavable linker probes. We show that the synthesis of the linker and its probe derivatives is straightforward, that the linker is compatible with tandem click labeling, that enrichment and release of probe targets is efficient, and that the identification of targets takes place with significantly lower background than in on-bead digestion protocols.  相似文献   

10.
11.
12.
13.
All of the 2,6-dideoxy sugars contained within the structure of chromomycin A3 are derived from d-glucose. Enzyme assays were used to confirm the presence of hexokinase, phosphoglucomutase, UDPG pyrophosphorylase (UDPGP), and UDPG oxidoreductase (UDPGO), all of which are involved in the pathway of glucose activation and conversion into 2,6-dideoxyhexoses during chromomycin biosynthesis. Levels of the four enzymes in Streptomyces spp. cell extracts were correlated with the production of chromomycins. The pathway of sugar activation in Streptomyces spp. involves glucose 6-phosphorylation by hexokinase, isomerization to G-1-P catalyzed by phosphoglucomutase, synthesis of UDPG catalyzed by UDPGP, and formation of UDP-4-keto-6-deoxyglucose by UDPGO.Dideoxy sugars occur commonly in the structures of cardiac glycosides from plants, in antibiotics like chromomycin A3 (Fig. (Fig.1),1), and in macrolides produced by microorganisms. On the basis of stable isotope-labeling experiments, biosynthetic studies conducted in Rosazza’s laboratory have indicated that all the deoxy sugars of chromomycin A3 are derived from d-glucose (21). While the assembly of the polyketide aglycone is reasonably well understood, relatively little is known of the details of 2,6-dideoxy sugar biogenesis in streptomycetes. Earlier studies with Streptomyces rimosus indicated that TDP-mycarose is synthesized from TDP-d-glucose (TDPG) and S-adenosyl-l-methionine (10, 23). The reaction requires NADPH as a cofactor, and TDP-4-keto-6-deoxy-d-glucose is an intermediate. Formation of TDP-4-keto-6-deoxy-d-glucose was catalyzed by the enzyme TDPG oxidoreductase (TDPG-4,6-dehydratase; EC 4.2.1.46). Similar 4-keto sugar nucleotides are intermediates for the biosynthesis of polyene macrolide antibiotic amino sugars (18). Similar pathways have been elaborated for the formation of 2,6-dideoxy-d-threo-4-hexulose of granaticin in Escherichia coli (6, 25) and 2,6-dideoxy-d-arabino-hexose of chlorothricin (12). The initial 6-deoxygenation of glucose during 3,6-dideoxy sugar formation involves a similar mechanism (32). In all of these processes, glucose is first activated by conversion into a sugar nucleotide such as UDPG followed by NAD+ oxidation of the 4 position to the corresponding 4-oxo derivative. Position 6 deoxygenation involves a general tautomerization, dehydration, and NADH,H+-catalyzed reduction process (6, 12, 25). A similar tautomerization and dehydration followed by reduction may produce C-3-deoxygenated products, such as CDP-3,6-dideoxyglucose (27). The pathway for formation of 3,6-dideoxyhexoses from CDPG in Yersinia pseudotuberculosis was clearly elucidated by Liu and Thorson (14). However, none of this elegant work was focused on the earlier steps of hexose nucleotide formation. Open in a separate windowFIG. 1Structures of chromomycins A2 and A3.On the basis of previous work (7), it is reasonable to postulate that the biosynthesis of 2,6-dideoxyglucose in Streptomyces griseus involves phosphorylation to glucose-6-phosphate by hexokinase (HK; E.C.2.7.7.1), as in glycolysis; conversion to glucose-1-phosphate by phosphoglucomutase (PGM; EC 2.7.5.1); reaction with UTP to form UDPG in a reaction catalyzed by UDPG pyrophosphorylase (UDPGP) (glucose-1-phosphate uridylyltransferase; EC 2.7.7.9), and C-6 deoxygenation catalyzed by UDP-d-glucose-4,6-dehydratase with NAD+ as a cofactor (Fig. (Fig.2).2). UDPG and GDPG have been detected in cell extracts of S. griseus and Streptomyces sp. strain MRS202, suggesting that these compounds are active sugar nucleotides involved in the formation of dideoxyhexoses (15). UDPGP genes from several bacteria have been cloned and sequenced (1, 3, 4, 11, 29, 30). Although nucleotidyl diphosphohexose-4,6-dehydratases (NDP-hexose-4,6-dehydratases) have been purified and characterized from several sources (5, 8, 9, 13, 19, 25, 26, 31, 33), the occurrence of the glucose-activating enzymes HK, PGM, UDPGP, and UDPG oxidoreductase (UDPGO) involved in 2,6-dideoxyhexose formation has not been established in streptomycetes. This work provides evidence for the presence of these enzymes involved in the biosynthetic activation of glucose to the 2,6-dideoxyhexoses in chromomycin A3.Open in a separate windowFIG. 2Proposed pathway for the formation of 2,6-dideoxy sugars in streptomycetes involving HK, PGM, UDPGP, and UDPGO.  相似文献   

14.
Saccharomyces cerevisiae plasma membrane H+-ATPase activity was stimulated during octanoic acid-induced latency, reaching maximal values at the early stages of exponential growth. The time-dependent pattern of ATPase activation correlated with the decrease of cytosolic pH (pHi). The cell population used as inoculum exhibited a significant heterogeneity of pHi, and the fall of pHi correlated with the loss of cell viability as determined by plate counts. When exponential growth started, only a fraction of the initial population was still viable, consistent with the role of the physiology and number of viable cells in the inoculum in the duration of latency under acid stress.The biological target sites of octanoic acid in Saccharomyces cerevisiae may be related to processes of transport across membranes, particularly the plasma membrane (21). Like other weak acids at low pH, octanoic acid, a highly toxic by-product of yeast alcoholic fermentation (23) and an antimicrobial food additive (6), leads to the reduction of cytosolic pH (pHi) due to its dissociation in the approximately neutral cytoplasm following the entrance of the undissociated toxic form into the cell by passive diffusion (5, 20, 23). It is likely that this highly liposoluble weak acid significantly affects the spatial organization of the plasma membrane, affecting its function as a matrix for enzymes and as a selective barrier, thereby leading to the dissipation of the proton motive force across the plasma membrane and to intracellular acidification (16, 18). Significantly, the H+-ATPase in the plasma membrane in yeast, which creates the electrochemical proton gradient that drives the secondary transport of solutes and is implicated in the maintenance of pHi around neutrality, has been pointed out as a critical component of yeast adaptation to weak acids (8, 19, 24). Indeed, yeast plasma membrane H+-ATPase is activated during exponential growth with octanoic acid (19, 24), and the duration of lag phase before yeast cells enter exponential growth in the presence of sorbic acid is significantly extended in a mutant with reduced levels of plasma membrane ATPase activity (8). The activation of the H+-ATPase in the plasma membrane in yeast cells exposed to other stresses that also lead to the dissipation of the H+ gradient and intracellular acidification (such as subcritical inhibitory concentrations of ethanol [12, 14, 15], supraoptimal temperatures below 40°C [25], presence of other organic acids at low pH [1, 5, 8], and deprivation of nitrogen source [2]) have also been observed. Several lines of evidence indicate that ATPase activation is due to posttranslational modifications of the PMA1 ATPase (2, 12, 24, 25). Considerable information has been obtained on the variation of plasma membrane ATPase activity during exponential growth and early stationary phase of yeast cells cultivated in media, at low pH, supplemented or not with octanoic acid (24). However, this is not the case during the period of latency preceding exponential growth at concentrations of octanoic acid close to the maximal concentration allowing growth. The main objective of the present work was to obtain information about the pattern of ATPase activity and the changes in pHi and cell viability during the lag phase necessary for yeast adaptation to the physiological effects of octanoic acid before exponential growth.

Duration of yeast growth latency in octanoic acid-supplemented media.

When cells of S. cerevisiae IGC3507III grown, at 30°C, in medium that had not been supplemented with octanoic acid were used to inoculate buffered YG media (30 g of glucose liter−1, 6.7 g of Yeast Nitrogen Base [Difco] liter−1) (pH 4.0) supplemented with increasing concentrations of this toxic acid up to around 0.35 mM total acid (19, 23), exponential growth was initiated without significant delay (Fig. (Fig.1a),1a), although a dose-dependent decrease in specific growth rate was observed (Fig. (Fig.1b).1b). However, for higher concentrations up to the maximal that allowed growth (0.42 mM), a lag phase was observed and its duration strongly increased with the severity of octanoic acid stress (Fig. (Fig.1a).1a). The duration of latency was drastically reduced when exponential cells used as inoculum were grown in medium with an identical concentration of octanoic acid (Fig. (Fig.1a),1a), but the specific growth rate was not modified (Fig. (Fig.1b).1b). At a concentration of total octanoic acid of 0.39 mM, a lag phase of around 55 h was necessary for yeast cells, which had been cultivated under nonstressing conditions, to adapt to the deleterious effects of octanoic acid and to initiate inhibited exponential growth (Fig. (Fig.2).2). Open in a separate windowFIG. 1Effect of the addition to the growth medium of increasing concentrations of octanoic acid on the duration of lag phase (a) and the specific growth rate of S. cerevisiae IGC 3507III (b) for exponentially growing cells (used as inoculum) cultivated at 30°C at pH 4.0 in the absence (○) or presence (•) of concentrations of toxic lipophilic acid identical to those present in the growth medium. Results are representative of the many growth experiments carried out.Open in a separate windowFIG. 2Specific activity of plasma membrane H+-ATPase (filled symbols) and growth curve (open symbols) of S. cerevisiae IGC 3507III during cultivation in the presence (a) or absence (b) of 0.39 mM total octanoic acid (at pH 4.0, 30°C). The data are averages with standard deviations for at least three enzyme assays using cells from at least two independent growth experiments. OD, optical density.

Activation of plasma membrane ATPase during octanoic acid-induced latency.

The specific activity of plasma membrane ATPase assayed in crude membrane suspensions prepared from nonadapted cells, as previously reported (19, 25), during cultivation in buffered medium (at pH 4.0) supplemented with 0.39 mM octanoic acid, increased during the 55 h of latency (Fig. (Fig.2a).2a). A peak of activity was reached during the early stages of exponential growth and values of ATPase activity were consistently higher (twofold) in cells grown under octanoic acid stress (Fig. (Fig.2),2), as described by Viegas et al. (24). Yeast cells must adapt to the physiological effects of octanoic acid during an extended lag period, the length of which depended on the severity of acid stress (Fig. (Fig.1a),1a), before eventually recovering and entering exponential growth; the activation of plasma membrane H+-ATPase observed during this period of latency reinforces the idea that this proton pump is an important component of this adaptative response (5, 8, 19, 24). In fact, the ability of yeast cells to grow in the presence of lipophilic acids at a low pH reflects their capacity to maintain control over their internal pH by excluding protons. This adaptative phenomenon, reported for the first time in the present work, complements the observation of Holyoak et al. (8) that a strain with reduced plasma membrane H+-ATPase activity displayed increased lag phase in the presence of the weak-acid preservative sorbic acid. Significantly, plasma membrane H+-ATPase activity was also pointed out to play a critical role in yeast tolerance of ethanol (15) or supraoptimal temperatures (13, 25). The mechanism underlying plasma membrane ATPase activation during octanoic acid-induced latency remains obscure at the present time, but it is likely that this is due to a posttranslational modification of ATPase, as proposed for ATPase activation during octanoic acid-stressed exponential growth (24). It is likely that during lag phase the amount of H+-ATPase in the plasma membrane slightly decreases, as found by Benito et al. (2) in yeast cells deprived of nitrogen source where ATPase activation also occurred (2), as the estimated half-life of the enzyme is about 11 h (2). ATPase activation during latency can hardly be attributed to the adaptative modification of the ATPase lipid environment in cells grown under lipophilic acid stress, as suggested by Alexandre et al. (1).

Changes in yeast pHi and viability during octanoic acid-stressed cultivation.

The change in pHi during cultivation of nonadapted cells with 0.39 mM octanoic acid was monitored by using an adaptation of the fluorescence microscopic image processing technique developed by Imai and Ohno (9); 5- (and 6)-carboxyfluorescein (cF) was used as the internal pH-dependent fluoroprobe. Cells washed and resuspended in cold CF buffer (citrate-phosphate buffer [at pH 4.0] with 50 mM glycine [Sigma], 110 mM NaCl, 5 mM KCl, and 1 mM MgCl2) to a cellular density of 2 × 108 ml−1 were loaded with cF by adding 20 μM of 5 (and 6)-carboxyfluorescein-diacetate (Sigma) and vortexing in two bursts of 1 min each, interspersed with 15 min on ice (9). After being washed twice with cold CF buffer, cF-loaded cells were immediately examined with a Zeiss Axioplan microscope equipped with adequate epifluorescence interference filters (Zeiss BP450-490 and Zeiss LP520) and connected to a video camera and to a computer with an image- analysis program (gel documentation system SW2000; UVP, San Gabriel, Calif.). Following a cell-by-cell analysis, the value of fluorescence intensity (fI) emitted by each cell, measured by direct densitometry, corresponded to the arithmetical mean value of fI measured in two or three different regions in the cytoplasm of the same cell, with the less fluorescent vacuole excluded. To estimate average pHi, an in vivo calibration curve was prepared (Fig. (Fig.3)3) by using cell suspensions grown in the absence of toxics which were loaded with cF as described above and incubated, at 30°C, with 0.5 mM carbonyl cyanide m-chlorophenylhydrazone (CCCP) to dissipate the plasma membrane pH gradient (4), before adjustment of external pH (in the range 3.5 to 7.5) by the addition of HCl or NaOH at 2 M. Fluorescence images were fixed 15 s after the occurrence of the excitation radiation in order to minimize interferences due to leakage of cF as well as fluorescence quenching (3, 7). Cells were kept on ice throughout the procedure, and CF buffer lacked glucose; therefore, the active efflux of cF (3) was minimized as confirmed by measuring the fluorescence in the medium surrounding the cells, which was negligible. Under the experimental conditions used and for the purpose of the study, this technique proved to be highly useful and suitable despite the limitations that might be raised (3, 7). It allowed a clear-cut picture of the pHi of individual cells, giving information about the distribution of pHi values of a yeast population (Fig. (Fig.44 and and5a5a to c), instead of solely an estimation of the average value of the whole population, as is the case with techniques based on the distribution of radioactive propionic acid (20) or on the in vivo 31P nuclear magnetic resonance (5). Moreover, values calculated for the average pHi of the whole yeast population during latency and exponential growth in medium with octanoic acid (Fig. (Fig.5d)5d) were close to, although slightly lower than, the values previously obtained based on the distribution of [14C]propionic acid (20, 22). Results revealed that the cell population used to inoculate octanoic acid-supplemented medium exhibited a significant heterogeneity (Fig. (Fig.4);4); around 31% showed a pHi in the optimal range (above 6.5) (Fig. (Fig.4),4), with the average pHi value of the whole population estimated to be approximately 6.0. This low pHi value results from cell cultivation in a rich medium with high production of organic acids (11) (external pH, 3.6), followed by washing of the cells with YG medium buffered at pH 4.0 (17). The introduction of the inoculum in octanoic acid-supplemented medium led to the very rapid (5-min) increase in the percentage of the cell population with pHi below 5.5, consistent with the rapid kinetics of cytosol acidification when yeast cells are exposed to weak acids (5). During extended incubation with octanoic acid and until the end of latency, the percentage of the population with a very low pHi (below 5.5) continued to increase, reaching 80% of the cell population, while the percentage of cell population with a pHi above 6.0 suffered a corresponding decrease (Fig. (Fig.5).5). During exponential growth, the opposite pHi modification was observed, consistent with a recovery of pHi to physiological levels (Fig. (Fig.5).5). The time-dependent pattern of internal acidification during lag phase correlated with plasma membrane ATPase activation (Fig. (Fig.2a2a and and5),5), suggesting that this activation was triggered by intracellular acidification, as proposed for acetic acid (5)- or nitrogen starvation (2)-induced activation. Immediately before yeast cells entered exponential growth, 80% of the initial viable population had lost viability, as assessed by the number of CFU (21) (Fig. (Fig.6),6), suggesting that octanoic acid-induced death during latency is related to internal acidification down to critical values (Fig. (Fig.55 and and6),6), in agreement with the relationship established by Imai and Ohno (10) between yeast viability and intracellular pH. Only about 20% of the initial population was able to start cell division in octanoic acid-supplemented medium, presumably those cells that in the inoculum exhibited pHi values around neutrality (Fig. (Fig.55 and and6).6). These results suggest that despite plasma membrane H+- ATPase activation, this system of pH homeostasis may not be able to fully counteract the physiological effects of increasing octanoic acid concentrations and eventually fails at very severe acid stress. Open in a separate windowFIG. 3In vivo calibration curve, showing the pH dependence of the fI of cF-loaded-cells of S. cerevisiae IGC 3507III. Intracellular and extracellular pHs were equilibrated by incubation of cF-loaded cells, for 10 min at 30°C, with 0.5 mM CCCP. At each pH, values of fI correspond to the average fI of about 20 cells. The data are averages with standard deviations for three independent experiments.Open in a separate windowFIG. 4Distribution of cells with different pHi values present in the inoculum of S. cerevisiae IGC 3507III prepared in growth medium without octanoic acid supplementation.Open in a separate windowFIG. 5Percentage of yeast cells with pHi below 5.5 (a), between 5.5 and 6.0 (b), or above 6.0 (c); average pHi of the whole cell population (▴) during S. cerevisiae IGC3507III cultivation in medium supplemented with 0.39 mM total octanoic acid (pH 4.0, 30°C); and the optical density (OD) of the culture at 600 nm (▪). The average pHi values estimated for the whole cell population are the arithmetical mean values of the various average pHi values calculated for individual cells. The percentage of cells present in the inoculum with pHi values within the three ranges (○) and the average pHi of the inoculum cell population (▵) are indicated.Open in a separate windowFIG. 6Concentration of viable cells (▴) and culture optical density (O.D.) at 600 nm (□) during lag and exponential phases of S. cerevisiae IGC 3507III growth in medium supplemented with 0.39 mM octanoic acid, at pH 4.0 and 30°C.

Adaptative response to octanoic acid.

The adaptation of yeast cells to octanoic acid at a low pH appears to depend on their H+-exporting ability, but this requires not only a highly active H+-ATPase in the plasma membrane but the provision of sufficient ATP to drive this energy-demanding process as indicated by the results of Holyoak et al. (8). It is likely that increased ATPase activity under octanoic acid stress may reduce cellular ATP levels and that ATP depletion contributes to the failure of the maintenance of pHi homeostasis, particularly among the subpopulation that in the inoculum exhibited the lowest pHi values. The loss of viability might occur for those cells where pHi decreased down to nonphysiological values. The eventual recovery of growth therefore depends on the remaining viable population, in agreement with the well-known critical role played by the physiology and number of viable cells in the inoculum in the duration of latency under acid stress. The observation that octanoic acid-adapted cells reinoculated into the same fresh medium can resume growth after a much shorter latency (Fig. (Fig.1a)1a) is a good example of the importance of the physiology of the inoculum cells. Besides the increased plasma membrane H+-ATPase activity of octanoic acid-adapted cells, other mechanisms may underlie the adaptation to acid stress, such as the increased cellular buffering capacity of octanoic acid-grown cells due to their lower intracellular volume (20), the more favorable plasma membrane lipid composition (1), and the possible induction of the active efflux of the anion (26).  相似文献   

15.
16.
The presence of a mannitol cycle in fungi has been subject to discussion for many years. Recent studies have found no evidence for the presence of this cycle and its putative role in regenerating NADPH. However, all enzymes of the cycle could be measured in cultures of Aspergillus niger. In this study we have analyzed the localization of two enzymes from the pathway, mannitol dehydrogenase and mannitol-1-phosphate dehydrogenase, and the expression of their encoding genes in nonsporulating and sporulating cultures of A. niger. Northern analysis demonstrated that mpdA was expressed in both sporulating and nonsporulating mycelia, while expression of mtdA was expressed only in sporulating mycelium. More detailed studies using green fluorescent protein and dTomato fused to the promoters of mtdA and mpdA, respectively, demonstrated that expression of mpdA occurs in vegetative hyphae while mtdA expression occurs in conidiospores. Activity assays for MtdA and MpdA confirmed the expression data, indicating that streaming of these proteins is not likely to occur. These results confirm the absence of the putative mannitol cycle in A. niger as two of the enzymes of the cycle are not present in the same part of A. niger colonies. The results also demonstrate the existence of spore-specific genes and enzymes in A. niger.Mannitol has been described as one of the main compatible solutes in fungi (20) and may play a role as a storage carbon source (3) or a protectant against a variety of stresses (10, 16, 20, 22). Mannitol metabolism in fungi has been the subject of study for decades. It was proposed to exist in the form of a cyclic pathway, the mannitol cycle (9). This cycle consists of four steps enabling the conversion of fructose into mannitol and back to fructose (Fig. 1). The main role proposed for this cycle was regenerating NADPH (9, 10). Subsequently, many studies have questioned the existence of a mannitol cycle (reviewed in reference 20), and it has been shown that a mannitol cycle is not involved in NADPH regeneration in Stagonospora nodorum (19), Aspergillus niger (16), and Alternaria alternata (21). However, all enzymes of the cycle were detected in both sporulating and nonsporulating mycelia in A. niger (16), suggesting that a cycle could operate in this fungus. Fungi are able to use mannitol as a sole carbon source but do so in various ways (7).Open in a separate windowFig. 1.Putative mannitol cycle in fungi as proposed by Hult and Gatenbeck (9). HXK, hexokinase (EC 2.7.1.1); MTD, mannitol dehydrogenase (EC 1.1.1.138); MPD, mannitol-1-phosphate dehydrogenase (EC 1.1.1.17); MPP, mannitol-1-phosphate phosphatase (EC 3.1.3.22).d-Mannitol plays an important role in germination of Aspergillus conidia. In A. niger (23) and Aspergillus oryzae (8), mannitol accumulates in conidiospores and is utilized during the initial stages of germination. Production of mannitol appears to be largely dependent on mannitol-1-phosphate dehydrogenase (MPD) while mannitol dehydrogenase (MTD) contributes to a lesser extent (16, 19, 20).In this study we demonstrate that MTD and MPD as well as the expression of the corresponding genes (mtdA and mpdA) are spatially separated in colonies of A. niger. This demonstrates that a mannitol cycle does not exist in this fungus and shows that spores express specific genes that are involved in germination.  相似文献   

17.
Hyperhomocysteinemia has long been associated with atherosclerosis and thrombosis and is an independent risk factor for cardiovascular disease. Its causes include both genetic and environmental factors. Although homocysteine is produced in every cell as an intermediate of the methionine cycle, the liver contributes the major portion found in circulation, and fatty liver is a common finding in homocystinuric patients. To understand the spectrum of proteins and associated pathways affected by hyperhomocysteinemia, we analyzed the mouse liver proteome of gene-induced (cystathionine β-synthase (CBS)) and diet-induced (high methionine) hyperhomocysteinemic mice using two-dimensional difference gel electrophoresis and Ingenuity Pathway Analysis. Nine proteins were identified whose expression was significantly changed by 2-fold (p ≤ 0.05) as a result of genotype, 27 proteins were changed as a result of diet, and 14 proteins were changed in response to genotype and diet. Importantly, three enzymes of the methionine cycle were up-regulated. S-Adenosylhomocysteine hydrolase increased in response to genotype and/or diet, whereas glycine N-methyltransferase and betaine-homocysteine methyltransferase only increased in response to diet. The antioxidant proteins peroxiredoxins 1 and 2 increased in wild-type mice fed the high methionine diet but not in the CBS mutants, suggesting a dysregulation in the antioxidant capacity of those animals. Furthermore, thioredoxin 1 decreased in both wild-type and CBS mutants on the diet but not in the mutants fed a control diet. Several urea cycle proteins increased in both diet groups; however, arginase 1 decreased in the CBS+/− mice fed the control diet. Pathway analysis identified the retinoid X receptor signaling pathway as the top ranked network associated with the CBS+/− genotype, whereas xenobiotic metabolism and the NRF2-mediated oxidative stress response were associated with the high methionine diet. Our results show that hyperhomocysteinemia, whether caused by a genetic mutation or diet, alters the abundance of several liver proteins involved in homocysteine/methionine metabolism, the urea cycle, and antioxidant defense.Homocysteine (Hcy)1 is a thiol-containing amino acid that is produced in every cell of the body as an intermediate of the methionine cycle (Fig. 1, Reactions 1–5) (1). Once formed, the catabolism of homocysteine occurs via three enzymatic pathways. 1) Hcy is remethylated back to methionine using vitamin B12-dependent methionine synthase (Fig. 1, Reaction 4) and/or 2) betaine-homocysteine methyltransferase (BHMT) (Fig. 1, Reaction 5), and 3) Hcy is converted to cysteine via the transsulfuration pathway using CBS and γ-cystathionase (Fig. 1, Reactions 6 and 7). Under normal conditions ∼40–50% of the Hcy that is produced in the liver is remethylated, ∼40–50% is converted to cysteine, and a small amount is exported (13). However, when Hcy production is increased (i.e. increased dietary methionine/protein intake) or when Hcy catabolism is decreased (i.e. CBS deficiency or B vitamin deficiencies), excess Hcy is exported into the extracellular space, resulting in hyperhomocysteinemia (15).Open in a separate windowFig. 1.Homocysteine metabolism in liver and kidney. In classical homocystinuria, the initial enzyme of the transsulfuration pathway, CBS (Reaction 6), is deficient. MTHF, methylenetetrahydrofolate; THF, tetrahydrofolate; DHF, dihydrofolate; MeCbl, methylcobalamin; DMG, dimethylglycine; PLP, pyridoxal 5′-phosphate.Homocystinuria was first described in the 1960s by Carson et al. (6): they observed 10 pediatric patients with severely elevated levels of Hcy in the urine and hypermethioninemia. Normal concentrations of plasma total homocysteine (tHcy) range from 5 to 12 μm (7); however, in homocystinuria, tHcy levels can exceed 100 μm. Homocystinuric patients present with mental retardation, abnormal bone growth, fine hair, malar flush, and dislocation of the lens of the eye, and most die from premature cardiovascular disease (6, 8). Autopsy findings indicate widespread thromboembolism, arteriosclerosis, and fatty livers (6, 8). Mudd et al. (9, 10) identified the cause of homocystinuria as a defect in the enzyme cystathionine β-synthase. A recent study of newborn infants in Denmark estimated the birth prevalence for CBS heterozygosity to be about 1:20,000 (11).Plasma tHcy concentrations are also directly correlated with dietary methionine/protein intake (12, 13). Guttormsen et al. (13) demonstrated that a protein-rich meal affected tHcy for at least 8–24 h. When normal subjects were fed a low protein-containing breakfast (12–15 g), plasma methionine levels increased slightly after 2 h (22.5–27.5 μm), but tHcy levels did not change significantly. However, when these same subjects were fed a high protein meal (52 g), plasma methionine levels peaked after 4 h (38 μm), and tHcy rose steadily until a maximum level was reached 8 h postmeal (7.6 versus 8.5 μm) (13). Thus, the following questions can be raised. How does the hepatic proteome respond to a hyperhomocysteinemic diet, and are the changes that accompany such a diet the same as or different from those that may be observed in gene-induced hyperhomocysteinemia?Because hyperhomocysteinemia is a strong independent risk factor for cardiovascular, cerebrovascular, and peripheral vascular disease, most of the current research has focused on the mechanisms involved in Hcy-induced endothelial dysfunction (1424). The results of those studies have concluded that Hcy induces intracellular oxidative stress by generating ROS, which in turn lead to decreased bioavailable nitric oxide (NO), altered gene expression, increased endoplasmic reticulum stress, and activation of cholesterol biosynthesis. Also, several studies have examined the association between hyperhomocysteinemia and alcoholic liver disease, but few have looked at the effect of Hcy on the non-alcoholic liver even though fatty liver is a constant finding in homocystinuria (6, 8), and the liver is the major source of circulating Hcy (4, 5, 10). We hypothesize that 1) the liver proteome will respond to hyperhomocysteinemia by altering the expression of proteins involved in methionine/homocysteine metabolism and antioxidant defense and that 2) the set of proteins that are expressed when hyperhomocysteinemia is induced by CBS deficiency will differ from those expressed as a result of a high methionine diet. In the present study, we use a well established mouse model of CBS deficiency to study the early changes in the liver proteome that accompany hyperhomocysteinemia (25).  相似文献   

18.
Drug target identification is a critical step toward understanding the mechanism of action of a drug, which can help one improve the drug''s current therapeutic regime and expand the drug''s therapeutic potential. However, current in vitro affinity-chromatography-based and in vivo activity-based protein profiling approaches generally face difficulties in discriminating specific drug targets from nonspecific ones. Here we describe a novel approach combining isobaric tags for relative and absolute quantitation with clickable activity-based protein profiling to specifically and comprehensively identify the protein targets of andrographolide (Andro), a natural product with known anti-inflammation and anti-cancer effects, in live cancer cells. We identified a spectrum of specific targets of Andro, which furthered our understanding of the mechanism of action of the drug. Our findings, validated through cell migration and invasion assays, showed that Andro has a potential novel application as a tumor metastasis inhibitor. Moreover, we have unveiled the target binding mechanism of Andro with a combination of drug analog synthesis, protein engineering, and mass-spectrometry-based approaches and determined the drug-binding sites of two protein targets, NF-κB and actin.As most drugs exert pharmacological effects by interacting with their target proteins, the identification of these target proteins is a critical step in unraveling the mechanisms of drug action. It is also imperative for our understanding of the pharmacodynamics of a known drug, suggesting potentially unrevealed actions and thus refining future clinical applications of the substance. Traditional approaches used to identify protein targets of a drug typically utilize immobilized drug affinity chromatography coupled with mass spectrometry (MS)1 (1, 2). These methods can be applied to cell lysates, but not in an in vivo setting, because of the requirement of a solid support. In vitro target profiling might not accurately reflect the drug''s actions in the in vivo physiological environment. To overcome this limitation, several groups have used activity-based protein profiling (ABPP) combined with bio-orthogonal click chemistry to identify drug targets both in vitro and in vivo (supplemental Fig. S1) (315). ABPP probes exert their functions via covalent reactions with the target proteins or photoaffinity-based labeling via the incorporation of photoreactive groups. With the increasing sensitivity of modern MS platforms, low-abundance protein targets can be successfully identified. Although both conventional affinity chromatography and recent ABPP-based methods allow us to detect a set of candidate protein targets for a drug, it remains difficult to discriminate specific interactions from nonspecific ones. Thus, more time and effort are needed for subsequent validation because of the presence of a large number of nonspecific binders. Therefore, there is an urgent need to develop comprehensive unbiased methods for specific target identification. Quantitative proteomics has been used to profile enriched kinases using cell-lysate-based kinobead pull-down. However, these types of experiments are mainly suitable for studying kinase inhibitors (16). Recently, proteomics methods based on stable isotope labeling of amino acids in cell culture (SILAC) have been applied to determine the specific binders of small molecules or proteins with certain post-translational modifications (1719). These studies have shed light on how quantitative proteomics can improve the specificity of target protein identification. Nevertheless, because of the inherent limitations of SILAC, the complete incorporation of isotopic amino acids via such an approach takes a long time. Furthermore, it is also extremely difficult to apply the SILAC approach to tissue and body fluid samples, which are of particular relevance to biomedical research.Here we introduce a clickable activity-based protein profiling (ICABPP) approach based on the use of isobaric tags for relative and absolute quantitation (iTRAQ) for the unbiased specific and comprehensive identification of target proteins in live cells. iTRAQ is a stable isotope labeling approach for multiplexed quantitative proteome profiling (20). An overview of the technique is illustrated in Fig. 1A. In this assay, cells are first incubated with a clickable probe or with DMSO, which serves as a negative control. After the probe permeates the cell, and covalently binds to its dedicated in situ targets, the washed cells are lysed, clicked with biotin-N3 tag, and enriched through avidin pull-down in parallel. The beads are washed thoroughly, and the bond proteins are directly digested on the beads with trypsin. The resulting peptides are labeled with their respective iTRAQ reagents, pooled together for further identification and quantification via LC-MS/MS. This technique enabled us to discriminate specific protein targets from nonspecific, and endogenously biotinylated proteins. Biological replicates of probe- or DMSO-treated samples are included to overcome experimental variations. As shown in Fig. 1B, nonspecific binding proteins'' iTRAQ reporters have equal or similar intensities, whereas specific target proteins enriched by the probe show highly differential intensities relative to the DMSO-treated control samples (as illustrated by the significantly higher reporter intensities of 116 and 117 versus 113 and 114 shown in Fig. 1B). The multiplexing nature of the iTRAQ-based chemical proteomics method allows replicated enrichments to be compared within a single LC-MS/MS analysis, thereby increasing the accuracy of specific target identifications, and minimizing experimental errors.Open in a separate windowFig. 1.Identifying specific drug targets using ICABPP approach in live cells. A, live cells were treated with DMSO, and clickable probe in duplicate. Cells were lysed and tagged with biotin-N3 using click chemistry in parallel. The biotin-bearing target proteins were enriched through avidin pull-down, and directly digested on beads. The resulting peptides of the two biological replicates of control pulled-down samples were labeled with 113 and 114, respectively, and two probe pulled-down samples were labeled with 116 and 117, respectively. The labeled peptides were combined in order to be identified, and quantified via LC-MS/MS. B, for the nonspecific targets, the iTRAQ reporters showed similar intensities, whereas for the specific targets, the reporters showed highly differential intensities.In this context, the ICABPP approach was applied to identify protein targets of andrographolide (Andro) (Fig. 2), a natural product with known anti-inflammation, and anti-cancer effects (2125), in live cancer cells. A spectrum of 75 potential Andro targets was identified with high confidence, which suggested that Andro may exert anti-cancer effects by acting on multiple targets to interfere with several cellular signaling pathways. Two targets, NF-κB and β-actin, were validated by in vitro binding assay, and direct binding site mapping. Furthermore, our data revealed a novel mechanism of Andro in suppressing tumor metastasis.Open in a separate windowFig. 2.Chemical structures of Andro, reduced Andro analog RA, and Andro-based clickable ABPP probes P1 and P2.  相似文献   

19.
Pelvic lymph node dissection (PLND) is a standard component of radical cystectomy (RC) for bladder cancer. The optimal anatomic PLND template remains undefined. An extended PLND template can potentially improve survival through the eradication of micrometastatic disease and improved pathologic staging. However, this benefit could be compromised by a potential increase in perioperative complications and cost. Two randomized controlled clinical trials that will clarify this question are ongoing. Many important retrospective studies have provided insights into the optimal PLND extent. Here the authors review the key evidence that informs how urologists may tailor the PLND template during RC depending on patient and tumor characteristics.Key words: Pelvic lymph node dissection, Radical cystectomy, Template, Micrometastatic diseaseAs described by Whitmore and Marshall in 1962,1 the original pelvic lymph node dissection (PLND) template during radical cystectomy (RC) included nodal/lymphatic tissue bounded by the external iliac artery, the distal ureter, and Cooper’s ligament. The nodal tissue enveloping the external iliac vein and occupying the obturator fossa nodes was also removed, followed by removal of the lymphatic tissue surrounding the internal iliac vessels and posterior pelvis, including presacral nodes.2In the ensuing decades, urologists at different centers have modified the original template. However, to date, no level I evidence exists to address the value of one dissection template over another; published guidelines, accordingly, are vague.3 Several names have been given to different templates and nodal regions. The limited template includes perivesical nodes and lymphatic tissue in the obturator fossa, limited laterally by the external iliac vein and medially by the obturator nerve.46 A larger limited lymphadenectomy template (referred to as standard by some) is bounded distally by the circumflex iliac vein and Cloquet’s node, laterally by the genitofemoral nerve, medially by the bladder and internal iliac vessels, posteriorly by the obturator fossa, and proximally by the bifurcation of (or distal aspect of) the common iliac artery.58Extended PLND has been variously defined to include a more cephalad proximal boundary located at or above the aortic bifurcation and may also include presacral and presciatic nodes.7,9,10 The extended template bounded proximally by the crossing of the ureter and external iliac artery6,11 is referred to as standard/extended in this review. Still others define the extended template to extend as proximally as the origin of the inferior mesenteric artery and also include lymphatic tissues in the triangle of Marcille, which involves medial retraction of the external iliac artery to reveal the space medial to the psoas major; dissection of this space reveals the proximal aspect of the obturator nerve as it emanates from the psoas.4,11,12An alternative nomenclature system for PLND defines three levels of operative intervention: levels I, II, and III (Figure 1).13 The level I template is similar to a standard template—the proximal extent of this is the common iliac bifurcation. Level II nodes are lateral to the common iliac artery, include presacral nodes, and extend proximally to the aortic bifurcation. Level III nodes are located between the ureters and great vessels and extend proximally to the inferior mesenteric artery. Investigators have also referred to super-extended PLND, which is the template that would include all level I, II, and III nodes (Figure 2).2,11,14Open in a separate windowFigure 1Anatomic pelvic lymph node dissection levels: I, II, and III. Percentages of patients with lymph node metastases at each level are shown (from a series of 290 patients who had super-extended PLND including levels I, II, and III). Adapted from Leissner J et al.13Open in a separate windowFigure 2(A) Super-extended pelvic lymph node dissection template (levels I, II, and III). (B) Standard/extended template (level I plus level II up to crossing of ureter and external iliac artery). Adapted from Zehnder P et al.11  相似文献   

20.
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