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.
Aldoxime dehydratase (Oxd) is a novel hemeprotein that catalyzes the dehydration reaction of aldoxime to produce nitrile. In this study, we studied the spectroscopic and substrate binding properties of two Oxds, OxdB from Bacillus sp. strain OxB-1 and OxdRE from Rhodococcus sp. N-771, that show different quaternary structures and relatively low amino acid sequence identity. Electronic absorption and resonance Raman spectroscopy revealed that ferric OxdRE contained a six-coordinate low-spin heme, while ferric OxdB contained a six-coordinate high-spin heme. Both ferrous OxdRE and OxdB included a five-coordinate high-spin heme to which the substrate was bound via its nitrogen atom for the reaction to occur. Although the ferric Oxds were inactive for catalysis, the substrate was bound to the ferric heme via its oxygen atom in both OxdB and OxdRE. Electronic paramagnetic resonance (EPR) and rapid scanning spectroscopy revealed that the flexibility of the heme pocket was different between OxdB and OxdRE, which might affect their substrate specificity.  相似文献   

2.
To the best of our knowledge, this is the first report on the structure of product-inhibited mammalian peroxidase. Lactoperoxidase is a heme containing an enzyme that catalyzes the inactivation of a wide range of microorganisms. In the presence of hydrogen peroxide, it preferentially converts thiocyanate ion into a toxic hypothiocyanate ion. Samples of bovine lactoperoxidase containing thiocyanate (SCN) and hypothiocyanate (OSCN) ions were purified and crystallized. The structure was determined at 2.3-Å resolution and refined to Rcryst and Rfree factors of 0.184 and 0.221, respectively. The determination of structure revealed the presence of an OSCN ion at the distal heme cavity. The presence of OSCN ions in crystal samples was also confirmed by chemical and spectroscopic analysis. The OSCN ion interacts with the heme iron, Gln-105 Nɛ1, His-109 Nɛ2, and a water molecule W96. The sulfur atom of the OSCN ion forms a hypervalent bond with a nitrogen atom of the pyrrole ring D of the heme moiety at an S–N distance of 2.8 Å. The heme group is covalently bound to the protein through two ester linkages involving carboxylic groups of Glu-258 and Asp-108 and the modified methyl groups of pyrrole rings A and C, respectively. The heme moiety is significantly distorted from planarity, whereas pyrrole rings A, B, C, and D are essentially planar. The iron atom is displaced by ≈0.2 Å from the plane of the heme group toward the proximal site. The substrate channel resembles a long tunnel whose inner walls contain predominantly aromatic residues such as Phe-113, Phe-239, Phe-254, Phe-380, Phe-381, Phe-422, and Pro-424. A phosphorylated Ser-198 was evident at the surface, in the proximity of the calcium-binding channel.  相似文献   

3.
In the recently identified cholesterol catabolic pathway of Mycobacterium tuberculosis, 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate hydrolase (HsaD) is proposed to catalyze the hydrolysis of a carbon-carbon bond in 4,5–9,10-diseco-3-hydroxy-5,9,17-tri-oxoandrosta-1(10),2-diene-4-oic acid (DSHA), the cholesterol meta-cleavage product (MCP) and has been implicated in the intracellular survival of the pathogen. Herein, purified HsaD demonstrated 4–33 times higher specificity for DSHA (kcat/Km = 3.3 ± 0.3 × 104 m−1 s−1) than for the biphenyl MCP 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) and the synthetic analogue 8-(2-chlorophenyl)-2-hydroxy-5-methyl-6-oxoocta-2,4-dienoic acid (HOPODA), respectively. The S114A variant of HsaD, in which the active site serine was substituted with alanine, was catalytically impaired and bound DSHA with a Kd of 51 ± 2 μm. The S114A·DSHA species absorbed maximally at 456 nm, 60 nm red-shifted versus the DSHA enolate. Crystal structures of the variant in complex with HOPDA, HOPODA, or DSHA to 1.8–1.9 Åindicate that this shift is due to the enzyme-induced strain of the enolate. These data indicate that the catalytic serine catalyzes tautomerization. A second role for this residue is suggested by a solvent molecule whose position in all structures is consistent with its activation by the serine for the nucleophilic attack of the substrate. Finally, the α-helical lid covering the active site displayed a ligand-dependent conformational change involving differences in side chain carbon positions of up to 6.7 Å, supporting a two-conformation enzymatic mechanism. Overall, these results provide novel insights into the determinants of specificity in a mycobacterial cholesterol-degrading enzyme as well as into the mechanism of MCP hydrolases.  相似文献   

4.
The crystal structure of the complex of lactoperoxidase (LPO) with its physiological substrate thiocyanate (SCN) has been determined at 2.4Å resolution. It revealed that the SCN ion is bound to LPO in the distal heme cavity. The observed orientation of the SCN ion shows that the sulfur atom is closer to the heme iron than the nitrogen atom. The nitrogen atom of SCN forms a hydrogen bond with a water (Wat) molecule at position 6′. This water molecule is stabilized by two hydrogen bonds with Gln423 Nε2 and Phe422 oxygen. In contrast, the placement of the SCN ion in the structure of myeloperoxidase (MPO) occurs with an opposite orientation, in which the nitrogen atom is closer to the heme iron than the sulfur atom. The site corresponding to the positions of Gln423, Phe422 oxygen, and Wat6′ in LPO is occupied primarily by the side chain of Phe407 in MPO due to an entirely different conformation of the loop corresponding to the segment Arg418–Phe431 of LPO. This arrangement in MPO does not favor a similar orientation of the SCN ion. The orientation of the catalytic product OSCN as reported in the structure of LPO·OSCN is similar to the orientation of SCN in the structure of LPO·SCN. Similarly, in the structure of LPO·SCN·CN, in which CN binds at Wat1, the position and orientation of the SCN ion are also identical to that observed in the structure of LPO·SCN.Lactoperoxidase (LPO4; EC 1.11.1.7) is a Fe3+ heme enzyme that belongs to the mammalian peroxidase family (1). The family of mammalian peroxidases comprises lactoperoxidase (2), eosinophil peroxidase (3), thyroid peroxidase (4), and myeloperoxidase (MPO) (5). LPO, eosinophil peroxidase, and MPO are responsible for antimicrobial function and innate immune responses (68), whereas thyroid peroxidase plays a key role in thyroid hormone biosynthesis (9). These peroxidases are different from plant and fungal peroxidases because unlike plant and fungal enzymes, the prosthetic heme group in mammalian peroxidases is covalently linked to the protein (10). There are also several striking structural and functional differences among the mammalian peroxidases (11). The heme group in MPO is attached to the protein via three covalent linkages (12), whereas LPO (12, 13), eosinophil peroxidase (12), and thyroid peroxidase (12) contain only two ester linkages. These covalent and various non-covalent linkages contribute differentially to the high stability of the heme core as well as for the peculiar values of their redox potentials (2, 14). Furthermore, MPO consists of two disulfide-linked protein chains, whereas LPO, eosinophil peroxidase, and thyroid peroxidase are single chain proteins, although their chain lengths differ greatly. In addition, their sequences contain several critical amino acid differences that may also contribute to the variations in the stereochemical environments of the substrate-binding sites. As a consequence of these differences, the mammalian enzymes oxidize various inorganic ions such as SCN, Br, Cl, and I with differing specificities and potencies. Biochemical studies have shown that LPO catalyzes preferentially the conversion of SCN to OSCN (15, 16), whereas MPO uses halides (17, 18) with a preference for chloride ion as the substrate. The preferences of eosinophil peroxidase and thyroid peroxidase are bromide and iodide, respectively. However, the stereochemical basis of the reported preferences for the substrates by mammalian heme peroxidases is still unclear. So far, the structures of only two mammalian enzymes, MPO and LPO, have been determined (12, 13). It is of considerable importance to identify the structural parameters that are responsible for the subtle specificities. In the present work, we have attempted to address this question through the new crystal structures of LPO complexes with SCN ions using goat, bovine, and buffalo lactoperoxidases. Because the overall structures of complexes of SCN with LPO from all three species were found to be identical, the structure of the complex of buffalo LPO with SCN and the ternary complex with SCN and CN will be discussed here, and buffalo LPO will be termed hereafter as LPO. To highlight the factors pertaining to binding specificity of SCN, a comparison of the structures of LPO·SCN and MPO·SCN has also been made, revealing many valuable differences pertaining to the observed orientations of the common substrate, SCN ion, when bound at the substrate-binding site in the distal heme cavity of the two structures. The structures of LPO·SCN and MPO·SCN clearly show that the bound SCN ions are present in the distal heme cavity of two enzymes with opposite orientations. In the structure of LPO·SCN, the sulfur atom is closer to the heme iron than the nitrogen atom, whereas in that of MPO·SCN, the nitrogen atom is closer to the heme iron than the sulfur atom. As a result of this, the interactions of the SCN ion in the distal site of two proteins differ drastically. Gln423, a conserved water (Wat) molecule at position 6′, and a well aligned carbonyl oxygen of Phe422 in the proximity of the substrate-binding site in LPO against a protruding Phe407 in MPO seem to play the key roles in inducing the observed orientations of SCN ions in LPO and MPO. The structure of LPO·SCN has also been compared with the structure of its ternary complex with SCN and CN ions.  相似文献   

5.
RlmM (YgdE) catalyzes the S-adenosyl methionine (AdoMet)-dependent 2′O methylation of C2498 in 23S ribosomal RNA (rRNA) of Escherichia coli. Previous experiments have shown that RlmM is active on 23S rRNA from an RlmM knockout strain but not on mature 50S subunits from the same strain. Here, we demonstrate RlmM methyltransferase (MTase) activity on in vitro transcribed 23S rRNA and its domain V. We have solved crystal structures of E. coli RlmM at 1.9 Å resolution and of an RlmM–AdoMet complex at 2.6 Å resolution. RlmM consists of an N-terminal THUMP domain and a C-terminal catalytic Rossmann-like fold MTase domain in a novel arrangement. The catalytic domain of RlmM is closely related to YiiB, TlyA and fibrillarins, with the second K of the catalytic tetrad KDKE shifted by two residues at the C-terminal end of a beta strand compared with most 2′O MTases. The AdoMet-binding site is open and shallow, suggesting that RNA substrate binding may be required to form a conformation needed for catalysis. A continuous surface of conserved positive charge indicates that RlmM uses one side of the two domains and the inter-domain linker to recognize its RNA substrate.  相似文献   

6.
Lipoate-protein ligase A (LplA) catalyzes the attachment of lipoic acid to lipoate-dependent enzymes by a two-step reaction: first the lipoate adenylation reaction and, second, the lipoate transfer reaction. We previously determined the crystal structure of Escherichia coli LplA in its unliganded form and a binary complex with lipoic acid (Fujiwara, K., Toma, S., Okamura-Ikeda, K., Motokawa, Y., Nakagawa, A., and Taniguchi, H. (2005) J Biol. Chem. 280, 33645–33651). Here, we report two new LplA structures, LplA·lipoyl-5′-AMP and LplA·octyl-5′-AMP·apoH-protein complexes, which represent the post-lipoate adenylation intermediate state and the pre-lipoate transfer intermediate state, respectively. These structures demonstrate three large scale conformational changes upon completion of the lipoate adenylation reaction: movements of the adenylate-binding and lipoate-binding loops to maintain the lipoyl-5′-AMP reaction intermediate and rotation of the C-terminal domain by about 180°. These changes are prerequisites for LplA to accommodate apoprotein for the second reaction. The Lys133 residue plays essential roles in both lipoate adenylation and lipoate transfer reactions. Based on structural and kinetic data, we propose a reaction mechanism driven by conformational changes.  相似文献   

7.
5-HT3A receptors select among permeant ions based on size and charge. The membrane-associated (MA) helix lines the portals into the channel’s cytoplasmic vestibule in the 4-Å resolution structure of the homologous acetylcholine receptor. 5-HT3A MA helix residues are important determinants of single-channel conductance. It is unknown whether the portals into the cytoplasmic vestibule also determine the size selectivity of permeant ions. We sought to determine whether the portals form the size selectivity filter. Recently, we showed that channels functioned when the entire 5-HT3A M3–M4 loop was replaced by the heptapeptide M3–M4 loop sequence from GLIC, a bacterial Cys-loop neurotransmitter gated ion channel homologue from Gloebacter violaceus. We used homomeric 5-HT3A receptors with either a wild-type (WT) M3–M4 loop or the chimeric heptapeptide (5-HT3A–glvM3M4) loop, i.e., with or without portals. In Na+-containing buffer, the WT receptor current–voltage relationship was inwardly rectifying. In contrast, the 5-HT3A–glvM3M4 construct had a negative slope conductance region at voltages less than −80 mV. Glutamine substitution for the heptapeptide M3–M4 loop arginine eliminated the negative slope conductance region. We measured the relative permeabilities and conductances of a series of inorganic and organic cations ranging from 0.9 to 4.5 Å in radius (Li+, Na+, ammonium, methylammonium, ethanolammonium, 2-methylethanolammonium, dimethylammonium, diethanolammonium, tetramethylammonium, choline, tris [hydroxymethyl] aminomethane, and N-methyl-d-glucamine). Both constructs had measurable conductances with Li+, ammonium, and methylammonium (size range of 0.9–1.8-Å radius). Many of the organic cations >2.4 Å acted as competitive antagonists complicating measurement of conductance ratios. Analysis of the permeability ratios by excluded volume theory indicates that the minimal pore radius for 5-HT3A and 5-HT3–glvM3M4 receptors was similar, ∼5 Å. We infer that the 5-HT3A size selectivity filter is located in the transmembrane channel and not in the portals into the cytoplasmic vestibule. Thus, the determinants of size selectivity and conductance are located in physically distinct regions of the channel protein.  相似文献   

8.
p-Hydroxyphenylacetate (HPA) 3-hydroxylase is a two-component flavoprotein monooxygenase that catalyzes the hydroxylation of p-hydroxyphenylacetate to form 3,4-dihydroxyphenylacetate. Based on structures of the oxygenase component (C2), both His-120 and Ser-146 are located ∼2.8 Å from the hydroxyl group of HPA. The variants H120N, H120Q, H120Y, H120D, and H120E can form C4a-hydroperoxy-FMN (a reactive intermediate necessary for hydroxylation) but cannot hydroxylate HPA. The impairment of H120N is not due to substrate binding because the variant can still bind HPA. In contrast, the H120K variant catalyzes hydroxylation with efficiency comparable with that of the wild-type enzyme; the hydroxylation rate constant for H120K is 5.7 ± 0.6 s−1, and the product conversion ratio is 75%, compared with values of 16 s−1 and 90% for the wild-type enzyme. H120R can also catalyze hydroxylation, suggesting that a positive charge on residue 120 can substitute for the hydroxylation function of His-120. Because the hydroxylation reaction of wild-type C2 is pH-independent between pH 6 and 10, the protonation status of key components required for hydroxylation likely remains unchanged in this pH range. His-120 may be positively charged for selective binding to the phenolate form of HPA, i.e. to form the Hisδ+·HPAδ− complex, which in turn promotes oxygen atom transfer via an electrophilic aromatic substitution mechanism. Analysis of Ser-146 variants revealed that this residue is necessary for but not directly engaged in hydroxylation. Product formation in S146A is pH-independent and constant at ∼70% over a pH range of 6–10, whereas product formation for S146C decreased from ∼65% at pH 6.0 to 27% at pH 10.0. These data indicate that the ionization of Cys-146 in the S146C variant has an adverse effect on hydroxylation, possibly by perturbing formation of the Hisδ+·HPAδ− complex needed for hydroxylation.  相似文献   

9.
The first structural and biophysical data on the folate biosynthesis pathway enzyme and drug target, 6-hydroxymethyl-7,8-dihydropterin pyrophosphokinase (SaHPPK), from the pathogen Staphylococcus aureus is presented. HPPK is the second essential enzyme in the pathway catalysing the pyrophosphoryl transfer from cofactor (ATP) to the substrate (6-hydroxymethyl-7,8-dihydropterin, HMDP). In-silico screening identified 8-mercaptoguanine which was shown to bind with an equilibrium dissociation constant, Kd, of ∼13 µM as measured by isothermal titration calorimetry (ITC) and surface plasmon resonance (SPR). An IC50 of ∼41 µM was determined by means of a luminescent kinase assay. In contrast to the biological substrate, the inhibitor has no requirement for magnesium or the ATP cofactor for competitive binding to the substrate site. The 1.65 Å resolution crystal structure of the inhibited complex showed that it binds in the pterin site and shares many of the key intermolecular interactions of the substrate. Chemical shift and 15N heteronuclear NMR measurements reveal that the fast motion of the pterin-binding loop (L2) is partially dampened in the SaHPPK/HMDP/α,β-methylene adenosine 5′-triphosphate (AMPCPP) ternary complex, but the ATP loop (L3) remains mobile on the µs-ms timescale. In contrast, for the SaHPPK/8-mercaptoguanine/AMPCPP ternary complex, the loop L2 becomes rigid on the fast timescale and the L3 loop also becomes more ordered – an observation that correlates with the large entropic penalty associated with inhibitor binding as revealed by ITC. NMR data, including 15N-1H residual dipolar coupling measurements, indicate that the sulfur atom in the inhibitor is important for stabilizing and restricting important motions of the L2 and L3 catalytic loops in the inhibited ternary complex. This work describes a comprehensive analysis of a new HPPK inhibitor, and may provide a foundation for the development of novel antimicrobials targeting the folate biosynthetic pathway.  相似文献   

10.
The DNA photosensitisers m-iodo Hoechst and m-iodo, p-methoxy Hoechst have been co-crystallised with the oligonucleotide d(CGCGAATTCGCG)2 and their crystal structures determined. The crystals were then subjected to slow dehydration, which reduced their solvent contents from 40 (normal) to 30 (partially dehydrated) and then 20% (fully dehydrated) and caused a reduction in cell volume from 68 000 to 60 000 then 51 000 Å3. The dehydration resulted in a dramatic enhancement of diffraction resolution from ~2.6 to beyond 1.5 Å. Crystal structures have also been determined for the partially and fully dehydrated states. The fully dehydrated crystals consist of an infinite polymeric network, in which neighbouring dodecamer duplexes are crosslinked through phosphate oxygens via direct bonding to bridging magnesium cations. This unique three-dimensional structure for DNA is described in detail in the following companion paper. The present paper details evidence from the sequence of crystal structures that the DNA is able to breathe locally, allowing the ligand to leave the minor groove, re-orient in the surrounding solvent medium and then re-enter the groove in a different orientation and location. The rearrangement of the minor groove binding ligands during the dehydration process mimics the binding behaviour of these ligands in solution and in vivo. We also present details of the DNA–ligand interactions that are consistent with a hydrogen atom abstraction mechanism for photocleavage of DNA.  相似文献   

11.
Undecaprenyl pyrophosphate synthase (UPPs) is an essential enzyme in a key bacterial cell wall synthesis pathway. It catalyzes the consecutive condensations of isopentenyl pyrophosphate (IPP) groups on to a trans-farnesyl pyrophosphate (FPP) to produce a C55 isoprenoid, undecaprenyl pyrophosphate (UPP). Here we report the discovery and co-crystal structures of a drug-like UPPs inhibitor in complex with Streptococcus pneumoniae UPPs, with and without substrate FPP, at resolutions of 2.2 and 2.1 Å, respectively. The UPPs inhibitor has a low molecular weight (355 Da), but displays potent inhibition of UPP synthesis in vitro (IC50 50 nM) that translates into excellent whole cell antimicrobial activity against pathogenic strains of Streptococcal species (MIC90 0.4 µg mL−1). Interestingly, the inhibitor does not compete with the substrates but rather binds at a site adjacent to the FPP binding site and interacts with the tail of the substrate. Based on the structures, an allosteric inhibition mechanism of UPPs is proposed for this inhibitor. This inhibition mechanism is supported by biochemical and biophysical experiments, and provides a basis for the development of novel antibiotics targeting Streptococcus pneumoniae.  相似文献   

12.

Background

The essential purine salvage pathway of Trypanosoma brucei bears interesting catalytic enzymes for chemotherapeutic intervention of Human African Trypanosomiasis. Unlike mammalian cells, trypanosomes lack de novo purine synthesis and completely rely on salvage from their hosts. One of the key enzymes is adenosine kinase which catalyzes the phosphorylation of ingested adenosine to form adenosine monophosphate (AMP) utilizing adenosine triphosphate (ATP) as the preferred phosphoryl donor.

Methods and Findings

Here, we present the first structures of Trypanosoma brucei rhodesiense adenosine kinase (TbrAK): the structure of TbrAK in complex with the bisubstrate inhibitor P1,P5-di(adenosine-5′)-pentaphosphate (AP5A) at 1.55 Å, and TbrAK complexed with the recently discovered activator 4-[5-(4-phenoxyphenyl)-2H-pyrazol-3-yl]morpholine (compound 1) at 2.8 Å resolution.

Conclusions

The structural details and their comparison give new insights into substrate and activator binding to TbrAK at the molecular level. Further structure-activity relationship analyses of a series of derivatives of compound 1 support the observed binding mode of the activator and provide a possible mechanism of action with respect to their activating effect towards TbrAK.  相似文献   

13.
The development of heme oxygenase (HO) inhibitors is critical in dissecting and understanding the HO system and for potential therapeutic applications. We have established a program to design and optimize HO inhibitors using structure-activity relationships in conjunction with X-ray crystallographic analyses. One of our previous complex crystal structures revealed a putative secondary hydrophobic binding pocket which could be exploited for a new design strategy by introducing a functional group that would fit into this potential site. To test this hypothesis and gain further insights into the structural basis of inhibitor binding, we have synthesized and characterized 1-(1H-imidazol-1-yl)-4,4-diphenyl-2-butanone (QC-308). Using a carbon monoxide (CO) formation assay on rat spleen microsomes, the compound was found to be ∼15 times more potent (IC50 = 0.27±0.07 µM) than its monophenyl analogue, which is already a potent compound in its own right (QC-65; IC50 = 4.0±1.8 µM). The crystal structure of hHO-1 with QC-308 revealed that the second phenyl group in the western region of the compound is indeed accommodated by a definitive secondary proximal hydrophobic pocket. Thus, the two phenyl moieties are each stabilized by distinct hydrophobic pockets. This “double-clamp” binding offers additional inhibitor stabilization and provides a new route for improvement of human heme oxygenase inhibitors.  相似文献   

14.
Gu W  Yang J  Lou Z  Liang L  Sun Y  Huang J  Li X  Cao Y  Meng Z  Zhang KQ 《PloS one》2011,6(1):e16262
Microbial ferulic acid decarboxylase (FADase) catalyzes the transformation of ferulic acid to 4-hydroxy-3-methoxystyrene (4-vinylguaiacol) via non-oxidative decarboxylation. Here we report the crystal structures of the Enterobacter sp. Px6-4 FADase and the enzyme in complex with substrate analogues. Our analyses revealed that FADase possessed a half-opened bottom β-barrel with the catalytic pocket located between the middle of the core β-barrel and the helical bottom. Its structure shared a high degree of similarity with members of the phenolic acid decarboxylase (PAD) superfamily. Structural analysis revealed that FADase catalyzed reactions by an “open-closed” mechanism involving a pocket of 8×8×15 Å dimension on the surface of the enzyme. The active pocket could directly contact the solvent and allow the substrate to enter when induced by substrate analogues. Site-directed mutagenesis showed that the E134A mutation decreased the enzyme activity by more than 60%, and Y21A and Y27A mutations abolished the enzyme activity completely. The combined structural and mutagenesis results suggest that during decarboxylation of ferulic acid by FADase, Trp25 and Tyr27 are required for the entering and proper orientation of the substrate while Glu134 and Asn23 participate in proton transfer.  相似文献   

15.
Wang Q  Zhang M  Ning G  Gu W  Su T  Xu M  Li B  Wang W 《PloS one》2011,6(6):e21006

Background

Recent studies have shown that adult human possess active brown adipose tissue (BAT), which might be important in controlling obesity. It is known that ß-adrenoceptor-UCP1 system regulates BAT in rodent, but its influence in adult humans remains to be shown. The present study is to determine whether BAT activity can be independently stimulated by elevated catecholamines levels in adult human, and whether it is associated with their adiposity.

Methodology/Principal Findings

We studied 14 patients with pheochromocytoma and 14 normal subjects who had performed both 18F-fluorodeoxyglucose positron emission tomography/computed tomography (18F-FDG PET/CT) and plasma total metanephrine (TMN) measurements during 2007–2010. The BAT detection rate and the mean BAT activity were significantly higher in patients with elevated TMN levels (Group A: 6/8 and 6.7±2.1 SUVmean· g/ml) than patients with normal TMN concentrations (Group B: 0/6 and 0.4±0.04 SUVmean· g/ml) and normal subjects (Group C: 0/14 and 0.4±0.03 SUVmean·g/ml). BAT activities were positively correlated with TMN levels (R = 0.83, p<0.0001) and were inversely related to body mass index (R = −0.47, p = 0.010), visceral fat areas (R = −0.39, p = 0.044), visceral/total fat areas (R = −0.52, p = 0.0043) and waist circumferences (R = −0.43, p = 0.019). Robust regression revealed that TMN (R = 0.81, p<0.0001) and waist circumferences (R = −0.009, p = 0.009) were the two independent predictors of BAT activities.

Conclusions/Significance

Brown adipose tissue activity in adult human can be activated by elevated plasma TMN levels, such as in the case of patients with pheochromocytoma, and is negatively associated with central adiposity.  相似文献   

16.
Prevention of tissue damages at the site of Leishmania major inoculation can be achieved if the BALB/c mice are systemically given L. major antigen (LmAg)-loaded bone marrow-derived dendritic cells (DC) that had been exposed to CpG-containing oligodeoxynucleotides (CpG ODN). As previous studies allowed establishing that interleukin-4 (IL-4) is involved in the redirection of the immune response towards a type 1 profile, we were interested in further exploring the role of IL-4. Thus, wild-type (wt) BALB/c mice or DC-specific IL-4 receptor alpha (IL-4Rα)-deficient (CD11ccreIL-4Rα−/lox) BALB/c mice were given either wt or IL-4Rα-deficient LmAg-loaded bone marrow-derived DC exposed or not to CpG ODN prior to inoculation of 2×105 stationary-phase L. major promastigotes into the BALB/c footpad. The results provide evidence that IL4/IL-4Rα-mediated signaling in the vaccinating DC is required to prevent tissue damage at the site of L. major inoculation, as properly conditioned wt DC but not IL-4Rα-deficient DC were able to confer resistance. Furthermore, uncontrolled L. major population size expansion was observed in the footpad and the footpad draining lymph nodes of CD11ccreIL-4Rα−/lox mice immunized with CpG ODN-exposed LmAg-loaded IL-4Rα-deficient DC, indicating the influence of IL-4Rα-mediated signaling in host DC to control parasite replication. In addition, no footpad damage occurred in BALB/c mice that were systemically immunized with LmAg-loaded wt DC doubly exposed to CpG ODN and recombinant IL-4. We discuss these findings and suggest that the IL4/IL4Rα signaling pathway could be a key pathway to trigger when designing vaccines aimed to prevent damaging processes in tissues hosting intracellular microorganisms.  相似文献   

17.
Previous studies have reported the association between multiple genetic variants in the enamel-formation genes and the risk of dental caries with inconsistent results. We performed a systematic literature search of the PubMed, Cochrane Library, HuGE and Google Scholar databases for studies published before March 21, 2020 and conducted meta-, gene-based and gene-cluster analysis on the association between genetic variants in the enamel-formation genes and the risk of dental caries. We identified 21 relevant publications including a total of 24 studies for analysis. The genetic variant rs17878486 in AMELX was significantly associated with dental caries risk (OR = 1.40, 95% CI: 1.02–1.93, P = 0.037). We found no significant association between the risk of dental caries with rs12640848 in ENAM (OR = 1.15, 95% CI: 0.88–1.52, P = 0.310), rs1784418 in MMP20 (OR = 1.07, 95% CI: 0.76–1.49, P = 0.702) and rs3796704 in ENAM (OR = 1.06, 95% CI: 0.96–1.17, P = 0.228). Gene-based analysis indicated that multiple genetic variants in AMELX showed joint association with the risk of dental caries (6 variants; P < 10−5), so did genetic variants in MMP13 (3 variants; P = 0.004), MMP2 (3 variants; P < 10−5), MMP20 (2 variants; P < 10−5) and MMP3 (2 variants; P < 10−5). The gene-cluster analysis indicated a significant association between the genetic variants in this enamel-formation gene cluster and the risk of dental caries (P < 10−5). The present meta-analysis revealed that genetic variant rs17878486 in AMELX was associated with dental caries, and multiple genetic variants in the enamel-formation genes jointly contributed to the risk of dental caries, supporting the role of genetic variants in the enamel-formation genes in the etiology of dental caries.  相似文献   

18.
The persistence of a reservoir of latently infected CD4 T cells remains one of the major obstacles to cure HIV. Numerous strategies are being explored to eliminate this reservoir. To translate these efforts into clinical trials, there is a strong need for validated biomarkers that can monitor the reservoir over time in vivo. A comprehensive study was designed to evaluate and compare potential HIV-1 reservoir biomarkers. A cohort of 25 patients, treated with suppressive antiretroviral therapy was sampled at three time points, with median of 2.5 years (IQR: 2.4–2.6) between time point 1 and 2; and median of 31 days (IQR: 28–36) between time point 2 and 3. Patients were median of 6 years (IQR: 3–12) on ART, and plasma viral load (<50 copies/ml) was suppressed for median of 4 years (IQR: 2–8). Total HIV-1 DNA, unspliced (us) and multiply spliced HIV-1 RNA, and 2LTR circles were quantified by digital PCR in peripheral blood, at 3 time points. At the second time point, a viral outgrowth assay (VOA) was performed, and integrated HIV-1 DNA and relative mRNA expression levels of HIV-1 restriction factors were quantified. No significant change was found for long- and short-term dynamics of all HIV-1 markers tested in peripheral blood. Integrated HIV-1 DNA was associated with total HIV-1 DNA (p<0.001, R² = 0.85), us HIV-1 RNA (p = 0.029, R² = 0.40), and VOA (p = 0.041, R2 = 0.44). Replication-competent virus was detected in 80% of patients by the VOA and it correlated with total HIV-1 DNA (p = 0.039, R² = 0.54). The mean quantification difference between Alu-PCR and VOA was 2.88 log10, and 2.23 log10 between total HIV-1 DNA and VOA. The levels of usHIV-1 RNA were inversely correlated with mRNA levels of several HIV-1 restriction factors (TRIM5α, SAMHD1, MX2, SLFN11, pSIP1). Our study reveals important correlations between the viral outgrowth and total and integrated HIV-1 DNA measures, suggesting that the total pool of HIV-1 DNA may predict the size of the replication-competent virus in ART suppressed patients.  相似文献   

19.
LigD 3′-phosphoesterase (PE) enzymes perform end-healing reactions at DNA breaks. Here we characterize the 3′-ribonucleoside-resecting activity of Candidatus Korarchaeum PE. CkoPE prefers a single-stranded substrate versus a primer–template. Activity is abolished by vanadate (10 mM), but is less sensitive to phosphate (IC50 50 mM) or chloride (IC50 150 mM). The metal requirement is satisfied by manganese, cobalt, copper or cadmium, but not magnesium, calcium, nickel or zinc. Insights to CkoPE metal specificity were gained by solving new 1.5 Å crystal structures of CkoPE in complexes with Co2+ and Zn2+. His9, His15 and Asp17 coordinate cobalt in an octahedral complex that includes a phosphate anion, which is in turn coordinated by Arg19 and His51. The cobalt and phosphate positions and the atomic contacts in the active site are virtually identical to those in the CkoPE·Mn2+ structure. By contrast, Zn2+ binds in the active site in a tetrahedral complex, wherein the position, orientation and atomic contacts of the phosphate are shifted and its interaction with His51 is lost. We conclude that: (i) PE selectively binds to ‘soft’ metals in either productive or non-productive modes and (ii) PE catalysis depends acutely on proper metal and scissile phosphate geometry.  相似文献   

20.
During the biosynthesis of heme d1, the essential cofactor of cytochrome cd1 nitrite reductase, the NirE protein catalyzes the methylation of uroporphyrinogen III to precorrin-2 using S-adenosyl-l-methionine (SAM) as the methyl group donor. The crystal structure of Pseudomonas aeruginosa NirE in complex with its substrate uroporphyrinogen III and the reaction by-product S-adenosyl-l-homocysteine (SAH) was solved to 2.0 Å resolution. This represents the first enzyme-substrate complex structure for a SAM-dependent uroporphyrinogen III methyltransferase. The large substrate binds on top of the SAH in a “puckered” conformation in which the two pyrrole rings facing each other point into the same direction either upward or downward. Three arginine residues, a histidine, and a methionine are involved in the coordination of uroporphyrinogen III. Through site-directed mutagenesis of the nirE gene and biochemical characterization of the corresponding NirE variants the amino acid residues Arg-111, Glu-114, and Arg-149 were identified to be involved in NirE catalysis. Based on our structural and biochemical findings, we propose a potential catalytic mechanism for NirE in which the methyl transfer reaction is initiated by an arginine catalyzed proton abstraction from the C-20 position of the substrate.  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